Lectures On Electromagnetism PDF
Lectures On Electromagnetism PDF
Lectures On Electromagnetism PDF
Electromagnetism
Second Edition
This page intentionally left blank
Lectures on
Electromagnetism
Second Edition
Ashok Das
University of Rochester, USA
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TA I P E I • CHENNAI
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
LECTURES ON ELECTROMAGNETISM
Second Edition
Copyright © 2013 Hindustan Book Agency (HBA)
Authorized edition by World Scientific Publishing Co. Pte. Ltd. for exclusive distribution worldwide except
India.
The distribution rights for print copies of the book for India remain with Hindustan Book Agency (HBA).
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means, electronic or
mechanical, including photocopying, recording or any information storage and retrieval system now known or to
be invented, without written permission from the Publisher.
ISBN 978-981-4508-26-1
1 Electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Coulomb’s law . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Electric field . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Gauss’ law . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Potential . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5 Electrostatic energy . . . . . . . . . . . . . . . . . . . 24
1.6 Selected problems . . . . . . . . . . . . . . . . . . . . 27
2 Potential for simple systems. . . . . . . . . . . . . . . . . . 31
2.1 Potential for a thin spherical shell . . . . . . . . . . . 33
2.2 Potential for an infinitely long wire . . . . . . . . . . 36
2.3 Potential for a circular charged disc . . . . . . . . . . 39
2.4 Potential for a charge displaced along the z-axis. . . 42
2.5 Dipole . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.6 Continuous distribution of dipoles . . . . . . . . . . . 50
2.7 Quadrupole . . . . . . . . . . . . . . . . . . . . . . . 53
2.8 Potential due to a double layer of charges . . . . . . 55
2.9 Conductors and insulators . . . . . . . . . . . . . . . 58
2.10 Capacitor . . . . . . . . . . . . . . . . . . . . . . . . 63
2.11 Selected problems . . . . . . . . . . . . . . . . . . . . 67
3 Boundary value problems . . . . . . . . . . . . . . . . . . . 69
3.1 Method of images . . . . . . . . . . . . . . . . . . . . 69
3.2 Boundary conditions for differential equations . . . . 80
3.2.1 Partial differential equations . . . . . . . . . . 82
3.2.2 Uniqueness theorem . . . . . . . . . . . . . . . 88
3.3 Solutions of the Laplace equation . . . . . . . . . . . 90
3.3.1 General properties of harmonic functions . . . 90
3.3.2 Solution in Cartesian coordinates . . . . . . . . 93
3.3.3 Solution in spherical coordinates . . . . . . . . 96
3.3.4 Circular harmonics . . . . . . . . . . . . . . . . 103
3.4 Solution of the Poisson equation. . . . . . . . . . . . 106
3.4.1 Green’s function . . . . . . . . . . . . . . . . . 107
vii
viii
xi
xii
Electrostatics
It was observed quite early that when particles carrying electric charge
are brought closer, they experience a force, and this force was called
the electric force. The main question that one studies in electrostat-
ics is the analysis of the electric force experienced by a given charge
due to a complicated distribution of static electric charges in space.
(As a side remark, let us simply note here that although at the mo-
ment it may seem like the course is only concerned with developing
techniques for solving problems in electromagnetism, the techniques
are quite general and are so powerful that they may be used in any
other field of research as well. In that sense, we can think of the
material of the course as developing powerful techniques for solving
theoretical problems through examples of electromagnetism.) Fun-
damental to this study, therefore, is the understanding of the force
between a pair of static charges separated by a given distance. This
question was, in fact, studied by Coulomb in a series of impressive
experiments and he found that the electric force between a pair of
static particles carrying electric charge
1
2 1 Electrostatics
3. and is a vector along the line joining the positions of the two
particles carrying charge. It is attractive if the charges of the
two particles have opposite sign and is repulsive otherwise.
Here, we have defined R = r − r1 (see Fig. 1.1) and have used the
notation (which we will use throughout the lectures) that a bold-
face quantity represents a vector while a boldface quantity with a
“hat” simply stands for a unit vector and, in the above equation, k
represents the constant of proportionality.
z q1
R
r1 q
r
y
x
There are several things to note about the force from the ex-
pression in (1.1). First, it is translationally invariant, namely, under
a translation of the coordinate system, the expression is unchanged.
Sometimes, one takes advantage of this to translate the coordinate
system such that q1 is at the origin, in which case, the force has the
simple form
qq1 qq1
F(r) = k 3
r = k 2 r̂, (1.2)
|r| |r|
N − m2
k = 10−7 c2
C2
105 × 104 dyne − cm2
= 10−7 c2 ×
α−2 esu2
= 102 c2 α2 = 1, (1.4)
From the form of the force in (1.5) it is clear that, even in the presence
of a distribution of static charges, the force is linearly proportional to
the charge of the test particle. Therefore, by dividing out the charge
of the test particle, we can define an auxiliary quantity which we call
the electric field. For a single source charge q1 at r1 , the electric field
at the coordinate r is given by (k = 1 in our units)
q1
E(r) = (r − r1 ), (1.6)
|r − r1 |3
for any well behaved function f (r). The second relation in (1.13), in
fact, follows from the first if we choose f (r) = 1. The two relations
imply that the delta function must vanish at points where its argu-
ment does not vanish and that at points where its argument vanishes,
it must diverge (see Fig. 1.2) in such a way that its integral is unity
(namely, the area under the curve is normalized).
This does not correspond to the behavior of any simple function
that we know of. In fact, the Dirac delta function is truly not a func-
tion, rather it can be thought of as a limit of a sequence of functions.
Without going into detail, let us note some explicit representations
for the delta function. In one dimension, for example, we can write
1 sin g(x − x′ )
δ(x − x′ ) = lim
g→∞ π (x − x′ )
Zg Z∞
1 ik(x−x′ ) dk ik(x−x′ )
= lim dk e = e ,
g→∞ 2π 2π
−g −∞
r
α −α(x−x′ )2
δ(x − x′ ) = lim e ,
α→∞ π
8 1 Electrostatics
r′ r
dθ(x − x′ )
δ(x − x′ ) = , (1.14)
dx
where θ(x − x′ ) is the step function defined to be
(
′
1, for x − x′ > 0,
θ(x − x ) = (1.15)
0, for x − x′ < 0.
ρ(r) = q1 δ3 (r − r1 ). (1.17)
It follows then, from the defining relation in (1.11), that the electric
field produced at the coordinate r by a point charge q1 located at the
point r1 is given by
Z
ρ(r′ )
E(r) = d3 r ′ (r − r′ )
|r − r′ |3
Z
q1
= d3 r ′ δ3 (r′ − r1 ) (r − r′ )
|r − r′ |3
q1
= (r − r1 ), (1.18)
|r − r1 |3
z ds
r
r′
y
x
electric field to be
Z Z Z
ρ(r′ )
ds · E(r) = d3 r ′ ds · (r − r′ ) , (1.19)
|r − r′ |3
S S
ds = dΩ |r − r′ |(r − r′ ), (1.20)
ds · (r − r′ ) = dΩ |r − r′ |3 . (1.21)
Let us add here some clarification on the expression for the sur-
face element in (1.20). First, consider an infinitesimal line element
vector in polar coordinates (see Fig. 1.4) which has the form
dℓ
r
θ dθ
x
dℓ = r dθ θ̂. (1.22)
θ
rd
r r sin θdφ
y
φ
x
Z Zπ Z2π
dΩ = dθ sin θ dφ = 4π, (1.24)
0 0
showing that the total solid angle around a point is 4π. An alternative
way to understand the surface element is to note that the line element
vector in spherical coordinates has the form
which is the result obtained earlier in (1.23). Using (1.21), the surface
integral on the right hand side of (1.19) simplifies and we have
Z Z Z
ds · E(r) = d3 r ′ ρ(r′ ) dΩ
S
Z
= 4π d3 r ′ ρ(r′ ) = 4π Qenclosed , (1.27)
12 1 Electrostatics
where we have used (1.24) (namely, the fact that the total solid angle
around r′ is 4π) and Qenclosed represents the total charge enclosed
inside the volume bounded by the surface. Note that if there are
electric charges outside the enclosing surface, then the flux of the
electric field, due to such charges, enters as well as exits the surface
the same number of times (can be more than once depending on the
topology of the volume), thereby canceling (since the directions of
the surfaces at the point of entry and exit are opposite) any further
contribution to the total flux due to charges external to the given
volume. Thus, independent of the shape of the enclosing surface, we
have the general result
Z
ds · E(r) = 4π Qenclosed . (1.28)
S
This is known as Gauss’ law which says that the total electric
flux out of a closed surface equals 4π times the total electric charge
enclosed in the volume bounded by the surface. In fact, (1.28) repre-
sents the integral form of Gauss’ law. We can also write a differential
form for Gauss’ law by appealing to Gauss’ theorem which says that,
for any vector function A(r),
Z Z
d3 r ∇ · A(r) = ds · A(r), (1.29)
V S
which also differs from the usual differential form of Gauss’ law. ◭
r
h
is no flux through the top or the bottom of the cylinder because of the radial
nature of the electric field),
Z
ds · E(r) = 4π Q,
S
Exercise. Compare this with the behavior of the electric field due to a point charge.
Z
ds · E(r) = 4π Q,
S
Z
or, r 2 dΩ |E(r)| = 4π Q,
Q
or, |E(r)| = , (1.36)
r2
where, Q = 4πR2 σ is the total charge carried by the spherical shell. Furthermore,
recalling that the electric field can only point radially, for r > R, we have
Q 4πR2 σ
E(r) = 2
r̂ = r̂. (1.37)
r r2
Namely, outside the spherical shell, the electric field behaves as if the entire charge
on the surface of the shell were located at the center of the shell.
On the other hand, for points inside the shell, if we apply Gauss’ law and
calculate the electric flux through a spherical shell of radius r < R (see Fig. 1.8),
we obtain,
Z
ds · E(r) = 0, (1.38)
S
since there is no charge inside the Gaussian surface and this leads to the fact that
inside the shell the electric field vanishes. Therefore, we determine the electric
field to have the general form
Q 4πR2 σ
E(r) = θ(r − R) r̂ = θ(r − R) r̂, (1.39)
r2 r2
which shows explicitly that the electric field is discontinuous across the charged
surface. This is, of course, reminiscent of the behavior of the gravitational field
which satisfies the inverse square law as well. Let us note from (1.39) that the
16 1 Electrostatics
discontinuity in the normal component of the electric field across the surface car-
rying charge is given by (R and L refer respectively to right and left of the shell,
or even more appropriate outside and inside the shell)
Q
r̂ · (ER − EL ) = En,R − En,L = 2 = 4πσ. (1.40)
R R R
◭
Z
ds · E = 4πQ,
S
where n̂ represents the outward unit vector normal to the plane. We see that, in
this case, the magnitude of the electric field is constant at every point in space
(although the direction changes on either side of the plane). We note again that
1.4 Potential 17
the normal component of the electric field is discontinuous across the surface
carrying charge with the discontinuity given by
n̂ · (ER − EL ) = En,R − En,L = 4πσ, (1.43)
where the restriction refers to the location of the plane and (1.43) can be compared
with (1.40). ◭
1.4 Potential
We have seen that the electric field produced by a static charge dis-
tribution is a vector given by (1.11). Let us discuss briefly the nature
of this vector for a static distribution of charges. The analysis of
this section obviously may not hold when the charge distribution is
non-static and we will come back to this question later. For the
time being, however, we are interested in solving problems only in
electrostatics and, therefore, this is quite meaningful as we will see
shortly.
18 1 Electrostatics
∇ × r = 0. (1.48)
so that the value of the line integral of the electric field between any
two points depends only on the end points independent of the path
of integration.
It is well known that any vector, which falls off rapidly at infinite,
can be decomposed uniquely into the sum of two vectors, one of which
is divergence free while the other has vanishing curl. This is com-
monly known as the Helmholtz theorem. Explicitly, the Helmholtz
theorem says that, if A is a vector function, we can write it as
A = B + C, (1.53)
where
∇ · B = 0 = ∇ × C. (1.54)
and the form of the scalar function follows from the general form of
an arbitrary vector in (1.55) (from the Helmholtz theorem) to be
Z ′ ′ Z ′
1 3 ′ ∇ · E(r ) 3 ′ ρ(r )
Φ(r) = d r = d r . (1.60)
4π |r − r′ | |r − r′ |
Here, we have used the differential form of Gauss’ law (1.30) in the
last step. Let us note that we could have derived the result in (1.60)
1.4 Potential 21
also directly from the definitions of the electric field in (1.49) and
(1.59), namely,
Z
ρ(r′ )
E(r) = −∇ d3 r ′ = −∇Φ(r),
|r − r′ |
Z
ρ(r′ )
or, Φ(r) = d3 r ′ , (1.61)
|r − r′ |
where we have ignored a constant of integration (which is related to
the choice of a reference point, as we will see). Alternatively, from
(1.59) we can also write
Zr
Φ(r) = − dℓ · E. (1.62)
∞
∇ × E(r) = 0, (1.66)
namely, the electric field (1.65) is curl free. Second, we obtain the divergence of
the electric field to be
r
∇ · E(r) = q∇ · a
r
1 1
=q ∇ · r + ∇ · r
ra ra
a 3 (3 − a)q
= q − a+2 r · r + a = . (1.67)
r r ra
Comparing this with the differential form of Gauss’ law (1.30), we identify the
charge density that produces such an electric field to be
(3 − a)q
ρ(r) = . (1.68)
4πr a
We can use (1.62) to calculate the potential as
Zr
Φ(r) = − dℓ · E
∞
Zr
dr ′ qr 2−a
= −q =− , (1.69)
(r ′ )a−1 2−a
∞
where we have used the fact that the electric field is conservative (curl free) and
correspondingly chosen a radial path to do the line integral. (Recall that the
electric field is along the radial direction. We have also thrown away a divergent
constant for the case a < 2.) Note that when a = 2, the above expression needs
to be calculated in a limiting manner and gives
∇ × E(r) = 0, (1.72)
1.4 Potential 23
qr 2−a
Φ(r) = − , (1.73)
2−a
as derived earlier. Furthermore, the divergence of the electric field is given by
2 q 2 2−a q 1 ∂ 2 ∂
∇ · E(r) = −∇ Φ(r) = ∇ (r )= r r 2−a
2−a 2 − a r 2 ∂r ∂r
q 1 ∂ (3 − a)q
= (2 − a)r 3−a = , (1.74)
2 − a r 2 ∂r ra
which coincides with the earlier result and leads to the charge density through
Gauss’ law. ◭
∂ e−µr
E(r) = −∇Φ(r) = −qr̂
∂r r
qr̂
= (1 + µr) e−µr . (1.76)
r2
We note that when µ = 0, this reduces to the Coulomb field (1.6) for a point
charge at the origin.
We can also calculate the charge density that produces this potential by
using the differential form of Gauss’ law. Namely, we note that
r
∇ · E(r) = q∇ · (1 + µr) e−µr
r3
h r r i
= q ∇ · 3 (1 + µr) e−µr + 3 · ∇ (1 + µr) e−µr
r r
1 r
= q − ∇2 (1 + µr) e−µr + 3 · −µ2 r e−µr
r r
2
µ −µr
= q 4πδ 3 (r) (1 + µr) e−µr − e
r
µ2 −µr
= q 4πδ 3 (r) − e = 4πρ(r). (1.77)
r
Here we have used (1.56) in the intermediate step. This determines the charge
density associated with the Yukawa potential to be
qµ2 −µr
ρ(r) = q δ 3 (r) − e . (1.78)
4πr
We note that for µ = 0, this reduces to (1.17). ◭
24 1 Electrostatics
Let us next calculate the electrostatic potential energy for a given dis-
tribution of charges. As we will see, the form of the result is different
1.5 Electrostatic energy 25
Zr2
(r − r1 )
= −q1 q2 dr ·
|r − r1 |3
∞
2 −r1
rZ
r̂
= −q1 q2 dr ·
|r|2
∞
|rZ
1 −r2 |
dr
= −q1 q2
r2
∞
q1 q2
= , (1.82)
|r1 − r2 |
where |r| = r and, since the integral is independent of the path, owing
to the conservative nature of the electric field, we have chosen a radial
path in evaluating the integral. We can now keep adding more and
more charges and since the electric force is additive, the calculation
simplifies. For example, to bring in a third charge q3 to the point r3 ,
the total work done is given by
Zr3
W123 = W12 − dℓ · (F13 + F23 )
∞
q1 q2 q1 q3 q2 q3
= + + . (1.83)
|r1 − r2 | |r1 − r3 | |r2 − r3 |
It is obvious that this expression is completely symmetric in the per-
mutation of any pair of charged particles (charges and coordinates)
26 1 Electrostatics
and hence, the order in which the charges are brought in does not
matter. Carrying this out for n charged particles, it is easily obtained
that the total work required is
1 X qi qj
W = , (1.84)
2 |ri − rj |
i,j,i6=j
where we have used the differential form of Gauss’ law in (1.63) (Pois-
son equation) as well as the relation between the electric field and the
potential. We have also neglected surface terms in the integration by
parts with the assumption that the electric field falls off rapidly at
infinity.
The difference between the two cases needs to be discussed.
First, we note that the energy for a continuous distribution of charges
is completely given in terms of the electric field and is non-negative.
On the other hand, for a distribution of point charges the sign of the
energy in (1.84) clearly depends on the signs of various charges and
can, in fact, be negative. Let us recall that the electrostatic energy
for a pair of similarly charged particles is positive while it is neg-
ative if the charges of the two particles are opposite in sign. This
difference in the behavior of the electrostatic energy between a dis-
crete and a continuous distribution of charges arises mainly because
of our choice of the reference system in the case of point charges.
We assumed the zero of the energy to correspond to the system of
1.6 Selected problems 27
and
∇ × (∇ × A) = ∇(∇ · A) − ∇2 A,
28 1 Electrostatics
∇ · (A × B) = B · (∇ × A) − A · (∇ × B),
∇ · r = 3, ∇ × r = 0,
1
∇2 = −4πδ3 (r).
|r|
3. Using the result from the last problem show that for an arbi-
trary vector A we have
Z Z
3
d r ∇ × A = ds × A,
V S
e−µr
Φ(r) = q ,
r
where r = |r| and µ is a mass parameter (in units of c, ~) is
produced by a point charge q at the origin. Would Gauss’ law
be valid for such a case? Show that, for r 6= 0, this potential
satisfies the equation
∇2 Φ(r) = µ2 Φ(r).
1.6 Selected problems 29
ρ(r) = k r −n ,
Let us next calculate the potential function for some simple systems
with known charge distributions, to get a feeling for its properties.
The simplest example, of course, is the potential for a point charge.
Let us assume that a charge q1 is located at the coordinate r1 . In
such a case, we can write the charge density as (see (1.17))
ρ(r) = q1 δ3 (r − r1 ), (2.1)
31
32 2 Potential for simple systems
which is to be expected since the charge density in (2.3) is simply q times the
(quantum mechanical) probability density in the ground state.
The calculation of the potential and the electric field can be carried out
in one of two ways. First, let us note that the charge distribution is spherically
symmetric. Consequently, we expect the electric field as well as the potential
to reflect this. Using Gauss’ law to integrate the electric field over a spherical
(Gaussian) surface of radius r we obtain
Z Z
ds · E = 4π d3 r ′ ρ(r′ )
Z Zr
2 q − 2ra′
or, 4πr |E(r)| = 4π dΩ r ′2 dr ′ e
πa3
0
Zr
q 2r′
or, |E(r)| = (4π) dr ′ r ′2 e− a
πa3 r 2
0
2r
Z
a
q ′
= dr ′ r ′2 e−r
2r 2
0
q h ′
i 2r
a
= 2 (−r ′2 − 2r ′ − 2)e−r
2r 0
q 2r 2r 2 2r
= 2 1− 1+ + 2 e− a . (2.5)
r a a
Zr
q 2r ′ 2r ′2 − 2r
′
=− dr ′ 1 − 1 + + e a
r ′2 a a2
∞
Zr
d 1 r′ 2r′
=q dr ′ 1− 1+ e− a
dr ′ r′ a
∞
qh r − 2r i
= 1− 1+ e a . (2.7)
r a
The second way of solving the problem is to note that given the charge
density, we can obtain the potential simply as
Z Z 2r′
ρ(r′ ) q e− a
Φ(r) = d3 r ′ = d3 r ′
|r − r′ | πa3 |r − r′ |
2.1 Potential for a thin spherical shell 33
Z∞ Z1
q ′ ′2 − 2r
′
d cos θ
= (2π) dr r e a
1
πa3 (r 2 + r ′2 − 2rr ′ cos θ) 2
0 −1
Z∞
q (−2π) 2r′
= dr ′ r ′ e− a |r − r ′ | − (r + r ′ )
πa3 r
0
Zr Z∞
4q − 2r
′
− 2r
′
= dr ′ r ′2 e a +r dr ′ r ′ e a
ra3
0 r
qh r − 2r i
= 1− 1+ e a . (2.8)
r a
This is the same result as obtained in the earlier method in (2.7). The electric
field now follows from the definition
∂ qh r − 2r i
E(r) = −∇Φ(r) = −r̂ 1− 1+ e a
∂r r a
2
qr̂ 2r 2r 2r
= 2 1− 1+ + 2 e− a , (2.9)
r a a
Z
σ(r′ )
Φ(r) = ds′ . (2.11)
|r − r′ |
34 2 Potential for simple systems
R σ
2πRσ
Φ(r) = (r + R − |r − R|)
r
4πR2 σ = Q for r > R,
r r
= (2.14)
4πR2 σ = Q
for r < R.
R R
There are several things to observe from the result in (2.15) (or
(2.14)). First, the potential is spherically symmetric, as it should be
because of the symmetry in the problem. Second, it is a continuous
function across r = R, namely, across the surface carrying the charge.
This has to be contrasted with the behavior of the electric field. Out-
side the shell, the potential behaves as if all the charge were located
at the origin. We also note that the potential is a constant inside
the shell. In fact, the value of the potential at the origin is readily
seen to be the average of the potential over any closed surface within
the shell. This, as we will see, is a general feature of the solutions
of Laplace equation. Finally, given the potential, we can determine
the electric field by taking the gradient (see (1.59)). Recalling from
(2.15) that the potential only depends on the radial coordinate, we
obtain
E(r) = −∇Φ(r)
∂ Q Q
= −r̂ θ(r − R) + θ(R − r)
∂r r R
Q
= θ(r − R) r̂. (2.16)
r2
The important thing to note in this derivation is that the derivatives
of the two theta functions give delta functions of opposite sign which
cancel each other. Eq. (2.16) is, of course, our previous result ob-
tained in (1.37) and (1.38), namely, the electric field is non-vanishing
only outside the shell and, at such points, it behaves as if the to-
tal charge were concentrated at the origin. Furthermore, expressed
as in (2.16), the discontinuity in the electric field across the surface
carrying charge is manifest.
36 2 Potential for simple systems
Figure 2.2: An infinitely long wire along the z-axis carrying a constant
linear charge density λ.
z
r
φ y
x ρ̃
x = ρ̃ cos φ,
y = ρ̃ sin φ,
z = z, (2.17)
2.2 Potential for an infinitely long wire 37
1
or, c = . (2.20)
2π ρ̃
With (2.19), we can now calculate the potential due to a thin
wire from the definition in (2.1).
Z
ρ(r′ )
Φ(r) = d3 r ′
|r − r′ |
Z
λ δ(ρ̃′ )
= ρ̃′ dρ̃′ dφ′ dz ′
2π ρ̃′ ((ρ̃ − ρ̃′ )2 + (z − z ′ )2 ) 12
Z∞
dz ′
=λ 1
(ρ̃2 + (z − z ′ )2 ) 2
−∞
Z∞
dz̄
=λ 1
(ρ̃2 + z̄ 2 ) 2
−∞
p ∞
= λ ln z̄ + ρ̃2 + z̄ 2 . (2.21)
−∞
38 2 Potential for simple systems
It is clear that the right hand side of (2.21) diverges and the reason
for this is not hard to see. In writing an expression for the poten-
tial, we had chosen the potential to vanish at infinity which we had
taken as a reference point. However, in the present problem, such
a choice is not consistent simply because the charge density extends
to spatial infinity. (We have an infinitely long wire as is clear from
the integration limits.) The proper way to analyze this problem is
to recognize that we must choose a different reference point for this
problem (or equivalently allow for a constant potential at infinity).
In particular, let us note that a different choice of the reference point
simply corresponds to adding a constant term to the potential, be it
infinite. Thus, we can extract the finite meaningful potential from
(2.21) by writing
p Λ
Φ(r) = lim λ ln z̄ + ρ̃2 + z̄ 2
Λ→∞ −Λ
p
ρ̃2 + Λ2 )
(Λ +
= lim λ ln p
Λ→∞ (−Λ + ρ̃2 + Λ2 )
p
(Λ + ρ̃2 + Λ2 )2
= lim λ ln
Λ→∞ ρ̃2
4Λ2 + 2ρ̃2 + O( Λ12 )
≈ lim λ ln
Λ→∞ ρ̃2
= −2λ ln ρ̃ + constant, (2.22)
where the constant, on the right hand side, is a divergent constant.
(The important thing to remember is that the potential is not observ-
able, but the electric field is through the electric force. The electric
field is obtained from the potential by the gradient operation so that
a constant term in the potential does not contribute to the electric
field.) Thus, discarding the constant we determine the potential for
the infinitely long charged wire to be
Φ(r) = −2λ ln ρ̃. (2.23)
We see that the potential has cylindrical symmetry and it is con-
tinuous. The electric field can be obtained from (2.23) (or (2.22))
by taking the gradient and since the potential only depends on the
coordinate ρ̃, we obtain
ˆ ∂ 2λ ˆ
E(r) = −∇Φ(r) = −ρ̃ (−2λ ln ρ̃) = ρ̃, (2.24)
∂ ρ̃ ρ̃
which is the same result that we had obtained earlier in (1.35).
2.3 Potential for a circular charged disc 39
R σ
Note that the theta function implements the finite extension of the
disc (and, therefore, the charge distribution).
We can now calculate the potential at a point on the z-axis
(x = 0 = y) simply as
Z
ρ(r′ )
Φ(x = 0, y = 0, z) = d3 r ′
|r − r′ |
Z
σδ(z ′ )θ(R − r ′ )
= r ′ dr ′ dφ′ dz ′ 1
(r ′2 + (z − z ′ )2 ) 2
ZR
r ′ dr ′
= 2πσ 1
(r ′2 + z 2 ) 2
0
40 2 Potential for simple systems
1 R
= 2πσ r ′2 + z 2 2
0
p
= 2πσ R2 + z 2 − |z| . (2.26)
πR2 σ Q
≈ = , (2.27)
|z| |z|
Furthermore, the electric field along the z-axis can be obtained from
the potential in (2.26) by taking the gradient which gives
E(x = 0, y = 0, z) = −∇Φ(x = 0, y = 0, z)
∂ p
= −ẑ 2πσ R2 + z 2 − |z|
∂z
z
= −2πσ √ − sgn(z) ẑ. (2.29)
R2 + z 2
Here, sgn (z) stands for the sign of z which can also be represented
as an alternating step function. It is obvious from (2.29) that very
close to the disc, namely, for z ≈ 0, the electric field has the leading
behavior
First of all, this explicitly shows that the electric field is discontinuous
across the surface. But, more interestingly, the form of the electric
field, at such points, is like the electric field for the infinite plane
which we have calculated earlier (see (1.42)). This can be understood
qualitatively as a consequence of the fact that very close to the disc,
2.3 Potential for a circular charged disc 41
σ y
x
Φ(x = R, y = 0, z = 0)
Z
ρ(r′ )
= d3 r ′
|r − r′ |
Z
σδ(z ′ )θ(R − r ′ )
= r ′ dr ′ dφ′ dz ′ 1
((R − r ′ cos φ′ )2 + r ′2 sin2 φ′ + z ′2 ) 2
ZR Z2π
r ′ dr ′ dφ′
=σ 1
(r ′2 − 2Rr ′ cos φ′ + R2 ) 2
0 0
Z2π p
′
=σ dφ r ′2 − 2Rr ′ cos φ′ + R2
0
p R
+ R cos φ′ ln 2 r ′2 − 2Rr ′ cos φ′ + R2 + 2r ′ − 2R cos φ′
0
42 2 Potential for simple systems
Z2π
′
p
= σR dφ 2(1 − cos φ′ ) − 1
0
p
′ 2(1 − cos φ′ ) + (1 − cos φ′ )
+ cos φ ln
(1 − cos φ′ )
Z2π !
φ′ 1
= σR dφ′ 2 sin − 1 + cos φ′ ln 1 + ′ . (2.31)
2 sin φ2
0
Here, in the intermediate steps, we have used some standard integrals
from the tables (for example, see Gradshteyn and Ryzhik, 2.261 and
2.264). Finally, integrating by parts the last term inside the bracket
in (2.31), we have (the first term in the integration by parts vanishes
at the limits)
! Z2π 2
Z2π φ′
1 cos 2
dφ′ cos φ′ ln 1 + φ′
= dφ′
sin 2 φ′
0 0
1 + sin 2
Z2π
′ φ′
= dφ 1 − sin . (2.32)
2
0
Using this in (2.31), we obtain
Z2π
′ φ′ φ′
Φ(x = R, y = 0, z = 0) = σR dφ 2 sin − 1 + 1 − sin
2 2
0
Z2π
φ′
= σR dφ′ sin
2
0
= 4σR. (2.33)
It is interesting to compare this with the value of the potential at
the center of the disc (2.28), which makes it clear that the potential
decreases as we move out of the center of the disc. Consequently,
there must exist a radial component of the electric field on the disc
itself.
The next example that we consider is really not that different from
what we have studied earlier and yet has many features which will be
2.4 Potential for a charge displaced along the z-axis 43
ẑ
r−R
q, R r b
y
x
Figure 2.6: A point charge q displaced from the origin along the z-axis
by a distance R.
series as
1
1
(r 2 + R2 − 2Rr cos θ) 2
2 !− 21
1 2R R
= 1− cos θ +
r r r
" 2 !
1 1 2R R
= 1− − cos θ +
r 2 r r
2 !2
3 2R R
+ − cos θ + + ···
8 r r
1 R cos θ R2 (3 cos2 θ − 1)
= 1+ + + ···
r r 2r 2
1 R cos θ R2 (3 cos2 θ − 1)
= + + + ··· . (2.36)
r r2 2r 3
Substituting this back into the potential in (2.35), we obtain (for
r >> R),
q qR cos θ qR2 (3 cos2 θ − 1)
Φ(r) = + + + ··· . (2.37)
r r2 2r 3
As we will see later, the angular coefficients of the expansion
of the denominator in (2.36) can be identified with the Legendre
polynomials, namely,
P0 (cos θ) = 1,
P1 (cos θ) = cos θ,
(3 cos2 θ − 1)
P2 (cos θ) = , (2.38)
2
and so on. Thus, very far away from the point charge, we can write
the potential due to a charge displaced along the z-axis, as
X∞
qRn
Φ(r) = Pn (cos θ). (2.39)
n=0
r n+1
This also follows from the fact that the expression for the potential
in (2.35) is symmetric under r ↔ R.)
Such an expansion of the potential is known as the multipole
expansion. We note that, very far away from the charge, the domi-
nant term is the first term which we recognize to be the potential due
to a point charge at the origin (also called a monopole term). How-
ever, if for some reason, the first term is absent (namely, if we have
a charge neutral system), then the dominant term will be the second
term which is the potential due to a dipole. Furthermore, if we have
a system for which the first two terms vanish, then the leading term
would be the third term which is the potential due to a quadrupole
and so on.
As a parenthetical discussion, let us analyze the expansion of
the denominator in (2.35) a bit more in detail. Let us recall that a
d
translation by an amount a in one dimension is implemented by ea dx
so that we can write
d
f (x + a) = ea dx f (x) . (2.41)
The exponential operator simply generates the Taylor series for the
expansion of the function. In higher dimensions, this generalizes so
that we can write
f (r + a) = ea·∇ f (r) . (2.42)
−R·∇ 1 −R·∇ 1
= e = e
|r| r
∂ sin θ ∂ 1
= e−R(cos θ ∂r − r ∂θ ) , (2.43)
r
which, in fact, generates the series in (2.36). The other thing to
observe is that, for r ≫ R, we can write the denominator in (2.35) as
1 1 1
1 = , (2.44)
(r 2 + R2 − 2Rr cos θ) 2 r (1 − 2z cos θ + z 2 ) 21
2.5 Dipole
q, R b
r
y
b
−q, −R
x
Figure 2.7: A dipole system with two point charges where charge q
is at z = R while charge (−q) is at z = −R.
The potential for this system can be simply obtained from what we
have already calculated (see (2.39) or (2.40)), namely, for r >> R,
we have (the potential due to the second charge, (−q) at z = −R, is
obtained by letting q → −q and θ → π − θ)
q qR cos θ qR2 (3 cos2 θ − 1)
Φ(r) = + + + ···
r r2 2r 3
q qR cos θ qR2 (3 cos2 θ − 1)
+ − + − + ···
r r2 2r 3
2qR cos θ
= + ··· . (2.47)
r2
Here, we note that the potential at very large distances does not
behave like that of a point charge. In fact, the total charge of the
system is zero as seen from large distances. Therefore, it is the second
2.5 Dipole 47
where d represents the vector from the negative charge to the positive
charge. Incidentally, a more complete definition of the electric dipole
moment for a continuous distribution of charges is given by
Z
p = d3 r r ρ(r), (2.49)
p · r̂ p·r
Φdipole (r) = 2
= 3
r r
1 p
= −p · ∇ = −∇ · , (2.50)
|r| |r|
which shows that the potential for the dipole can, in fact, be written
as a divergence (p is constant and hence can be taken inside the
gradient operation). Thus, comparing with the potential for a point
charge, we realize that the potential for the dipole behaves more
like the electric field of a charge (just the gradient nature or the
dependence on the distance and not the vector aspect). We can
also calculate the electric field associated with the dipole system by
recalling that in spherical coordinates,
∂ θ̂ ∂ φ̂ ∂
∇ = r̂ + + , (2.51)
∂r r ∂θ r sin θ ∂φ
so that the electric field for a dipole has the form
r
θ θ̂
y
x
q
ℓ
(−q)
r
In this case, the negative and the positive charges of the dipole will experi-
ence an electric force given respectively by
where ℓ denotes the vector from the negative charge to the positive charge. As a
result, the total force acting on the dipole can be written as
Here we have used the definition of the dipole moment in (2.48) to identify p =
qℓ. We note that the dipole experiences an electrostatic force in the presence of
a nonuniform electric field. On the other hand, if the electric field is uniform
(constant), then it is clear that the net force on the dipole vanishes, namely, the
positive and the negative charges experience equal and opposite forces.
We can also calculate the torque exerted on the dipole due to such a force
given in (2.55). The torque around the origin, for example, will have the form
τ = r × F(−q) + (r + ℓ) × F(q)
We note that when the electric field is uniform, the second term on the right hand
side vanishes, but the first term is nonzero. Therefore, the dipole experiences a
torque even though the net force acting on the dipole is zero in this case. ◭
50 2 Potential for simple systems
It is worth pointing out here that there are many physical systems in
nature that behave like a dipole. In many molecules, for example, the
centers of positive and negative charges may not coincide giving rise
to an associated dipole moment even though the molecule as a whole
is charge neutral. A prime example of this is the water molecule
which behaves like a strong dipolar molecule and for these reasons,
the study of dipoles is quite significant. In fact, just as we can define
a continuous distribution of charges, for such dipolar material we can
also define a continuous distribution of dipole moments. Let P(r)
represent the dipole moment density centered at r so that we can
write the total dipole moment associated with such a physical system
as
Z
P = d3 r P(r), (2.57)
V
Here, in the last step we have used the fact that P(r′ ) does not
depend on the coordinate r and hence the gradient operator does not
act on it and can be taken outside the integral. Thus, we see that
the potential for a continuous distribution of dipole moments can be
written as a divergence as well.
Given a potential we can always relate it to a charge distribu-
tion through the Poisson equation. Therefore, it is meaningful to
2.6 Continuous distribution of dipoles 51
Z
3 ′ ′ ′ 1
= d r P(r ) · ∇
|r − r′ |
V
Z Z
3 ′ ′ P(r′ ) ∇′ · P(r′ )
= d r ∇ · − d3 r ′
|r − r′ | |r − r′ |
V V
Z Z
ds′ · P(r′ ) ∇′ · P(r′ )
= − d3 r ′ . (2.60)
|r − r′ | |r − r′ |
S V
σ(r) = n̂ · P(r),
ρ(r) = −∇ · P(r), (2.61)
The integral can be evaluated in the following way. (We use the standard trick
to simplify the evaluation of this integral, namely, let us assume that r lies along
the z-axis, or alternatively that the angle θ′ is measured from r.)
Z
θ(R − r ′ )
d3 r ′
|r − r′ |
Z
θ(R − r ′ )
= r ′2 dr ′ sin θ′ dθ′ dφ′ 1
(r 2 + r ′2 − 2rr ′ cos θ′ ) 2
ZR
1 1 π
= 2π dr ′ r ′2 (r 2 + r ′2 − 2rr ′ cos θ′ ) 2
rr ′ 0
0
ZR
2π
= dr ′ r ′ (r + r ′ ) − |r − r ′ | . (2.63)
r
0
Z ZR
θ(R − r ′ ) 2π
d3 r ′ = dr ′ r ′ (r + r ′ − r + r ′ )
|r − r′ | r
0
ZR
4π 4πR3
= dr ′ r ′2 = , (2.64)
r 3r
0
2.7 Quadrupole 53
r
Z ZR
4π 4π r 3 r(R2 − r 2 )
= dr ′ r ′2 + dr ′ rr ′ = +
r r 3 2
0 r
r 2
4π(3R2 − r 2 ) 4πR3
= = 3− . (2.65)
6 6R R
Therefore, we can write the potential (2.62) for the dipole distribution to
be
r 2
1 1
Φdipole (r) = −P · ∇ θ(r − R) + θ(R − r) 3− , (2.66)
r 2R R
where we have defined the total dipole moment of the sphere (V is the volume of
the sphere) as
4π R3 P
P = PV = . (2.67)
3
Furthermore, since the polarization is along the z-axis, we obtain
Φdipole (r)
r 2
∂ 1 1
= −|P| θ(r − R) + θ(R − r) 3−
∂z r 2R R
hz z z z i
= |P| 3
θ(r − R) − 2 δ(r − R) + 2 δ(r − R) + 3 θ(R − r)
r R R R
P ·r P ·r
= θ(r − R) + θ(R − r). (2.68)
r3 R3
It is clear from the above calculation that the potential is continuous. Outside the
sphere the potential behaves as if we have a single dipole with moment P centered
at the origin while the potential inside the sphere is that of a uniform electric field
(− RP ) (the negative sign is from the definition of the potential (−E · r) for a
3
uniform field, or alternatively, from the definition E = −∇Φ). ◭
2.7 Quadrupole
q, R b
r
−2q b
y
b
q, −R
x
qR2 (3 cos2 θ − 1)
= + ··· . (2.69)
r3
This shows that for this system of charges both the monopole and
the dipole terms in the potential vanish. The total charge of the
system is zero which is why the monopole term in (2.69) vanishes,
but we can also think of the system of four charges in Fig. 2.11 as
two dipoles with opposite dipole moments which makes the dipole
term vanish as well (total dipole moment is zero). As a result it is
the third term in the multipole expansion which gives the leading
behavior of the potential for large distances. Such a configuration
of charges is known as a quadrupole. We note that unlike the case
of dipoles, other configurations of quadrupoles are possible and this
is not the unique quadrupole configuration. Furthermore, while the
dipole moment is a vector, the quadrupole moment, in general, is a
second rank symmetric traceless tensor. In fact, the n-th term in the
multipole expansion (see, for example, (2.39)) leads to the 2n-th pole
moment which isR an n-th rank tensor. The monopole gives a 0-th
rank tensor q = d3 r ρ(r) which is a scalar, the dipole gives a rank
one tensor which is a vector and so on. In Cartesian coordinates the
2.8 Potential due to a double layer of charges 55
with all other components vanishing. (Note that knowing Q11 and
Q22 , we could have predicted the value for Q33 from the tracelessness
condition of the quadrupole moment tensor.) In general, the potential
for the quadrupole can be written in terms of the quadrupole moment
as (for large values of r = |x|)
1X 3xi xj − δij x2
Φquadrupole (r) = Qij , (2.72)
6 r5
i,j
which can be readily checked to give the leading term (the quadrupole
potential) in (2.69) for the system under consideration. We can again
obtain the electric fields associated with the quadrupole system by
taking the gradient. But, we will not get into the details of this
except for noting that the potential as well as the electric field for
the quadrupole decrease even faster than those for the dipole.
z=0
where we recognize that the constant on the right hand side of (2.74)
is a divergent constant. We also understand the origin of this diver-
gent constant, namely, we have a charge distribution which extends
to infinity. Consequently, the reference point for the potential has to
be chosen differently or one has to allow for a (possibly) divergent
constant in the potential. However, an additive constant is not mean-
ingful if we are interested in physical quantities such as the electric
field and, therefore, ignoring the constant we can write the potential
for such a system of charge distribution to be
∂
E(r) = −∇Φ(r) = −ẑ (−2πσ|z|) = 2πσ sgn(z) ẑ, (2.76)
∂z
σ
−σ
z = − d2 z= d
2
d d
Φ(r) = Φ1 (r) + Φ2 (r) = −2πσ z − − z+ . (2.77)
2 2
−− +
+
−
− ++
− +
En,R = 4π σ, (2.80)
E
B
curve cutting the inner surface (see Fig. 2.15), we must have
I ZB ZB
dℓ · E = dℓ · E = dℓ |E| = 0, (2.81)
C
A A
−
++ −
−
−
− q
b
−
+ −
++
r
b
q
2.10 Capacitor
V = Φ1 − Φ2 , (2.83)
Q = C V, (2.84)
geometrical property of the system. We also note here that the MKS
unit of capacitance is a farad (F) which is defined as
(−Q)
e dQ
Q e
e (Φ
dW = dQ e1 − Φ
e 2 ) = Ve dQ
e= . (2.90)
C
Integrating this we can obtain the total work necessary to charge the capacitor
plates starting from the uncharged state and the result is
ZQ e e
Q dQ Q2 (CV )2 1
W = = = = CV 2 . (2.91)
C 2C 2C 2
0
This work must, of course, be stored in the capacitor as electrostatic energy. Now,
let us recall that for the parallel plate system
V = |E|d,
A
C= . (2.92)
4πd
Using these we can also write
1 A 1 2 1 2
W = × (|E|d)2 = E × Ad = E × volume, (2.93)
2 4πd 8π 8π
which is exactly the result we had obtained earlier for the electrostatic energy for
a continuous distribution of charges in the last chapter (see (1.85)). ◭
where
Q1 Q2
σ1 = , σ2 = . (2.95)
4πR12 4πR22
66 2 Potential for simple systems
R2
Q2
R1 Q1
This is quite general so far. Let us next assume that we have a spherical
capacitor system in which case
C = R2 . (2.102)
Let us note here that capacitors are widely used to store charge. ◭
3Q
ρ(r) = (R − r), r ≤ R,
πR4
where the coordinate origin is assumed to coincide with the
center of the sphere. What is the total charge carried by the
sphere? Calculate the electric fields both inside and outside the
sphere.
69
70 3 Boundary value problems
d as shown in Fig. 3.1. Let us assume for concreteness that the charge is positive
although this is not essential for our discussion.
b q
It is clear that the electric field is meaningful and nonzero only above the
plane. This is because, for an insulated conductor, the point charge would induce
charges of opposite sign on the two sides of the conductor in such a way as to
cancel the field within the conductor. The charges on the lower surface of the
conductor, positive charges in this case, would then give rise to an electric field on
the lower half of the plane. Here, however, we have a conductor that is grounded
and the ground has an infinite supply of negative charges which would move on to
the conductor to annihilate all the positive charges. As a result of this, the surface
would have a net negative charge and all the field lines originating from the point
charge q would terminate on the plane and there will be no field lines below the
plane. Namely, the electric field cannot penetrate a grounded conducting plane
of infinite extent. Another way to see this is to note that, by grounding, the
surface of the conductor is maintained at zero potential as is the surface of the
plane infinitely below the conducting plane. Consequently, the potential difference
(voltage) across these two planes is zero and there cannot be an electric field in
this region. It is only at points above the plane that the electric field will be
nonzero.
Therefore, to determine the potential and the electric field above the plane
for this physical system, we need to find a fictitious charge distribution which can
reproduce the boundary condition. Without loss of generality, we can assume that
the conducting plane lies in the x − y plane and that the charge q is located at
a height d on the positive z-axis. Clearly the problem has cylindrical symmetry
and, consequently, it is meaningful to use cylindrical coordinates to analyze this
problem. Let us next assume that a second point charge q ′ is located at a height
d′ below the plane on the negative z-axis (see Fig. 3.2). In the presence of these
two charges, the potential at any point can be easily calculated and in particular
on the plane at z = 0 it is given by
q q′
Φ(r, φ, z = 0) = √ +√ , (3.1)
r2 +d 2 r + d′2
2
3.1 Method of images 71
b q
d′
q′ b
where r represents the radial coordinate on the plane (in cylindrical coordinates).
Requiring the potential to vanish on the plane (which is our boundary condition),
we obtain
q 2 r 2 + d′2 − q ′2 r 2 + d2 = 0,
or, q 2 − q ′2 r 2 + q 2 d′2 − q ′2 d2 = 0. (3.2)
Requiring this to be true for any value of r (namely, at any point on the plane),
we determine
Putting this back into the expression for Φ in (3.1) and noting that both d and d′
are positive, we determine that the potential vanishes on the plane only for
q ′ = −q, d′ = d. (3.4)
q q
Φ(r) = p −p . (3.5)
r 2 + (z − d)2 r 2 + (z + d)2
Consequently, the electric field at any point above the plane (z > 0) is obtained
to be
72 3 Boundary value problems
E(r) = −∇Φ(r)
!
∂ ∂ q q
= − r̂ + ẑ p −p
∂r ∂z r 2 + (z − d)2 r 2 + (z + d)2
!
r r
= qr̂ 3 − 3
(r 2 + (z − d)2 ) 2 (r 2 + (z + d)2 ) 2
!
(z − d) (z + d)
+qẑ 3 − 3 , (3.6)
(r 2 + (z − d)2 ) 2 (r 2 + (z + d)2 ) 2
where we have used the fact that the potential is independent of the azimuthal
angle φ and correspondingly have dropped the angular derivative in the gradient.
r
q b
−q b
Figure 3.3: The potential at a point above the plane due to the charge
q and the “image” charge (−q) at a distance d above and below the
plane respectively.
Let us emphasize that the expressions for both the potential and the electric
field are meaningful only for z > 0. As is obvious, they give the wrong result for
z < 0 as they should because we do not have a real charge for z < 0. It is worth
noting here that for z → 0+ , the electric field (3.6) takes the form
2qd
E(r)|z→0+ = − 3 ẑ. (3.7)
(r 2 + d2 ) 2
That is, there is only a normal component of the electric field on the surface along
the z-axis as we would expect. It is not a uniform electric field on the plane, but
has cylindrical symmetry and, in fact, the field becomes weaker as we move away
from the origin. (By the way, remember that r is the radial coordinate on the
plane.) On the lower surface (z < 0), of course, there is no electric field and,
consequently, the normal component of the electric field is discontinuous across
the surface. And as we have seen (see, for example, (2.10)), the discontinuity is
3.1 Method of images 73
Namely, the presence of the point charge q induces a charge density of opposite
sign on the surface of the grounded conductor. (For our case, with the choice
q > 0, there would be a negative induced surface charge.) The surface charge
density is invariant under rotations on the plane, but is not uniform. In fact, like
the electric field, it falls off rapidly as we move away from the center. We can, of
course, obtain the total induced surface charge by integrating over the entire area
and we have
Z
Qinduced = ds σ(r)
Z
qd r dr dφ
=− 3
2π (r 2 + d2 ) 2
Z∞
r dr
= −qd 3
(r 2 + d2 ) 2
0
∞
1
= qd √ = −q. (3.9)
r 2 + d2 0
Namely, the total induced charge on the plane is equal to the “image” charge
(which, in this case, is equal in magnitude to the physical charge, but opposite in
sign).
The induced charge, of course, would lead to an attractive force between the
point charge and the plane and this can be calculated in the following way. Let
us note that at any point on the z-axis, the electric field produced by the induced
surface charge density on the plane would be along the z-axis by symmetry. We
can calculate it in a simple manner. The potential at any point along the z-axis
due to the surface charges is given by (z > 0)
Z
σ(r ′ )
Φ(x = 0, y = 0, z) = r ′ dr ′ dφ′ 1
(r ′2 + z 2 ) 2
Z
qd r ′ dr ′
=− (2π) 3 1
2π (r + d ) 2 (r ′2 + z 2 ) 2
′2 2
Z∞
q ∂ dx
= 1 1
2 ∂d (x + d2 ) 2 (x + z 2 ) 2
0
Z∞
q ∂ dx
= p
2 ∂d x2 + (d2 + z 2 )x + d2 z 2
0
q ∂ h p i∞
= ln 2 x2 + (d2 + z 2 ) x + d2 z 2 + 2x + d2 + z 2
2 ∂d 0
q ∂ q
= (−) ln(z + d)2 = − . (3.10)
2 ∂d z+d
(It can be seen in two different, but equivalent, ways that the upper limit does
∂
not contribute. If we take the derivative ∂d first and then the limit, it is obvious.
74 3 Boundary value problems
Alternatively, we note that the upper limit gives rise to an infinite constant which
∂
vanishes upon taking the derivative ∂d .) We recognize this to be the potential
at a point along the z-axis in the upper half plane due to the fictitious “image”
charge (see (3.5)). The electric field due to the induced surface charge can now
be calculated and gives the value for z = d to be (actually, this can be obtained
from the general expression for the E field derived in (3.6) as well)
∂ q q
E(x = 0, y = 0, z = d) = −ẑ − = − 2 ẑ. (3.11)
∂z z + d z=d 4d
Therefore, the force of attraction experienced by the point charge has the value
(this is the force of attraction between the conducting plane and the point charge)
q2
F = qE(x = 0, y = 0, z = d) = − ẑ. (3.12)
4d2
Once again, we see that this is exactly the attractive force between the point charge
and the “image” charge and, consequently, on the positive z-axis the “image”
charge truly reproduces the effect of the induced surface charge on the plane.
Finally, let us note that we can calculate the electrostatic energy of the
system in a simple way. Let us recall that if the point charge is at a distance
z from the plane on the positive z-axis, then, the force between the conducting
plate and the charge is given by (see (3.12))
q2
F(z) = − ẑ. (3.13)
4z 2
Using this, we can calculate the work that must be done to bring the point charge
from infinity to a distance d above the plane on the z-axis. From the defining
relation, we have
Zdẑ Zd d
q2 dz q 2 q2
W =− dℓ · F = 2
=− =− . (3.14)
4 z 4z ∞ 4d
∞ ∞
In deriving this result, we have used the fact that the work is independent of the
choice of the path and, consequently, we have chosen a path along the z-axis for
simplicity. It is worth noting here that this energy is half the energy between the
point charge and its “image”. This can be understood in the following way. The
“image” charge is not real. But, if we had calculated the work done to bring the
point charge and its “image” from infinity to the final position, we would have
done twice the work because we have to move the “image” charge as well. This
is, of course, wrong because the “image” charge is not real. So, had we calculated
the electrostatic energy using the “image” charge, we would have obtained an
erroneous result. (An alternative way to see why one would get twice the result
using the “image” charge is to note that the electrostatic energy is related to
the square of the electric field integrated over the entire volume, as discussed in
chapter 1. With the “image” charge, of course, there is an associated electric
field even in the lower half of the plane and by symmetry, it contributes an exact
amount as the physical electric field above the plane. That is why the electrostatic
energy calculated with the “image” charge would give twice the actual value. In
reality, of course, there is no electric field in the lower half plane and that is how
the error creeps in.)
Although we have found a solution to the problem of a point charge above
a grounded conducting plane of infinite extent by the method of images, it is not
3.1 Method of images 75
clear whether this solution is unique. The uniqueness of the solution can be seen
only from an analysis of the Laplace or the Poisson equation which we will discuss
later in this chapter. ◭
y
d1
−q b b q
d2
2d2 x
2d1
q b b
−q
Let the two infinite, orthogonal and intersecting conducting planes be de-
scribed by x = 0, y ≥ 0 and y = 0, x ≥ 0 respectively. If we assume that the point
charge is on the plane z = 0, then, it is easy to conclude that all the “image”
charges would also lie on the same plane. In fact, it is easy to check that with the
choice of the “image” charges shown in the figure, the potential at any point on
the plane at x = 0 is
q q
Φ(x = 0, y, z) = 1 − 1
(d21 + (d2 − y)2 + z 2 ) 2 (d21 + (d2 − y)2 + z 2 ) 2
q q
+ 1 − 1
(d21 + (d2 + y)2 + z 2 ) 2 (d21 + (d2 + y)2 + z 2 ) 2
= 0. (3.15)
Thus, these image charges indeed reproduce the boundary condition of vanishing
potential on the two infinite planes.
76 3 Boundary value problems
Once we have determined the “image” charges, we can forget about the
conducting planes and determine the potential in the region x > 0 and y > 0
simply to be
q
Φ(x, y, z) = 1
((x − d1 )2 + (y − d2 )2 + z 2 ) 2
q q
− 1 − 1
((x + d1 )2 + (y − d2 )2 + z2) 2 ((x − d1 )2 + (y + d2 )2 + z 2 ) 2
q
+ 1 . (3.17)
((x + d1 )2 + (y + d2 )2 + z 2 ) 2
The electric field can be obtained from (3.17) by taking the gradient. It has a
general structure (namely, all the x̂, ŷ, ẑ components are nonzero) at an arbitrary
point. However, close to the planes, the electric field takes a simpler form. For
example, when x = 0, the electric field has the form
E(x = 0, y > 0, z)
" #
1 1
= 2qd1 x̂ 3 − 3 , (3.18)
(d21 + (y + d2 )2 + z 2 ) 2 (d21 + (y − d2 )2 + z 2 ) 2
E(x > 0, y = 0, z)
" #
1 1
= 2qd2 ŷ 3 − 3 . (3.19)
((x + d1 )2 + d22 + z 2 ) 2 ((x − d1 )2 + d22 + z 2 ) 2
The surface charge densities now follow from the discontinuities of the elec-
tric field across the two planes. Namely,
1
σ(x = 0, y > 0, z) = x̂ · E(x = 0, y > 0, z)
4π
" #
qd1 1 1
= 3 − 3 ,
2π (d21 + (y + d2 )2 + z 2 ) 2 (d21 + (y − d2 )2 + z 2 ) 2
1
σ(x > 0, y = 0, z) = ŷ · E(x > 0, y = 0, z)
4π
" #
qd2 1 1
= 3 − 3 . (3.20)
2π ((x + d1 )2 + d22 + z 2 ) 2 ((x − d1 )2 + d22 + z 2 ) 2
Using (3.20), the total induced charge on the conducting planes can now be de-
termined as follows.
Z∞ Z∞
Qinduced = dy dz σ(x = 0, y > 0, z)
0 −∞
Z∞ Z∞
+ dx dz σ(x > 0, y = 0, z)
0 −∞
3.1 Method of images 77
ZZ " #
∞
qd1 1 1
= dydz 3 − 3
π 0 (d21 + (y + d2 )2 + z 2 ) 2 (d21 + (y − d2 )2 + z 2 ) 2
ZZ " #
∞
qd2 1 1
+ dxdz 3 − 3
π 0 ((x + d1 )2 + d22 + z 2 ) 2 ((x − d1 )2 + d22 + z 2 ) 2
Z∞
2qd1 1 1
=− dy − 2
π y 2 + 2d2 y + d21 + d22 y − 2d2 y + d21 + d22
0
Z∞
2qd2 1 1
− dx − 2
π x2 + 2d1 x + d21 + d22 x − 2d1 x + d21 + d22
0
2qd1 1 −1 d2 2qd2 1 −1 d1
=− tan − tan
π d1 d1 π d2 d2
2q d2 d1
=− tan−1 + tan−1
π d1 d2
2q π
=− × = −q. (3.21)
π 2
In deriving this result, we have used some standard integrals from the tables
(Gradshteyn and Ryzhik, 2.172 and 2.271) as well as the trigonometric relation
that, for x > 0,
1 π
tan−1 x + tan−1 = . (3.22)
x 2
Once again, we see that the total induced charge (on the two plates) is equal to
the sum of all the “image” charges (which is equal in magnitude to the physical
point charge, but opposite in sign). However, the amount of charges on the two
plates depends on the ratio of the perpendicular distances d1 and d2 of the point
charge from the two planes. ◭
(q ′ , d′ ) (q, d)
b b
r−
dẑ
R r
q q′
Φ(r) = +
|r − dẑ| |r − d′ ẑ|
q q′
= 1 + 1 . (3.23)
(r 2 + d2 − 2rd cos θ) 2 (r 2 + d′2 − 2rd′ cos θ) 2
If we require that this potential vanishes when r = R for any angle θ and φ
(namely, on the surface of the sphere), we obtain,
Since this must hold for any θ, the coefficient of the last term in (3.24) must vanish
leading to
d′
q ′2 = q 2 . (3.25)
d
Substituting this back into (3.24), the other two terms lead to
R2 R
d′ = , q ′ = −q . (3.26)
d d
Actually, there are two solutions for the charge (as is clear from (3.25)), but this is
the one which gives a vanishing potential (see (3.23)) on the surface of the sphere
r = R.
Thus, we see that, studying the problem of a point charge outside a con-
ducting sphere which is grounded, is equivalent to studying the point charge in
the presence of an “image” charge inside the sphere. The potential at an arbitrary
point outside the sphere is now easily determined to be
3.1 Method of images 79
q q′
Φ(r) = 1 + 1
(r 2 + d2 − 2rd cos θ) 2 (r 2 + d′2 − 2rd′ cos θ) 2
!
1 R
=q 1 − 1 . (3.27)
(r 2 + d2 − 2rd cos θ) 2 (r 2 d2 + R4 − 2rdR2 cos θ) 2
The electric field can, of course, be calculated from this by taking the gradient
and it is clear that it would, in general, have both a radial and an angular compo-
nent. However, near the surface of the sphere, namely, when r = R, the angular
components cancel out (The simplest way to see this is to note that when taking
derivative with respect to θ, we can set r = R, but then the potential vanishes and
so does the θ component of the gradient at such points. Physically, of course, this
means that there is no tangential component of E on the surface of the conductor
as we would expect.) and we have only a radial component
∂
E(r)|r=R = − ∇Φ(r)|r=R = − r̂ Φ(r)
∂r r=R
!
(R − d cos θ) R(Rd2 − R2 d cos θ)
= qr̂ 3 − 3
(d2 + R2 − 2dR cos θ) 2 R3 (d2 + R2 − 2dR cos θ) 2
q(d2 − R2 ) r̂
=− 3 . (3.28)
R (d2 + R2 − 2dR cos θ) 2
Consequently, we can determine the surface distribution of the charges from the
discontinuity of the electric field, namely,
1
σ(r = R, θ, φ) = r̂ · E(r)|r=R
4π
q(d2 − R2 ) 1
=− 3 . (3.29)
4πR 2 2
(d + R − 2dR cos θ) 2
The total induced charge on the surface of the sphere can now be obtained
from (3.29) through a simple integration and we obtain,
Z
Qinduced = R2 sin θ dθ dφ σ(r = R, θ, φ)
Z1
q(d2 − R2 ) dx
=− × 2πR2 3
4πR (d2 + R2 − 2dRx) 2
−1
1
q(d2 − R2 )R 1 1
=− 1
2 (d + R − 2dRx) −1
dR
2 2 2
q(d2 − R2 ) 1 1
=− −
2d d−R d+R
R
= −q = q′ . (3.30)
d
This again shows that the total charge induced on the sphere is identical to the
“image” charge (which is not equal in magnitude to the point charge in this case).
80 3 Boundary value problems
The force of attraction between the point charge and the sphere can again
be calculated directly or from the “image” charge.
qq ′ q Rd
F(x = 0, y = 0, z = d) = ẑ = −q × 2 ẑ
(d − d′ )2 2
d − Rd
q 2 dR
=− ẑ. (3.31)
(d2 − R2 )2
It is clear from this that, at any point on the z-axis, the force experienced by the
point charge is
q 2 zR
F(x = 0, y = 0, z) = − ẑ. (3.32)
(z 2− R 2 )2
Therefore, the work done in bringing the charge from infinity is easily obtained
to be
Zdẑ Zd
q 2 zR
W = − dℓ · F = − dz − 2
(z − R2 )2
∞ ∞
d
q2 R 1 q2 R
= × − 2 =− . (3.33)
2 z − R2 ∞ 2(d2 − R2 )
Once again, this is half of the energy that we would have found from a calculation
using the “image” charge for reasons which we have discussed earlier. The method
of images works well for all quantities except for the energy of the system. ◭
"
dm f 1 df
=− −F (x0 ) + am (x0 )f (x0 ) + am−1 (x0 )
dxm x0 a0 (x0 ) dx x0
#
dm−1 f
+ · · · + a1 (x0 ) , (3.36)
dxm−1 x0
Here, we are going to assume that the equation is linear and that the
coefficients A, B and C are, in general, functions of x, y. Furthermore,
the two independent variables x, y can both be space coordinates or
one space and one time coordinate.
Unlike the one dimensional ordinary differential equation that
we discussed earlier, here specifying the function and its (two) first
derivatives at a point will not be enough to determine the solution
uniquely. Rather, we need to specify appropriate boundary condi-
tions on a curve. (In general, the solution of a partial differential
equation in n variables needs boundary conditions specified on a
(n−1) dimensional hypersurface.) Furthermore, as we will see Cauchy
boundary conditions may not always work in these cases because they
may over-specify the solution.
To understand the nature of boundary conditions and the curve
on which they must be specified, let us represent the boundary curve
parametrically by ξ = x(s) and η = y(s) where s is the distance of a
point on the curve from some reference point. At any point on the
curve, there are two orthogonal directions – one along the tangent to
the curve and the other normal to it (see Fig. 3.6). The unit vectors
along these directions are easily determined to be
y
êt
ên
Figure 3.6: Unit vectors êt and ên which are respectively tangential
and normal to the curve at a given point.
dξ dη
êt = x̂ + ŷ,
ds ds
dη dξ
ên = ẑ × êt = − x̂ + ŷ. (3.38)
ds ds
84 3 Boundary value problems
which follows from the fact that the infinitesimal distance between
two points on the curve can be written as
ds2 = dξ 2 + dη 2 . (3.40)
df
≡ , (3.41)
ds
dη ∂f dξ ∂f
fn = ên · ∇f |x(s),y(s) =− + .
ds ∂x x(s),y(s) ds ∂y x(s),y(s)
This analysis makes it clear that once we know f (s) along the
curve, we also know its derivative along the curve, since ft = df ds .
Therefore, the first order derivative which needs to be specified as
an independent boundary condition (for the Cauchy problem) is the
derivative normal to the curve or fn . A solution at any point will, of
course, have a Taylor expansion of the form
∂f dξ dη
= ft − fn = a(s),
∂x x(s),y(s) ds ds
∂f dη dξ
= ft + fn = b(s). (3.43)
∂y x(s),y(s) ds ds
Thus, we see that given f (s) and fn , we can determine the (two) first
derivatives directly from the data (remember ft = dfds ). Furthermore,
since a(s) and b(s) are known functions, by taking their derivative as
well as using the differential equation, we have
dξ ∂ 2 f dη ∂ 2 f da(s)
2
+ = ,
ds ∂x x(s),y(s) ds ∂x∂y x(s),y(s) ds
dξ ∂2f dη ∂2f db(s)
+ = ,
ds ∂x∂y x(s),y(s) ds ∂y 2 x(s),y(s) ds
∂2f ∂2f ∂2f
A(s) 2 + 2B(s) + C(s) 2 = F (s). (3.44)
∂x ∂x∂y ∂y x(s),y(s)
√
dη B± B 2 − AC dξ
= ,
ds A ds
√
B± B 2 − AC
or, dη = dξ. (3.47)
A
If this holds, then the Cauchy problem cannot be uniquely solved.
These two equations are equations for two curves which in the present
case are known as the characteristic curves and we see that the
Cauchy problem cannot be solved if the Cauchy data are specified
on any of the characteristics. This is reminiscent of the behavior in
the case of ordinary differential equations.
For a given second order equation, A, B, and C are known func-
tions and depending on the behavior of the radical in (3.47), partial
differential equations can be classified into three different groups. If
B 2 > AC, then we see that there are two real characteristic curves
of the equation. Such equations are known as hyperbolic equations.
The most familiar of the hyperbolic equations is the wave equation
of the form (in 1 + 1 dimensions)
∂2f 1 ∂2f
− = 0, (3.48)
∂x2 v 2 ∂t2
where v represents the speed of propagation of the wave.
y y
x x
Figure 3.7: A curve intersects the characteristic curves (on the left)
while it is tangential to a characteristic (on the right).
∂2f ∂2f
∇2 f = + = 0, (3.49)
∂x2 ∂y 2
∂2f ∂f
=α , (3.50)
∂x2 ∂t
is an example of a parabolic equation. In the case of a parabolic
equation, there is only one characteristic curve and a unique solution
can be obtained only for either Dirichlet or Neumann (or a mixed)
boundary condition on an open curve.
Although our discussion so far has been within the context of
two dimensions, it can be generalized to higher dimensions as well.
For a general second order partial differential equation in n dimen-
sions, we can always find a suitable coordinate transformation to
diagonalize the equation. When, only one of the coefficients of the
diagonalized equation is negative, the equation is known as a hyper-
bolic equation. If none of the coefficients of the diagonalized equation
is negative, then, it is known as an elliptic equation and if any one
of the coefficients vanishes, then, the equation is called a parabolic
equation. In general, hyperbolic equations are solved uniquely by
specifying Cauchy boundary conditions on an open hypersurface, el-
liptic equations by specifying Dirichlet/Neumann (or mixed) bound-
ary conditions on a closed hypersurface and parabolic equations by
specifying Dirichlet/Neumann (or mixed) boundary conditions on an
open hypersurface.
∇2 Φ = 0, (3.51)
Φ = Φ1 − Φ2 , (3.52)
3.2 Boundary conditions for differential equations 89
∇2 Φ = ∇2 Φ1 − ∇2 Φ2 = 0. (3.53)
Here, we have used the fact that Φ satisfies the Laplace equation
(3.53) so that the second term on the left hand side does not con-
tribute. Furthermore, for either Dirichlet or Neumann boundary con-
ditions on S, either Φ or ∂Φ
∂n vanishes on S so that the right hand side
identically vanishes. The final result, as it stands, shows that the in-
tegral of a positive quantity vanishes and, consequently, the integrand
on the left in (3.54) must vanish, namely,
∇Φ = 0. (3.55)
n̂ · ∇Φ = 0, (3.56)
∇Φ = 0. (3.58)
Φ = 0, (3.59)
and, consequently, the two solutions are the same. However, for
Neumann boundary conditions, Φ = constant which for the purposes
of calculating electric field etc. still implies that the solution is unique.
∇2 Φ = 0, (3.63)
2. Let us note that, given any two functions A and B, we have the
identity
1
This is known as Green’s identity. If we now choose A = |r|
and B = Φ, namely, if B represents a solution of the Laplace
equation (3.63) and A satisfies (this is known as the Green’s
function for the Laplacian)
1
∇2 A = ∇2 = −4π δ3 (r), (3.68)
|r|
then, from (3.67) (we are going to assume that S is the surface
of a sphere of radius R), we obtain
Z Z
3 3 1 r̂
4π d r Φ(r) δ (r) = ds · (∇Φ) + 2 Φ ,
V S R R
Z
1
or, Φ(r = 0) = ds Φ(R, θ, φ). (3.69)
4πR2 S
92 3 Boundary value problems
L1
L3 y
L2
x
boundary conditions
Φ(0, y, z) = Φ(x, 0, z) = Φ(x, y, 0) = 0,
Φ(L1 , y, z) = Φ(x, L2 , z) = 0,
Φ(x, y, L3 ) = f (x, y), (3.73)
in this region.
In trying to solve the Laplace equation for this system, let us
note that the Laplacian in Cartesian coordinates (see (3.72)) is really
a sum of three terms which commute with one another. In such a case,
it is a general result that the solution can be written as a product of
three terms, each depending on only one coordinate, namely, in such
a case, we expect the solution to have the factorized form
Φ(x, y, z) = X(x)Y (y)Z(z). (3.74)
Substituting this form of the solution into the Laplace equation (3.72)
and dividing by Φ throughout, we obtain
1 d2 X(x) 1 d2 Y (y) 1 d2 Z(z)
+ + = 0. (3.75)
X(x) dx2 Y (y) dy 2 Z(z) dz 2
Since each of the three terms in the above expression depends on only
one coordinate x, y or z, their sum can vanish for arbitrary values of
the coordinates only if each of the terms equals a constant such that
1 d2 X(x)
= α1 ,
X(x) dx2
1 d2 Y (y)
= α2 ,
Y (y) dy 2
1 d2 Z(z)
= α3 , α1 + α2 + α3 = 0. (3.76)
Z(z) dz 2
The boundary conditions (3.73) for Φ can now be translated to
conditions on the individual component functions as
X(0) = X(L1 ) = Y (0) = Y (L2 ) = Z(0) = 0,
2
nπy (n) nπ
Yn (y) = sin , α2 =− , (3.78)
L2 L2
2 2
sinh(αmn z) (m,n) mπ nπ
Zmn (z) = , α3 = α2mn = + ,
sinh(αmn L3 ) L1 L2
with m, n = 1, 2, . . . and we have normalized the z solution for later
convenience. The determination of these solutions uses only the five
homogeneous boundary conditions on Φ(x, y, z) with the condition
for the surface z = L3 in (3.73) yet to be implemented.
A general solution of the problem can now be written as a linear
superposition of the form (it is not unique yet)
∞
X
Φ(x, y, z) = Amn Xm (x)Yn (y)Zmn (z)
m,n=1
∞
X
mπx nπy sinh(αmn z)
= Amn sin sin ,(3.79)
L1 L2 sinh(αmn L3 )
m,n=1
Substituting this back into the Laplace equation and dividing through-
out by rΦ2 , we obtain
1 d 2 dR 1 d dΘ 1 1 d2 Q
r + sin θ + = 0.
R dr dr Θ sin θ dθ dθ sin2 θ Q dφ2
(3.85)
Since the expression within the parenthesis in the last term of (3.85)
is the only term which depends on φ, this equation cannot be satisfied
(for arbitrary values of φ) unless this term equals a constant, namely,
we must have
1 d2 Q
= −m2 , (3.86)
Q dφ2
where the choice of the sign of the constant is for convenience. Equa-
tion (3.86) can be readily integrated to give
Here we have used the fact that since the left hand side and the right
hand side of (3.89) depend on independent variables, this relation
will hold only if each side equals a constant which we denote by k.
We see from (3.89) that the θ equation takes the simple form
1 d dΘ m2
sin θ + k− Θ = 0. (3.90)
sin θ dθ dθ sin2 θ
is known as the Legendre equation and the solutions, Pℓ (x), are poly-
nomials of order ℓ known as the Legendre polynomials. It can be eas-
ily checked that the associated Legendre polynomials are related to
the Legendre polynomials through the relation (associated Legendre
polynomials are not really polynomials for odd values of m as is clear
from (3.94))
1 dℓ 2
Pℓ (x) = (x − 1)ℓ , (3.95)
2ℓ ℓ! dxℓ
from which the explicit forms of the first few Legendre polynomials
can be easily determined to be
P0 (x) = P0 (cos θ) = 1,
P1 (x) = P1 (cos θ) = x = cos θ,
1 1
P2 (x) = P2 (cos θ) = (3x2 − 1) = (3 cos2 θ − 1), (3.96)
2 2
and so on. These are precisely the functions that we encountered in
the last chapter in connection with the expansion of the potential
for a point charge displaced along the z-axis (see (2.38)). It is also
clear that we can write a closed form expression for the associated
Legendre polynomials from the Rodrigues’ formula as well.
It is also possible to write a generating function for the Legendre
polynomials in a simple manner. Consider a function of two variables
1
T (x, s) = 1 . (3.97)
(1 − 2sx + s2 ) 2
1 1
′
= 1
|r − r | (r 2 − 2rr ′ cos θ + r ′2 ) 2
1 1
=
r 2r ′
r′ 2
1
2
1− r cos θ + r
3.3 Solutions of the Laplace equation 99
∞
1 X r′ ℓ
= Pℓ (cos θ), (3.99)
r r
ℓ=0
where we have assumed that r > r ′ . This is, of course, the expansion
of the potential that we have discussed earlier in (2.39) and (2.46).
A similar expansion is trivially obtained when r < r ′ , simply by
interchanging r and r ′ in the previous expression (which is symmetric
in r, r ′ ).
The Legendre polynomials can be shown to satisfy the orthonor-
mality relation
Z1
2
dx Pℓ (x) Pℓ′ (x) = δℓℓ′ . (3.100)
2ℓ + 1
−1
where Aℓ,m ’s are constants, which can be determined from the given
boundary conditions of a physical system. We note here that if we
have a physical problem where the potential does not depend on the
azimuthal angle φ (namely, when m = 0), the corresponding solutions
are known as zonal harmonics.
◮ Example (Sphere in uniform electric field). As an example of a physical sys-
tem where spherical solutions of the Laplace equation may be used, let us con-
sider space without any free charge consisting of a uniform background electric
field along the z-axis. Thus, the field lines can be drawn as parallel lines of the
same magnitude and we can write
Φ(r) = −Ez + C = −Er cos θ + C = −ErP1 (cos θ) + CP0 (cos θ), (3.107)
Here even though we are interested in the solution outside the sphere (the origin
of the coordinate system is chosen to be at the center of the sphere), we have
allowed for both the independent forms of the radial solution since the potential
for a constant electric field is linear in r in this region. Furthermore, since there
is no free charge anywhere (so that Φ satisfies the Laplace equation) and
1
∇2 = −4πδ 3 (r), (3.109)
r
A0 = C, A1 = −E. (3.111)
All the terms with coefficients Bℓ vanish asymptotically and hence there is no
constraint on these coefficients from the asymptotic condition. Thus, we can
write for r ≥ R,
∞
X
Φ(r) = C − Er cos θ + Bℓ r −(ℓ+1) Pℓ (cos θ). (3.112)
ℓ=1
102 3 Boundary value problems
We still have to satisfy the boundary condition that the potential on the surface
of the sphere is a constant Φ0 . Requiring this, we obtain
∞
X
Φ0 = C − ER cos θ + Bℓ R−(ℓ+1) Pℓ (cos θ), (3.113)
ℓ=1
which determines
E(r) = −∇Φ(r)
!
∂ θ̂ ∂ R3
→ − r̂ + Φ0 − E 1 − 3 r cos θ
∂r r ∂θ r
2R3 R3
= r̂ E 1 + 3 cos θ − θ̂ E 1 − 3 sin θ. (3.116)
r r
Thus, we see that, in general, the electric field has a radial as well as a θ
component. However, on the surface of the sphere (r = R), the theta component
vanishes, so that the electric field is normal to the surface. Furthermore, it is not
a uniform electric field on the surface, rather its value depends on the angle θ.
From this, we can determine the surface charge density induced on the sphere to
be (we are using the fact that E = 0 inside the condocting sphere)
1 3E
σ(R, θ) = r̂ · E(R) = cos θ. (3.117)
4π 4π
Integrating this over the surface of the sphere, we obtain
Z
Qinduced = R2 dΩ σ(R, θ)
Zπ Z2π
3ER2
= dθ sin θ cos θ dφ = 0. (3.118)
4π
0 0
This is consistent with our earlier discussion in the last chapter. Namely, the
conducting sphere remains neutral, the positive and the negative charges simply
rearrange themselves so as to cancel the electric field inside the sphere. Note from
(3.116) that asymptotically for large r, the electric field becomes
consistent with (3.106). In deriving the surface charge density we have assumed
that there is no electric field inside the conducting sphere. However, this analysis
can also be carried out in the interior of the sphere as well to show that the
potential is a constant Φ0 and there is no electric field inside which also follow
from the general properties of harmonic functions that we discussed earlier. ◭
3.3 Solutions of the Laplace equation 103
Here, we have used the fact that both sides of (3.122) are functions
of independent variables and, therefore, the relation can be satisfied
for arbitrary values of r, φ only if both sides equal a constant which
we have identified with k. The solution for the φ equation is straight-
forward. We note that we will have a single valued function only if
k = n2 , n = 0, 1, . . ., with the φ solution written as
so that a general solution for the radial equation (3.124) can be writ-
ten (for n 6= 0) as
Rn (r) = Cn r n + Dn r −n , (3.126)
104 3 Boundary value problems
R0 (r) = C0 + D0 ln r, (3.127)
so that the potential will have the form (in polar coordinates)
where C is a constant and r is the radial coordinate on the plane. In the presence
of the conducting cylinder, the field lines will be distorted near the surface of the
cylinder, but asymptotically they will have the form (3.131). The surface of the
cylinder, of course, would be at a constant potential which we take to be Φ0 .
The general solution for the potential in the presence of the cylinder (out-
side) would have the form (since there are no free charges, the potential will satisfy
Laplace equation and in this case will have a solution of the form (3.129))
Φ(r) = C0 + D0 ln r
∞
X
+ (Cn r n + Dn r −n )(An cos nφ + Bn sin nφ). (3.132)
n=1
However, comparing with the asymptotic form of the potential (3.131) and using
the orthonormality relations for the trigonometric functions, we determine that
the only coefficients that are nontrivial are A1 , C0 , C1 , D1 and satisfy
C0 = C, A1 C1 = −E, (3.133)
so that we can write the form of the potential outside the cylinder to be (r ≥ R)
Φ(r) = C − Er cos φ + A1 D1 r −1 cos φ. (3.134)
We see that the term with the coefficient D1 vanishes asymptotically and, as a
consequence, there is no constraint on this coefficient from the asymptotic condi-
tion. (Remember that r is the radial coordinate in cylindrical coordinates and,
therefore, there is no inconsistency with the absence of free charges. Furthermore,
this term also involves an angular function.)
Let us next impose the boundary condition on the surface of the cylinder.
Namely, on the surface of the cylinder (r = R) we have,
Φ0 = C − ER cos φ + A1 D1 R−1 cos φ, (3.135)
which determines
C = Φ0 , A1 D1 = ER2 , (3.136)
so that we can write the potential outside the cylinder (3.134) to be
R2
Φ(r, φ) = Φ0 − E r − cos φ. (3.137)
r
We can now determine the electric field which has the form
E(r, φ) = −∇Φ(r, φ)
!
∂ φ̂ ∂ R2
= − r̂ + Φ0 − E r − cos φ
∂r r ∂φ r
R2 R2
= r̂E 1 + 2 cos φ − φ̂E 1 − 2 sin φ. (3.138)
r r
Once again we see that the electric field outside the cylinder has both a radial
and an angular component. However, on the surface of the cylinder (r = R), only
the normal (radial) component is nontrivial. This allows us to determine the
induced surface charge density on the cylinder to be
1 E
σ(R, φ) = r̂ · E(R) = cos φ, (3.139)
4π 2π
106 3 Boundary value problems
where we are assuming that there is no electric field inside the cylinder since the
potential is a constant. Consequently, the total induced charge on the surface of
the cylinder is obtained to be (L is the length of the cylinder which is to be taken
to infinity at the end)
Z Z2π
ELR
Qinduced = LR dφ σ(R, φ) = dφ cos φ = 0. (3.140)
2π
0
We see that the cylinder remains neutral. The charges simply rearrange them-
selves on the surface to yield zero electric field inside the cylinder, which can be
explicitly checked by carrying out a similar analysis inside the cylinder. Let us
also note that asymptotically,
lim E → E (r̂ cos φ − φ̂ sin φ) = E x̂, (3.141)
r→∞
as we would expect. ◭
= −4πρ(r). (3.145)
which is the familiar relation for the potential due to a charge distri-
bution that we have discussed earlier (see (1.60)).
The relation between the solution of an inhomogeneous problem
and the Green’s function can be easily seen from the Green’s identity
which we have derived earlier (see (3.67)), namely,
Z Z
′ ′
d3 r ′ A∇ 2 B − B∇ 2 A = ds′ · A∇′ B − B∇′ A . (3.151)
V S
V S
Here, we have used the Poisson equation (3.142) as well as the equa-
tion satisfied by the Green’s function (3.143). Furthermore, if we use
∂
the notation ês · ∇ = ∂n to represent the normal derivative, we can
rewrite (3.153) also as
Z ′ ′
1 ′ ′ ∂Φ(r ) ′ ∂G(r, r )
Φ(r) = ds G(r, r ) − Φ(r )
4π ∂n′ ∂n′
S
Z
+ d3 r ′ G(r, r′ )ρ(r′ ). (3.154)
V
This can be compared with the form of the result obtained earlier in
(3.149). Incidentally, the simplest way to check that the first term
in (3.154) must be a solution of the homogeneous equation is to note
that
Z
∇2 Φ(r) − d3 r ′ G(r, r′ )ρ(r′ ) = 0. (3.155)
V
A direct verification of this relation for the surface integral is, how-
ever, tricky.
which is indeed a well defined solution for the given Dirichlet bound-
ary value problem. It is worth noting that this solution is valid even
in the region which does not contain any charges (compare this with
the method of images where one does not calculate the potential in
the region containing the image charge). In such a case, the second
term vanishes and the solution is given completely by the surface
integral. (Let us also emphasize here that even though the bound-
ary condition for the solution may be inhomogeneous, the Green’s
function is required to satisfy only the homogeneous condition.)
The main question to answer now is what is the form of the
Green’s function satisfying the homogeneous Dirichlet boundary con-
dition (3.156). The exact structure of the Green’s function would, of
course, depend on the particular problem under consideration. How-
ever, let us note that we can write a general Green’s function to be
of the form
where
∇2 H(r, r′ ) = 0. (3.159)
Although there are two possible choices for H(r, r′ ) we note that for the second
choice of the sign in the denominator in (3.161) we will have GD (r, r′ ) = 0 which
is a trivial solution. Therefore, the first choice of the sign is the natural one and
we obtain
1
GD (r, r′ ) = p
(x − x ′ )2 + (y − y ′ )2 + (z − z ′ )2
1
−p . (3.162)
(x − x′ )2 + (y − y ′ )2 + (z + z ′ )2
There are two things to note from (3.162). First, with our choice of H(r, r′ ),
we note that
!
2 ′ 2 1
∇ H(r, r ) = −∇ p
(x − x′ )2 + (y − y ′ )2 + (z + z ′ )2
Therefore, it appears that H(r, r′ ) does not satisfy the homogeneous equation.
However, note that in the region that we are interested in, namely, z, z ′ > 0, the
right hand side of (3.163) indeed vanishes and H(r, r′ ) satisfies the homogeneous
equation in this region. The second thing to note is that the Green’s function
in (3.162) is reminiscent of the structure of the potential for a grounded plane
obtained by the method of images in (3.5). This should not be surprising since
the Green’s function is the potential for a unit source charge with a homogeneous
Dirichlet boundary condition.
Substituting (3.162) into the right hand side of the solution (3.157), we
obtain (S in this case is the plane z ′ = 0. Actually, S is the closed surface
bounding the upper half plane. However, as is clear from the form of the Green’s
function, the Green’s function as well as its derivative vanish when any of the
coordinates is at infinity. Consequently, the surface S is effectively the plane at
z ′ = 0.)
Z Z
1 ∂GD (r, r′ )
Φ(r) = − ds′ Φ(r′ ) + d3 r ′ GD (r, r′ )ρ(r′ )
4π ∂n′
S
Z
Φ0 2zdx′ dy ′
= 3
4π ((x − x′ )2 + (y − y ′ )2 + z 2 ) 2
Z
+ d3 r ′ GD (r, r′ )qδ(x′ )δ(y ′ )δ(z ′ − d)
Z
Φ0 z 2dx′ dy ′
= 3
4π (x′2 + y ′2 + z 2 ) 2
!
1 1
+q 1 − 1 . (3.164)
(x2 + y 2 + (z − d)2 ) 2 (x2 + y 2 + (z + d)2 ) 2
(Note that there is an extra negative sign in the surface term because the sur-
∂ ∂
face area points along the negative z-axis so that ∂n ′ = − ∂z ′ .) Here, we have
translated the coordinates of integration in the first integral for simplicity and
112 3 Boundary value problems
Z∞
dr ′2
= 2π 3
(r ′2 + z 2 ) 2
0
∞
1
= 2π (−2) √ = 4π . (3.165)
r + z 0
′2 2 z
Substituting this back into the solution (3.164), we have
!
1 1
Φ(r) = Φ0 + q 1 − 1 . (3.166)
(x2 + y 2 + (z − d)2 ) 2 (x2 + y 2 + (z + d)2 ) 2
For Φ0 = 0, this reduces to the solution (3.5). However, we now have the solution
for the case when the conducting plane is not grounded, but is held at a nonzero
constant potential. Incidentally, from (3.166) (or (3.164)), we see that the poten-
tial in the lower half plane is given by (the second term vanishes in the lower half
plane because there is no free charge in that region)
Φ(r) = Φ0 , for z ≤ 0, (3.167)
which is consistent with what we expect, namely, that there is no electric field in
the lower half plane. All the field lines from the point charge terminate on the
conducting plane. (An alternative way to see that the surface term in (3.164) for
a constant potential Φ0 on the boundary is simply equal to Φ0 is as follows. By
definition
∇′2 GD (r, r′ ) = −4πδ 3 (r − r′ ). (3.168)
Consequently, using Gauss’ theorem we can write
Z Z
ds′ · ∇′ GD (r, r′ ) = d3 r ′ ∇′2 GD (r, r′ )
S
Z
= d3 r ′ (−4π)δ 3 (r − r′ ) = −4π, (3.169)
◮ Example (Point charge inside a conducting sphere). Let us next analyze the
problem of a conducting sphere of radius R which contains a point charge q in-
side, at a distance d from the center of the sphere. The surface of the sphere is
maintained at a constant potential Φ0 and we are interested in determining the
potential within the sphere.
For simplicity, let us choose the origin to coincide with the center of the
sphere and the point charge q to lie on the z-axis so that we are interested in the
solution of the equation
∇2 Φ = −4π qδ(x)δ(y)δ(z − d), 0 ≤ x, y, z ≤ R, d < R, (3.170)
subject to the boundary condition that
Φ(R, θ, φ) = Φ0 . (3.171)
3.4 Solution of the Poisson equation 113
Therefore, this defines a Dirichlet problem. Once again, the Green’s function for
the problem, satisfying the homogeneous Dirichlet boundary condition, can be
written as (see (3.158))
1
GD (r, r′ ) = + H(r, r′ ). (3.172)
|r − r′ |
Since H(r, r′ ) has to satisfy the homogeneous equation (for r, r ′ < R), we note
that we can write it as
α
H(r, r′ ) = − , (3.173)
|r − r′′ |
where we assume that α is a constant (namely, independent of r) and r′′ = r′′ (r′ )
and lies outside the sphere (namely, r ′′ > R). Now, requiring that GD (r, r′ )|r=R =
0, we obtain
1 α
1 = 1 ,
(R2 + r ′2 − 2Rr ′ cos γ ′ ) 2 (R2 + r ′′2 − 2Rr ′′ cos γ ′′ ) 2
1 R
GD (r, r′ ) = 1 − 1 .
(r 2 + r ′2 − 2rr ′ cos γ ′ ) 2 (r 2 r ′2 + R4 − 2R2 rr ′ cos γ ′ ) 2
(3.177)
It is manifestly symmetric in r ↔ r′ and it is straightforward to verify explicitly
that it satisfies the homogeneous Dirichlet boundary condition whenever r or r′
lies on the surface of the sphere. Furthermore, its structure can be compared with
the potential obtained from the method of images.
The solution of the Poisson equation inside the sphere is now straightforward
∂ ∂ ′
(r < R, ∂n ′ = ∂r ′ ). (For simplicity of evaluation, we will measure the angle θ
′ ′
with respect to the vector r in the first integral so that γ = θ .)
Z Z
1 ∂GD (r, r′ )
Φ(r) = − ds′ Φ(r′ ) + d3 r ′ GD (r, r′ )ρ(r′ )
4π ∂n′
S V
Zπ Z2π
Φ0 (R2 − r 2 ) R2 sin θ′ dθ′ dφ′
= 3
4πR (r 2 + R2 − 2rR cos θ′ ) 2
0 0
Z
+ d3 r ′ GD (r, r′ )qδ(x′ )δ(y ′ )δ(z ′ − d)
114 3 Boundary value problems
Z1
Φ0 (R2 − r 2 )R dx
= 3
2 (r 2 + R2 − 2rRx) 2
−1
!
1 R
+q 1 − 1
(r 2 + d2 − 2rd cos θ) 2(d2 r 2 + R4 − 2R2 dr cos θ) 2
1
Φ0 (R2 − r 2 )R 1 1
= 1
2 Rr (r 2 + R2 − 2rRx) 2
−1
!
1 R
+q 1 − 1
(r 2 + d2 − 2rd cos θ) 2 (d2 r 2 + R4 − 2R2 dr cos θ) 2
!
1 R
= Φ0 + q 1 − 1 .
(r 2 + d2 − 2rd cos θ) 2 (d2 r 2 + R4 − 2R2 dr cos θ) 2
(3.178)
This can be compared with the solution obtained from the method of images. ◭
Here, we have used the fact that Φ satisfies the Poisson equation while
GN satisfies (3.184). Noting that for any function A(r) the average
over a given volume is defined to be
Z
1
Aavg =A= d3 r A(r), (3.187)
V
V
116 3 Boundary value problems
∂Φ 2qd
or, = 3 . (3.189)
∂z (r 2 + d2 ) 2
Here the restriction stands for z = 0, namely, the location of the plane. Conse-
quently, we can use this as a Neumann boundary condition which automatically
satisfies the appropriate constraint and try to solve for the potential of the prob-
lem. We are interested in the solution in the upper half plane so that the volume
is infinite implying that the average term in (3.188) can be ignored.
The Green’s function satisfying the homogeneous Neumann boundary con-
dition can be determined as before and has the form
1
GN (r, r′ ) = p
(x − x′ )2 + (y − y ′ )2 + (z − z ′ )2
1
+p . (3.190)
(x − x ′ )2 + (y − y ′ )2 + (z + z ′ )2
The additional term as we have seen earlier in (3.163), is a solution of the homo-
geneous equation and the sign is chosen such that the z (or z ′ ) derivative vanishes
if either z or z ′ vanishes (on the surface of the plane). The solution can now
be obtained from (3.188). However, unlike the Dirichlet boundary condition it is
clear that the surface term now gives a coordinate dependent term and to compare
with what we have already done by the method of images, let us calculate the
potential along the z-axis (x = 0, y = 0, z ≥ 0). For such points we obtain (the
negative sign in the surface term arises from the direction of the outward normal
∂ ∂
to the surface as discussed earlier, namely, ∂n ′ = − ∂z ′ )
Z Z
1 ∂Φ
Φ(x = 0 = y, z) = ds′ GN (r, r′ ) ′ + d3 r ′ GN (r, r′ )ρ(r′ )
4π ∂n
S
3.5 Selected problems 117
Z
2qd 2r ′ dr ′ dφ′
=− 3 1
4π (r ′2 + d2 ) 2 (r ′2 + z 2 ) 2
Z
+ d3 r ′ GN (r, r′ )qδ(x′ )δ(y ′ )δ(z ′ − d)
Z∞
q ∂ dr ′2
= (2π) 1 1
2π ∂d (r ′2 + d2 ) 2 (r ′2 + z 2 ) 2
0
1 1
+q +
|z − d| z+d
2q q q
=− + +
z+d |z − d| z+d
q q
= − , (3.191)
|z − d| z+d
which is what we had calculated earlier in (3.5) using the method of images. (See,
for example, Gradshteyn and Ryzhik 2.261 for the value of the first integral.) ◭
∇2 Φ1 = −4πρ1 ,
∇2 Φ2 = −4πρ2 .
∂2Φ 1 ∂2Φ
2
− 2 = 0,
∂x v ∂t2
where v represents the speed of propagation of the wave. This
is a hyperbolic equation, and at every point in space-time, there
are two characteristics ξ(x, t) and η(x, t).
a) Determine the characteristics as functions of (x, t).
3.5 Selected problems 119
Dielectrics
Pi = χij Ej . (4.1)
121
122 4 Dielectrics
P = χE, (4.2)
Here, the subscript “b” simply stands for the fact that these charges
are bound and are not free to move around unlike the charges in a
conductor. It follows from the identification in (4.4) that the total
charge in the dielectric is
Z Z
Qb = d3 r ρb (r) + ds σb
V S
Z Z
3
=− d r∇·P+ ds n̂ · P(r)
V S
Z Z
3
=− d r ∇ · P(r) + ds · P(r) = 0, (4.5)
V S
where the last identity follows from Gauss’ theorem. This is, of
course, what we would expect. Namely, the dielectric is charge neu-
tral, all that happens in the presence of an external electric field is
that the charges are displaced slightly to give it a macroscopic dipole
moment.
It is intuitively clear that because a polarized dielectric develops
a volume as well as a surface density of bound charges, the differential
form of Gauss’ law satisfied by the electric field in a dielectric would
4.1 Electric displacement field 123
si b
qi b
S
b
s s2
q11 q2
Z
ds · E = 4π(Q + Qb ), (4.6)
S
where Q represents the sum of free charges embedded inside the di-
electric and Qb is the total bound charge within the Gaussian vol-
ume (which need not be zero unlike the case of the whole dielectric).
By definition (ρb is contained only in the volume excluding the free
charges)
Z Z
Qb = d3 r ρb (r) + ds σb (r)
P
V Si
i
Z Z
3
=− d r (∇ · P) + ds · P
P
V Si
i
Z Z
=− ds · P + ds · P
P P
S+ Si Si
i i
Z
=− ds · P. (4.7)
S
124 4 Dielectrics
Here Si represents the surface area of the interface between the charge
qi and the dielectric. Putting this back into the integral form of
Gauss’ law (4.6), we obtain
Z
ds · (E + 4πP) = 4πQ,
S
Z
or, ds · D = 4πQ,
S
Here, we have used Gauss’ theorem in deriving the last line and have
defined a new vector field
∇ · D = 4πρ. (4.10)
In spite of its similarity with the differential form of Gauss’ law (1.30)
satisfied by the electric field in the absence of a dielectric, the two
vector fields may have quite different characters in general. For ex-
ample, as we have seen in chapter 1 the electric field is conservative,
but D may not be, namely,
∇ × E = 0, ∇ × D 6= 0. (4.11)
This follows because the polarization vector may not have vanishing
curl in general. Furthermore, note that since the polarization vector
is parallel to the electric field, (for isotropic dielectrics) we can write
b q
From the (spherical) symmetry of the problem we see that the electric field
will be radial (with the charge q at the center) at every point. Furthermore,
drawing a Gaussian sphere of radius r we note that the magnitude of the electric
field will be the same at every point on the surface of this sphere. Therefore, we
126 4 Dielectrics
determine trivially from (4.8) that (remember that D is parallel to the electric
field and, consequently, is radial as well)
b q
Figure 4.3: The induced negative charges in the interface of the free
charge in the dielectric medium.
All of this charge lies on the inner surface as shown in Fig. 4.3 and, as a
result, at a distance far away from the origin the total effective (free) charge seen
4.1 Electric displacement field 127
is
(ǫ − 1) q
qeff = q + Qb = q 1 − = . (4.20)
ǫ ǫ
In other words, induced bound charges have a tendency to reduce the magni-
tude of the (free) point charge and, consequently, lead to a weaker electric field.
Conventionally, this is known as the screening of a point charge by a dielectric
medium. ◭
◮ Example (Capacitor filled with dielectric). Once we have Gauss’ law (4.10) in
the presence of a dielectric (or the integral form of it in (4.8)) solving electrostatic
problems involving dielectrics is no more difficult than what we have already done
in chapters 1 and 2. Let us recall that Gauss’ law involves the field D and the free
charge distribution. Consequently, just as we determined E earlier (in chapter 1)
from Gauss’ law we can now determine D. Furthermore, from the relation between
D and E in (4.12) we can then obtain the electric field as well.
−Q Q
D
n̂
Thus, we can write the displacement field between the two plates to be
where we have defined n̂ to be the unit vector normal to the plates as in Fig. 4.4
(say, along the z-axis). Here the negative sign arises because the direction of the
D field is opposite to the vector n̂.
The electric field between the plates can now be determined from relation
(4.12)
1 4πσ
E(r) = D(r) = − n̂. (4.23)
ǫ ǫ
Thus, as before, we see that the electric field continues to be a constant between
the plates. However, its strength is reduced from the case where there was no
dielectric between the plates. From (4.23) we determine the potential difference
(voltage) between the two plates to be
4πσ 4πd −1
V = |E|d = d= σA = Cdiel. Q. (4.24)
ǫ ǫA
Namely, we determine the capacitance in the presence of the dielectric to be
ǫA
Cdiel. = . (4.25)
4πd
In other words, the capacitance of the system increases in the presence of a di-
electric (compare with (2.89)). ◭
ǫ2 (ǫL ) ǫ1 (ǫR )
n̂
Figure 4.5: The interface of two distinct dielectric media with the
dashed curve representing the surface of a rectangular Gaussian vol-
ume.
∇ · D(r) = 4πρ(r),
E2
ǫ2
θ2
θ1
ǫ1
E1
r
b
r
q b
ǫ2
z=0
r′
ǫ1
q′ b
Figure 4.8: A point charge q in the region z > 0 with the image
charge q ′ in the region z < 0 for calculating the electric field in the
upper half plane.
Without loss of generality, we can assume that the point charge lies on
the z-axis at a height z = d from the interface. Unlike the case of a grounded
conducting plane, here we will have nonvanishing electric fields present in both the
dielectric media and, consequently, we need to calculate these in both the regions
z ≥ 0 and z ≤ 0. Let us recall from our earlier study involving the method of
images that we need an image charge in a region where we are not calculating
the electric field. As a result, since we have to calculate the potential and the
fields in both the regions z ≥ 0 and z ≤ 0, we need two sets of image charges -
one for each calculation. When we are calculating the field in the region z ≥ 0,
we need an image charge q ′ located at z = −d as shown in Fig. 4.8. On the
other hand, when we calculate the field in the region z ≤ 0, we need an image
charge in the region z ≥ 0. In fact, if we think for a moment, we realize that the
dielectric will be polarized because of the presence of the charge q. As a result,
as we have discussed in an earlier example, the effective charge seen in the region
z ≤ 0 will be modified. Consequently, in calculating the field in region z ≤ 0, we
can imagine an image charge present at z = d so as to give rise to an effective
charge q ′′ at that point (namely, the image charge has the value q ′′ − q at that
point), as shown in Fig. 4.9.
With these introductory remarks, the calculation is now straightforward.
Let us use cylindrical coordinates for our calculations. From Fig. 4.8, we see that
we can write the potential in the region z ≥ 0 as (see, for example, the discussion
4.2 Boundary conditions in dielectric 133
q ′′ b
ǫ2
z=0
r
ǫ1
Figure 4.9: For calculating the electric field in the lower half plane the
image charge can be chosen to lie on top of q leading to an effective
charge q ′′ .
Here ρ represents the radial coordinate on the plane. Similarly, the potential in
the region z ≤ 0 can be determined from Fig. 4.9 to be
1 q ′′ 1 q ′′
Φ(r; z ≤ 0) = = p . (4.38)
ǫ1 |r| ǫ1 ρ2 + (d − z)2
The components of the electric fields can now be calculated easily and have
the forms
" #
∂ 1 q(z − d) q ′ (z + d)
Ez (z ≥ 0) = − Φ(z ≥ 0) = + ,
∂z ǫ2 (ρ2 + (z − d)2 ) 23 3
(ρ2 + (z + d)2 ) 2
∂ 1 q ′′ (z − d)
Ez (z ≤ 0) = − Φ(z ≤ 0) = ,
∂z ǫ1 (ρ2 + (d − z)2 ) 32
" #
∂ 1 qρ q′ ρ
Eρ (z ≥ 0) = − Φ(z ≥ 0) = + ,
∂ρ ǫ2 (ρ2 + (z − d)2 ) 32 3
(ρ2 + (z + d)2 ) 2
∂ 1 q ′′ ρ
Eρ (z ≤ 0) = − Φ(z ≤ 0) = . (4.39)
∂ρ ǫ1 (ρ2 + (d − z)2 ) 32
Since there are no free surfaces charges, the boundary conditions (see (4.26)
134 4 Dielectrics
ǫ2 Ez (ρ, z = 0+ ) = ǫ1 Ez (ρ, z = 0− ),
d d
or, − 3 (q − q ′ ) = − 3 q ′′ ,
(ρ2 + d2 ) 2 (ρ2 + d2 ) 2
′ ′′
or, (q − q ) = q ,
Eρ (ρ, z = 0+ ) = Eρ (ρ, z = 0− ),
ρ (q + q ′ ) ρ q ′′
or, 3 = 3 ,
(ρ2 + d2 ) 2 ǫ2 (ρ2 + d2 ) 2 ǫ1
or, ǫ1 (q + q ′ ) = ǫ2 q ′′ . (4.40)
(ǫ1 − 1)
Pz (ρ, z = 0− ) = Ez (ρ, z = 0− )
4π
(ǫ1 − 1) 2ǫ1 qd
=− . (4.42)
4πǫ1 ǫ1 + ǫ2 (ρ2 + d2 ) 23
It follows from this that the net density of surface polarization charge is given by
(see (4.4))
σb = n̂ · P = Pz (ρ, z = 0+ ) − Pz (ρ, z = 0− )
(ǫ1 − ǫ2 ) qd
= . (4.43)
ǫ2 (ǫ1 + ǫ2 ) 2π(ρ2 + d2 ) 23
electric field along the z-axis. Thus, in the absence of the dielectric sphere we
have
where C is a constant and we are assuming the coordinate origin to coincide with
the center of the dielectric sphere. When we introduce the dielectric sphere, the
sphere will be polarized and would modify the electric field around and within
the sphere. However, asymptotically, the form of the potential in (4.45) would
continue to hold.
To determine the potential in the presence of the dielectric sphere let us note
that outside the dielectric sphere we simply have to solve the Laplace equation in
vacuum which has the form
and whose solutions can be written, in general, as (see (3.105) and note that the
present problem has azimuthal symmetry leading to m = 0)
∞
X
Φ> (r) = C − Er P1 (cos θ) + Aℓ r −(ℓ+1) Pℓ (cos θ), (4.47)
ℓ=1
where we have kept a linear term in r with the asymptotic condition (4.45) in
mind. (Here we have also used the fact that for Φ> (r) to satisfy the Laplace
equation, we must have A0 = 0.) Inside the dielectric sphere we also have to solve
the Laplace equation (there are no free charges inside the sphere and note also
that within the dielectric sphere ǫ is a constant so that it can be taken out of the
gradient operation)
We note that in writing these solutions, we have used the fact that the region inside
the sphere contains the origin and, consequently, should have regular solutions,
while outside the sphere the potential should fall off except for the asymptotic
behavior required by a constant electric field.
Now matching the solutions in (4.47) and (4.49) across the surface of the
sphere we obtain (see (4.30))
∞
X
+ Aℓ R−(ℓ+1) Pℓ (cos θ), (4.50)
ℓ=1
136 4 Dielectrics
which determines
B0 = C,
B1 R = −ER + A1 R−2 ,
Similarly, the condition on the normal derivatives in (4.30) at the boundary surface
gives (remember, there are no free charges on the surface and the space outside
the sphere is vacuum)
∂Φ< (r) ∂Φ> (r)
ǫ = ,
∂r r=R ∂r r=R
∞
X
or, ǫℓBℓ R(ℓ−1) Pℓ (cos θ) = − EP1 (cos θ)
ℓ=1
∞
X
− (ℓ + 1)Aℓ R−(ℓ+2) Pℓ (cos θ), (4.52)
ℓ=1
which determines
Aℓ = 0 = Bℓ , for ℓ ≥ 2,
B0 = C,
ǫ−1
A1 = ER3 ,
ǫ+2
3
B1 = − E, (4.54)
ǫ+2
so that we can write
3 !
ǫ−1 R
Φ> (r) = C − E 1− r cos θ,
ǫ+2 r
3
Φ< (r) = C − Er cos θ. (4.55)
ǫ+2
The electric fields can be determined from these and we note, in particular, that
inside the dielectric sphere the electric field is given by (z = r cos θ)
3
E< (r) = −∇Φ< (r) = Eẑ. (4.56)
ǫ+2
Namely, as a result of the external electric field the electric field present inside the
dielectric sphere is uniform. It has a reduced strength, but is along the z-axis like
the asymptotic field. This can be contrasted with the case of a conducting sphere
where there is no field inside the sphere. (Namely, even though the dielectric
is polarized, the polarization is not large enough to completely cancel the field
4.2 Boundary conditions in dielectric 137
inside.) This shows that the boundary value problems involving dielectrics are
solved much the same way by imposing appropriate boundary conditions at the
surface separating two dielectric media.
We note here that the problem of a spherical cavity inside an infinite
isotropic dielectric medium in the presence of a uniform electric field can also
be solved exactly in the same manner. In fact, let us note that the solution is
similar except that since the electric field inside a dielectric is reduced compared
to that in vacuum, we can obtain the solution simply by letting ǫ → 1ǫ . (Namely,
the boundary conditions, in this case, lead to an inverted field.)
◭
2ℓ
Similarly, the electric field along the z-axis in the region between the two
disks is obtained to be (|z| ≤ ℓ)
" ! !#
ℓ−z ℓ+z
E(z) = −2πP ẑ −p +1 − p −1
R2 + (ℓ − z)2 R2 + (ℓ + z)2
" #
z+ℓ z−ℓ
= 2πP −2 + p −p . (4.61)
R2 + (z + ℓ)2 R2 + (z − ℓ)2
σb = α cos θ, ρb = 0,
Magnetostatics
141
142 5 Magnetostatics
This can be taken as the defining relation for the magnetic field in
the sense that the magnetic field can be determined from the velocity
dependent part of the force experienced by a charged particle.
We have seen that the sources of electric fields are point charges
(or monopole charges or charge distributions). Correspondingly, we
can ask what are the sources of magnetic fields in nature. Exper-
imentally we know that there are no magnetic monopoles (yet) in
nature. The simplest sources of magnetic fields are known as mag-
netic dipoles. In fact, every magnet in nature including our own earth
has two magnetic poles (conventionally called north and south poles)
which cannot be separated into magnetic monopoles no matter how
hard we try. Thus, the center piece (the basic element) in the study
of magnetic phenomena is the magnet or the magnetic dipole. There
are some substances in nature which have permanent magnetic mo-
ments (or are permanent magnets). One can study the properties of
such materials and the effect of the magnetic fields, they produce,
on charged particles. However, there is an alternate mechanism for
producing magnetic fields which is what we will concern ourselves
with. It was observed in a series of experiments by Oersted, Ampere
etc. that a current produces a magnetic field. In fact, one can think
of closed current loops as magnetic dipoles which is the approach we
will take in studying magnetic phenomena.
5.2 Current
J = ρ v, (5.4)
where ρ is the volume charge density and v represents the net velocity
of the charge flow. The current density may or may not be uniform.
In either case, the current carried by a conductor (wire) is simply the
total charge flowing across a given cross-section of the conductor per
unit time
Z
I = ds · J. (5.5)
S
For a thin wire, the variation of J over the cross-sectional area may
not be appreciable in which case we can write
I = |J|A, (5.6)
∂ρ
or, + (∇ · J) = 0, (5.7)
∂t
where we have used Gauss’ theorem as well as the fact that the in-
tegral identity must hold for any volume V and, consequently, the
144 5 Magnetostatics
∂ρ
= 0, or, ∇ · J = 0. (5.8)
∂t
Namely, in such a case the system has reached equilibrium and as
much charge enters a given volume as leaves through a cross-sectional
area. In a later chapter, when we study time dependent phenomena
we will also analyze currents that are not steady state.
dℓ
| |
z
r1
r2
y
x
I dℓ′ × (r − r′ ) 1 ′ ′
′ J(r ) × (r − r )
dB(r) = = dV . (5.14)
c |r − r′ |3 c |r − r′ |3
are known as the Biot-Savart law and are observed to hold experi-
mentally. Comparing (5.15) with (1.11) we conclude that an electric
current is a source for the magnetic field just as an electric charge is a
source for the electric field. (One should be careful with the infinites-
imal form of the law. Namely, the current element here is assumed to
be a part of a current in a conductor. An independent element which
is not part of a current loop would violate the continuity equation.)
◮ Example (Magnetic field due to a long straight wire). As an application of the
Biot-Savart law let us determine the magnetic field, produced by an infinitely long
straight thin wire carrying a current I, at a perpendicular distance r from the wire.
For simplicity, let us assume that the wire lies along the z-axis as shown in Fig.
5.4 and that the base of the perpendicular from the point of observation on the
z-axis defines the origin.
We know from (5.15) that the magnetic field produced by the current can
be written as
I
I dℓ′ × (r − r′ )
B(r) = . (5.16)
c |r − r′ |3
Furthermore, it is obvious from the geometry that since dℓ′ is along the z-axis, the
magnetic field lines would be along the polar angle φ̂, field lines forming circles
surrounding the wire. This can be seen explicitly from the fact that, if we define
R = r − r′ = r − z ′ ẑ, (5.17)
I z
r
r′ θ ′
r
r−
Figure 5.4: The magnetic field due to a current carrying long straight
wire carrying current.
Z∞ Z∞
I dz ′ r 2I dz ′ r
|B| = 3 = 3
c (r 2 + z ′2 ) 2 c (r 2 + z ′2 ) 2
−∞ 0
∞
2Ir 1 z′
= 2I .
= × 2√ (5.19)
c r r 2 + z ′2 0 cr
Thus, using (5.18) and (5.19), we can write the magnetic field produced by the
current as
2I
B(r) = φ̂. (5.20)
cr
Incidentally, it is quite easy to see now that if there are two infinite parallel
wires separated by a distance r and carrying currents I1 and I2 respectively along
the same direction (say the z-axis), then there will be a force acting between the
two. For, we can think of the current I2 as producing a magnetic field which gives
rise to a magnetic force on the wire with current I1 and we can write (see (5.12))
Z
I1
F= dℓ × B. (5.21)
c
The magnetic field is along the polar direction φ̂ and the current (or dℓ) is along
the z-axis. Consequently, the force would be along the radial direction connecting
the two wires and would be attractive. Namely, in cylindrical coordinates
r̂ × φ̂ = ẑ, φ̂ × ẑ = r̂, ẑ × r̂ = φ̂, (5.22)
so that we can write the force (5.21) as
ZL/2
I1 2I2 2I1 I2 L
F = −r̂ dz = − r̂. (5.23)
c cr c2 r
−L/2
Here we have assumed the two wires to be of length L (each) which is to be taken
to L → ∞ at the end. Therefore, we obtain the force per unit length between the
two currents to be
F 2I1 I2
= − 2 r̂. (5.24)
L c r
148 5 Magnetostatics
The force is clearly attractive. However, if we reverse the direction of one of the
currents, then, the direction of the force would reverse as well and leads to the
familiar fact that two parallel currents attract while two anti-parallel currents
repel each other. ◭
◮ Example (Magnetic field due to a circular current loop). The important result
to note from the previous example is that a straight wire carrying a steady current
produces concentric circular magnetic fields around the axis of the wire, whose
strength falls off inversely as the radius of the circle. Let us next analyze the
magnetic field produced by a circular current loop of radius R as shown in Fig.
5.5. We assume that the the current loop is in a plane perpendicular to the z-axis
and that the current moves in a clockwise direction when seen from below. The
magnetic field due to this current at an arbitrary point is difficult to calculate.
Therefore, we will calculate the magnetic field at any point on the z-axis which
has a simpler form.
r
y
φ R
x
Once again, we use Biot-Savart law (5.15) and choosing the center of the
current loop to be the origin of our coordinate system, we have
I
I dℓ × (r − R)
B(r) = . (5.25)
c |r − R|3
The simplest way to evaluate this is to note that (dℓ is orthogonal to R)
dℓ = Rdφ φ̂, R̂ = cos φ x̂ + sin φ ŷ,
where we have used the usual rules for cross products, namely, φ̂×ẑ = R̂, R̂× φ̂ =
ẑ. It follows now that
Z2π
I R dφ [z(x̂ cos φ + ŷ sin φ) + Rẑ]
B(z) = 3
c (R2 + z 2 ) 2
0
2πIR2
= 3 ẑ. (5.27)
c(R2 + z 2 ) 2
5.3 Force on a current due to a magnetic field 149
Thus, along the axis of the loop, the magnetic field is completely parallel to
the axis. At the center of the loop (z = 0), we note from (5.27) that the magnetic
field has the value
2πI
B(z = 0) = ẑ, (5.28)
cR
while very far away, namely, when z >> R, we have
2πIR2
B(z >> R) ≈ ẑ, (5.29)
cz 3
which is reminiscent of the electric field due to a dipole. Thus, we suspect that
a circular current somehow produces a magnetic field which has dipole charac-
teristics. We will see this shortly, but let us note that this suggests that the
magnitude of the magnetic dipole moment is proportional to the current times
the area enclosed by the current loop.
Figure 5.6: A long solenoid of radius R with n turns per unit length.
From (5.27), we can also calculate the magnetic field due to a long solenoid
along its axis. Let us consider a solenoid of radius R with n turns per unit length
with the z-axis representing the axis of the solenoid (see Fig. 5.6). In a length
interval dz there will be n dz loops of wire each producing a magnetic field as
derived in (5.27). Consequently, the total magnetic field produced by the solenoid
along the axis is given by
Z∞
2πnIR2 dz
B= ẑ 3
c (R2 + z 2 ) 2
−∞
∞
2πnIR2 1 z
= ẑ × 2 √
c R R2 + z 2 −∞
4πnI
= ẑ. (5.30)
c
Namely, the magnetic field is a constant along the axis of an infinitely long solenoid
determined completely by the current and the number of turns per unit length.
In fact, even though we have not shown this, the magnetic field is really constant
150 5 Magnetostatics
at any point inside the solenoid. Long solenoids are often used to produce and
maintain constant magnetic fields over short distances. ◭
We have already seen that the Biot-Savart law (5.15) expresses the
magnetic field in terms of the current density as
Z
1 J(r′ ) × (r − r′ )
B(r) = d3 r ′ , (5.31)
c |r − r′ |3
∇ · B(r) = 0. (5.33)
∇ · B(r) = 0,
4π
∇ × B(r) = J(r), (5.36)
c
which should be compared with the laws of electrostatics in vacuum,
∇ · E(r) = 4πρ(r),
∇ × E(r) = 0. (5.37)
4πI
= . (5.38)
c
The second of these relations (in either the differential or the integral
form) is known as Ampere’s law and says that the line integral of the
magnetic field around any closed loop is proportional to the current
flowing through the cross-sectional area of the loop. It is useful in
calculating magnetic fields for problems with symmetry, much like
the Gauss’ law in calculating electric fields.
152 5 Magnetostatics
I
z
Figure 5.7: The magnetic field due to a long straight wire carrying
current.
enough symmetry so that to begin with, we know that the magnetic field at any
point would be in the tangential direction to the circle drawn around the axis of
the wire. Furthermore, the magnitude of the magnetic field would be the same at
points (perpendicularly) equidistant from the wire. With this information, let us
draw an “Amperian” loop of radius r around the axis of the wire (clockwise as
seen from below). Then, according to Ampere’s law (5.38), we have
I
4πI
dℓ · B = ,
c
4πI
or, |B(r)| 2πr = ,
c
2I
or, |B(r)| = , (5.39)
cr
and the magnetic field is along the direction of the polar angle. This is the result
we had obtained earlier in (5.20) by explicitly evaluating the integral in the Biot-
Savart law. ◭
hand, we know that the magnetic field has vanishing divergence (see
(5.36)). Consequently, in this case, we expect that we can write it as
the curl of a vector and, in fact, we have already seen in (5.32) that
this is true, namely,
Z
1 J(r′ ) × (r − r′ )
B(r) = d3 r ′
c |r − r′ |3
Z
1 3 ′ ′ 1
=− d r J(r ) × ∇
c |r − r′ |
Z
1 J(r′ )
=∇× d3 r ′ . (5.40)
c |r − r′ |
where α(r) is an arbitrary scalar function would give rise to the same
magnetic field since the curl of a gradient vanishes. This is the first
manifestation of what we would see later as the gauge invariance
of Maxwell’s equations. For the present, let us simply note that
the vector potential A obtained in (5.42) appears to be unique only
because it also satisfies the condition
Z
1 3 ′ ′ 1
∇ · A(r) = d r J(r ) · ∇
c |r − r′ |
Z
1 1
= d3 r ′ ∇′ · J(r′ ) = 0, (5.44)
c |r − r′ |
154 5 Magnetostatics
which follows from the fact that the currents are steady state cur-
rents in magnetostatics (the surface terms arising from integration
by parts are assumed to vanish for localized currents as in (5.35)).
Such conditions, as we will see later, are called gauge conditions and
are necessary when dealing with a system of equations which has
gauge invariance.
◮ Example (Vector potential of a long straight wire). As an example of calcula-
tions of the vector potential, let us consider again the example of an infinitely
long straight wire carrying a current I along the z-axis as shown in Fig. 5.8. The
vector potential is defined in (5.42) to be
I z
r
r′ θ ′
r
r−
Figure 5.8: Vector potential for a long straight wire carrying current.
Z Z
1 J(r′ ) I dr′
A(r) = d3 r ′ = . (5.45)
c |r − r′ | c |r − r′ |
In the present case, the current is along the z-axis. Consequently, only the z-
component of the vector potential will be nonzero. If we define the perpendicular
distance of a point from the wire as r, then the vector potential takes the form
(here we assume that the point at which the field is being calculated lies in the
z = 0 plane with the origin at the foot of the perpendicular to the wire)
Z∞
I dz ′
Az (r) = √
c r2+ z ′2
−∞
Z∞
2I dz ′
= √
c r2+ z ′2
0
2I p ∞
= log r 2 + z ′2 + z ′
c 0
2I
=− log r + constant, (5.46)
c
where we have used the standard integration formula (see Gradshteyn and Ryzhik
2.261). Note that the constant of integration is a divergent constant, much like the
5.6 Multipole expansion 155
case of the scalar potential for an infinitely long wire carrying charge. As in the
example in electrostatics the constant in (5.46) does not matter in the calculation
of physical fields.
From the form of p the vector potential (5.46), we obtain the magnetic field
to be (in this case, r = x2 + y 2 )
B(r) = ∇ × A(r)
∂Az ∂Az
= x̂ − ŷ
∂y ∂x
2I (−y x̂ + x ŷ) 2I r(− sin φ x̂ + cos φ ŷ)
= =
c r2 c r2
2I
= φ̂. (5.47)
cr
This is, of course, the result we had obtained earlier for the magnetic field by
directly evaluating the integral in the Biot-Savart law (see (5.20)). ◭
Let us consider a small circular current loop with the center at the
origin of the coordinate system (see Fig. 5.9). We note from (5.42)
that we can write the vector potential as
r
r′
r−
r′
Figure 5.9: The vector potential for a small circular current loop.
Z I
1 3 ′ J(r′ ) I dr′
A(r) = d r ′
= , (5.48)
c |r − r | c |r − r′ |
where we are assuming that the wire is thin so that we can assume
the current density to be uniform. We note that, as in the case of
electrostatics (see, for example, (2.40) and discussions there), we can
expand the denominator in (5.48) for r ≫ r ′ , so that the vector
156 5 Magnetostatics
5.7 Magnetization
also have materials where the magnetic moments due to each of the
electron currents in an atom do not quite cancel. In such a case, every
atom in the material may have a net magnetic dipole moment. How-
ever, the magnetic moments of different atoms in the material may be
randomly distributed, leading to a zero net magnetic moment for the
material. In the presence of an external magnetic field, once again,
the magnetic moments would orient themselves. However, in such a
case, the magnetic moments align parallel to the external field and
this is the behavior of paramagnetic materials. There are, of course,
also ferromagnetic materials. Here, the atoms do have a net magnetic
moment like paramagnetic materials. Furthermore, in such materials
the atoms are quite close together and the magnetic moments are
aligned so that inside the material there are domains with large mag-
netic moments which, however, are randomly distributed and can give
rise to a net zero magnetic moment. In the presence of a magnetic
field, however, they all align to give rise to a large magnetic moment
which does not vanish even when the magnetic field is switched off.
(This phenomenon is known as hysteresis.)
Just as in the case of dielectric materials, for a magnetic ma-
terial we can define a magnetic dipole moment M per unit volume
(analogous to polarization, see for example, (2.57) and discussions
there), which is also known as the magnetization of the material.
Experimentally, it is observed, for both diamagnetic and paramag-
netic materials, that in the presence of an external magnetic field the
magnetization is linearly related to the applied magnetic field so that
we can write
M ∝ B. (5.54)
where n̂ represents the unit vector normal to the surface and we have
used the identity from vector calculus that for any arbitrary vector
A,
Z Z
3
d r ∇ × A = ds × A. (5.56)
V S
The relation (5.55) is quite interesting, for it says that the vector
potential produced by a magnetized material can be thought of as
due to both a volume current density and a surface current density
of the forms
J(r) = c ∇ × M(r),
Jσ (r) = −c n̂ × M(r), (5.57)
B(r) = ∇ × A(r)
Z
3 ′ ′ 1
= −∇ × d r M(r ) × ∇
|r − r′ |
Z
3 ′ ′ 2 1
=− d r M(r ) ∇
|r − r′ |
′ 1
−(M(r ) · ∇)∇
|r − r′ |
Z Z
M(r′ ) · (r − r′ )
= d r M(r ) 4πδ (r − r ) − ∇ d3 r ′
3 ′ ′ 3 ′
|r − r′ |3
Z
M(r′ ) · (r − r′ )
= 4π M(r) − ∇ d3 r ′ . (5.59)
|r − r′ |3
4π
∇ × B = 4π ∇ × M = J, (5.61)
c
where J represents the volume current density due to the current
loops in the magnetized material.
5.8 Magnetic field intensity 161
Let us now ask what will be the modifications in the laws of magneto-
statics (5.36) in the presence of a magnetic material. The discussion
is completely parallel to the case of the laws of electrostatics in the
presence of a dielectric material. First, we note that when we have
a magnetic material present we will have two kinds of currents, one
that is maintained by batteries etc. and the other due to the inter-
nal motion of bound electrons inside the magnetic material. Thus,
analogous to the case of the dielectrics, let us refer to them as the
free and the bound currents respectively. In this case, Ampere’s law
would say that the magnetic field integrated around any closed curve
will be related to the total current,
I
4π
dℓ · B = (I + Ib ), (5.62)
C c
where the right hand side represents the total current through the
surface bounded by the curve, with I denoting the free current and
Ib representing the current due to the bound electrons of the magnetic
material completely in analogy with the case of electrostatics in the
presence of a dielectric. We can, of course, write
Z
Ib = ds · Jb , (5.63)
S
I
4π
or, dℓ · H = I, (5.65)
C c
where we have defined (analogous to the case of electrostatics) a new
field
H = B − 4π M. (5.66)
The new field, H, is known as the magnetic field intensity (also
known simply as the magnetic field) and is the analog of the electric
displacement field. Ampere’s law (5.65), in the presence of a magnetic
material, is written in terms of this field and the right hand side, in
this case, involves only the free currents in a magnetic material. The
differential form of Ampere’s law can now be obtained using Stokes’
theorem and takes the form
4π
∇×H= J, (5.67)
c
where J represents the free current density in the system. Thus, the
laws of magnetostatics, in the presence of a magnetic material, take
the forms
∇ · B = 0,
4π
∇×H= J, (5.68)
c
where the fields are related as
H = B − 4π M, or, B = H + 4π M. (5.69)
Since the magnetization is parallel to B all the vectors B, M,
and H are parallel. Let us define
M = χm H. (5.70)
The constant of proportionality χm is known as the magnetic sus-
ceptibility of the material. For diamagnetic materials it is negative,
while it is positive for paramagnetic materials. Furthermore, for both
diamagnetic and paramagnetic materials, the magnitude of the mag-
netic susceptibility is quite small (of the order of 10−5 − 10−4 ) which
should be compared with the electric susceptibility which is positive
and is much larger in magnitude. Let us also note that with this
definition, we can now write
B = (1 + 4πχm ) H = µ H, (5.71)
where we have identified
µ = 1 + 4πχm . (5.72)
This is known as the permeability of a magnetic material.
5.9 Boundary condition 163
L R
R z
∞
X
Φm,> (r) = Cℓ r −(ℓ+1) Pℓ (cos θ). (5.78)
ℓ=0
We can now match the boundary condition for the tangential components
of the H fields. The tangential component is along the θ direction giving (we can
also equivalently write this as Φm,> |r=R − Φm,< |r=R = 0 as in (4.50))
This determines
Cℓ = Aℓ R2l+1 , ℓ ≥ 1. (5.80)
The matching of the normal component of the B field can be done as follows.
We recall the relation (5.69) between the B and the H fields. Since there is no
magnetization outside the sphere, we can write (remember that B = H + 4πM
and that H = −∇Φm )
Cℓ = 0 = A ℓ , for ℓ ≥ 2,
C0 = 0,
4πM
A1 = C1 R−3 = , (5.83)
3
so that, we can write
3
4πM R
Φm,> (r) = r cos θ,
3 r
4πM
Φm,< (r) = r cos θ. (5.84)
3
Here we have used the fact that even though A0 is an undetermined constant, a
constant term in the potential does not influence the fields and accordingly we
have chosen to set it to zero. It follows from this that
3
4πM R
H> (r) = −∇Φm,> (r) = (2r̂ cos θ + θ̂ sin θ),
3 r
166 5 Magnetostatics
4πM
H< (r) = −∇Φ< (r) = − ẑ. (5.85)
3
It is clear from this that outside the sphere, the magnetic field behaves like that
of a magnetic dipole of moment (recall that 3(ẑ · r̂)r̂ − ẑ = (2r̂ cos θ + θ̂ sin θ))
There are several things to note from this relation. First of all,
we have seen in electrostatics that
I
dℓ · E = 0. (5.88)
1 ∂B
or, ∇ × E = − . (5.89)
c ∂t
This is the differential form of Faraday’s law and shows that when
we have time dependent phenomena, electric and magnetic fields es-
sentially become coupled.
5.11 Inductance
dI
= −L , (5.90)
dt
where we have defined
Z
1 ∂
L= ds · B. (5.91)
c ∂I
S
Then, denoting by Φm,i the total magnetic flux through the ith circuit
due to the currents in all the circuits, we have
n
X
Φm,i = Φm,ij . (5.93)
j=1
Here Φm,ij represents the magnetic flux through the ith circuit due
to the current in the jth circuit,
Z
Φm,ij = dsi · Bj . (5.94)
Si
It is clear now from Faraday’s law that the emf in the ith circuit due
to time dependent currents in the circuits takes the form (recall that
Φm,ij ∝ Ij )
1 dΦm,i
(emf)i = −
c dt
n
1 X ∂Φm,ij dIj
=−
c ∂Ij dt
j=1
n
X dIj
=− Mij , (5.95)
dt
j=1
Maxwell’s equations
171
172 6 Maxwell’s equations
and must hold for any time dependent sources. Furthermore, Gauss’
law has the general form
∇ · D = 4πρ. (6.4)
S1
S2
know that
I Z Z
4π
dℓ · H = ds · (∇ × H) = ds · J. (6.7)
c
S S
The first relation follows from the fact that there is no conduction
current within the dielectric. The two relations in (6.8) are, there-
fore, inconsistent. On the other hand, the additional term in (6.5)
will remove this inconsistency since inside the dielectric there is a dis-
placement field (even though there is no conduction current) whose
time rate of change provides the effect of a current when integrated
over a surface, namely, (here we will assume the surface to be closed)
Z Z
1 ∂D ∂ 1
ds · = ds · D
4π ∂t ∂t 4π
S S
Z Z
d 1 3 d
= d x ∇ · D = d3 x ρ
dt 4π dt
V V
dQ
= = ID . (6.9)
dt
We know that a current produces a magnetic field and so, in
keeping with this concept Maxwell identified the new term added to
the right hand side of Ampere’s law with a current (since it con-
tributes to the magnetic field) known as the displacement current,
1 ∂D
JD = . (6.10)
4π ∂t
The simplest way to think of this current is to recall that in the
presence of an applied field (including an alternating one) the charge
centers in a dielectric are displaced leading to the effect of a current.
174 6 Maxwell’s equations
Such currents are, however, different from the usual conduction cur-
rents that we are used to, since the charges never really leave the
nucleus (or the atom). They are known as displacement currents
since they arise from a displacement of charge centers. (Sometimes,
they are also known as polarization currents. Actually, during the
time of Maxwell, it was believed that space is filled with ether which
acts like a dielectric and Maxwell himself believed that this effect
should arise in ether as well.) Maxwell’s proposal (6.5), of course,
had a purely theoretical origin and the experimental verification of
this would not come until two decades later in the experiments of
Hertz. The main difficulty in the experimental verification lies in the
fact that in a conductor, where we know how to measure a current,
the conduction current is overwhelmingly dominant over the displace-
ment current unless the frequency of the time dependent current is
extremely high. However, Hertz’s experiments clearly demonstrated
the existence of a displacement current in dielectrics and the validity
of Maxwell’s modification of Ampere’s law in (6.5).
Together with this modification, we can write all the laws of elec-
tricity and magnetism for general time dependent fields (and sources)
as
∇ · D = 4πρ,
∇ · B = 0,
1 ∂B
∇×E=− ,
c ∂t
4π 1 ∂D 4π
∇×H= J+ = (J + JD ) . (6.11)
c c ∂t c
These are the fundamental laws of electrodynamics and they are
known as Maxwell’s equations. They hold for both time dependent
as well as time independent fields and sources. As we see, these
are coupled differential equations (incidentally, one can also write
the integral forms for these equations using Gauss’ and Stokes’ the-
orems, but we would not go into this), which can be checked to be
self-consistent and which become decoupled in the static limit. The
equations (6.11), of course, have to be supplemented further by the
continuity equation (6.3) as well as various other equations describing
the effects of the medium, namely,
∂ρ
+ ∇ · J = 0,
∂t
J = σE,
6.2 Plane wave solution 175
(
q E + 1c v × B for point charges,
F= (6.12)
ρE + 1c J × B for continuous distributions.
Here, we note that the second of the supplementary equations in
(6.12) is simply Ohm’s law with σ representing the conductivity of
the conductor, while the third describes the Lorentz force law (for
continuous distributions, the third condition corresponds to the force
density).
Equations (6.16) and (6.17) show that both the electric field as
well as the magnetic field satisfy the three dimensional wave equation,
with the velocity (speed) of propagation given by (see (3.48))
c
v=√ . (6.18)
ǫµ
In particular, since ǫ = 1 = µ in vacuum we see that the speed of
propagation of these waves in vacuum coincides with the speed of
light. This was the first evidence that light waves arise because of
time dependent electric and magnetic fields or that light waves are
simply electromagnetic waves.
To further understand the behavior of these waves, let us con-
sider for simplicity a plane wave solution for the electric field. A
plane wave by definition is a wave where the wave variable has the
same phase at any point on the wavefront which is an infinite plane.
Thus, we note that a plane wave solution of the equation involving
the electric field (6.16) will have the form
E(x, t) = E(0) e∓iωt+ik·x , (6.19)
where E(0) is a constant vector, provided the parameters ω and k
satisfy
ω c
= √ = v, (6.20)
|k| ǫµ
which is precisely the relation satisfied by traveling waves. Conven-
tionally, we say that ω = 2πν represents the angular frequency of
the wave, while |k| = 2π λ with λ the wavelength denotes the wave
number. This is seen by noting that the direction of propagation of
the wave is along k and along that direction, points separated by a
distance of λ (the wavelength) are in phase. With this identification,
(6.20) leads to the familiar relation of wave phenomena, namely,
c
νλ = v = √ . (6.21)
ǫµ
As we have noted, we are analyzing plane wave solutions. The
wavefronts (points where the phases are the same), in this case, are
given by
∓ωt + k · x = constant, (6.22)
which, for a fixed time, correspond to planes of infinite extent satis-
fying
k · x = constant. (6.23)
6.2 Plane wave solution 177
∓ ωt + |k|ξ = constant,
dξ ω
or, =± = ±v, (6.24)
dt |k|
showing that the velocity of propagation in one case is along k while
it is in the opposite direction in the other case.
Let us emphasize here that the wave solutions that we have con-
structed in (6.19) are known as monochromatic plane waves since they
involve only a single frequency ω (also known as harmonic waves).
A monochromatic plane wave can consist of a linear superposition
of both forward and backward moving waves of the same frequency.
A general solution, on the other hand, would involve a sum (or an
integral) over distinct frequencies as well, in which case it is not a
monochromatic wave. The velocity of propagation for a monochro-
matic wave is known as the phase velocity, while for a wave packet
consisting of distinct frequencies the velocity of propagation is known
as the group velocity. The two velocities can, in general, be different.
Furthermore, let us note that while the four Maxwell’s equations
(6.14) lead to the wave equations (6.16) and (6.17) (for E and B
fields), the two wave equations are not equivalent to the set of four
Maxwell’s equations. (Namely, the solutions of (6.16) and (6.17)
would not automatically satisfy all the equations in (6.14).) There-
fore, a plane wave solution of Maxwell’s equations has to satisfy fur-
ther conditions. Thus, for example, we see from the first equation of
(6.14)
∇ · E = 0, (6.25)
that the plane wave solution for the electric field (6.19) must satisfy
k · E = 0. (6.26)
k · E = 0 = k · B. (6.28)
Namely, both the electric and the magnetic fields are perpendicular
to the direction of propagation of the wave which shows that electro-
magnetic waves are transverse waves (unlike sound waves which are
longitudinal). It is also worth emphasizing here that both the electric
and the magnetic fields are real quantities. In writing a solution in
the form (6.27), the assumption is that the electric and the magnetic
fields correspond to either the real or the imaginary parts of the com-
plex solutions which would respectively give cosine or sine solutions.
Furthermore, from the third equation in (6.14)
1 ∂B
∇×E=− , (6.29)
c ∂t
we obtain for the forward moving wave,
ω
k×E= B. (6.30)
c
k
B
Figure 6.2: The electric and the magnetic fields as well as the direc-
tion of propagation as defining an orthogonal system.
In other words, the electric and the magnetic fields are not only
perpendicular to the direction of propagation, but they are orthogonal
to each other as well (see, for example, Fig. 6.2). It is also clear from
6.2 Plane wave solution 179
(6.20) and (6.30) that if the wave is propagating along the z-axis,
then (B = µH),
r
|Ex | |Ey | ωµ µ µ
= = =√ = , (6.31)
|Hy | |Hx | c|k| ǫµ ǫ
x x x
z z z
E E E
y y y
In this case, it is easy to see that not only the magnitude of the electric
field, but also its direction changes with time tracing out an ellipse
in the x − y plane and we say that the wave is elliptically polarized.
(0) (0)
Finally, if |E1 | = |E2 | and the magnitude of the relative phase
between the two components is π2 , then we can write from (6.33)
(0)
E(z, t) = Re |E1 |(x̂ ± iŷ)e−iωt+ikz
(0)
= |E1 | (x̂ cos(ωt − kz) ± ŷ sin(ωt − kz)) . (6.37)
(0)
In this case, the magnitude of the electric field is a constant (|E1 |),
but the direction changes with time tracing out a circle in the x − y
plane and we say that the wave is circularly polarized. If the rotation
of the electric field is clockwise to an observer facing the incoming
wave, the wave is said to be right circularly polarized. For an op-
posite rotation, the wave is correspondingly known as left circularly
polarized. (Namely, the two terms x̂±iŷ in (6.37) denote respectively
left and right circular polarizations, see Fig. 6.3.) The different po-
larizations are sketched in Fig. 6.3.
boundary surface separating the two media. In such a case, the first
two of Maxwell’s equations in (6.11)
∇ · D = ∇ · (ǫE) = 0 = ∇ · B, (6.38)
hold in the two dielectric media and tell us that the normal compo-
nents of the electric displacement field and the magnetic field must
be continuous across the boundary. Specifically, we have
L R
In the limit of vanishing width of the rectangular loop, the right hand
side of the equation vanishes because the area enclosed by the loop
does. In the same limit, the left hand side simply gives the difference
in the tangential components of the electric field on the two sides
multiplied by the horizontal length of the curve. Therefore, in this
limit, (6.41) leads to
EL = Einc + ERefl ,
(0) (0)
Einc = Re E1 e−iωt+ikL z = E1 cos(ωt − kL z),
(0) (0)
ERefl = Re E2 eiωt+ikL z = E2 cos(ωt + kL z),
(0) (0)
ER = Etrans = Re E3 e−iωt+ikR z = E3 cos(ωt − kR z). (6.44)
Here, we are assuming that the vector amplitudes are real for simplicity. Further-
more, for these to represent solutions of Maxwell’s equations in the two media
with dielectric constants and permeabilities (ǫL , µL ) and (ǫR , µR ) respectively, we
must have (see (6.20))
√ √
ω ǫL µL ω ǫR µR
kL = = nL k, kR = = nR k, (6.45)
c c
184 6 Maxwell’s equations
z
kL kR
z=0
where k represents the wave number in vacuum. (Incidentally, the fact that the
waves are all moving along the z-axis follows from the fact that the tangential
components of the fields have to be continuous across the boundary. The copla-
narity of the waves is a general property that we will see in more detail in the
next example where we consider reflection and refraction for oblique incidence.)
Furthermore, from (6.30)
ω
k×E= B, (6.46)
c
BL = Binc + BRefl ,
√ √ (0)
Binc = ǫL µL ẑ × Einc = ǫL µL ẑ × E1 cos(ωt − kL z),
√ √ (0)
BRefl = − ǫL µL ẑ × Erefl = − ǫL µL ẑ × E2 cos(ωt + kL z),
√ √ (0)
BR = Btrans = ǫR µR ẑ × Etrans = ǫR µR ẑ × E3 cos(ωt − kR z). (6.47)
We note here that the electric and the magnetic fields are in the plane
orthogonal to the direction of propagation which is along the z-axis. Consequently,
there are no components of these fields normal to the boundary plane. Matching
the tangential components of the electric and the magnetic fields at z = 0 (see
(6.42) and (6.43)), we obtain from (6.44) and (6.47)
q
2 µǫLL
(0) (0)
E3 = q q E1 . (6.49)
ǫL ǫR
µL
+ µR
◮ Example (Oblique incidence). Let us next consider the case where a plane elec-
tromagnetic wave is incident on a boundary surface separating two dielectric me-
dia at an oblique angle (see Fig. 6.6). We choose the boundary surface to be the
plane z = 0 and assume that the wave is incident at an angle θi (with the z-axis).
Without loss of generality, we can assume the plane of incidence to be the x − z
plane (plane of incidence is defined to be the plane containing the direction of
propagation of the incident wave k̂i and the normal to the boundary surface n̂
which corresponds to ẑ in the present example). Thus, as before, we can write
the incident wave to have the form (see (6.44))
θi
θr z
θt
x
z=0
(0)
Einc = E1 cos(ωt − ki · x), (6.52)
186 6 Maxwell’s equations
where, as we have noted, k̂i denotes the direction of propagation of the incident
wave which we have chosen to be in the x − z plane so that it has the form (see
Fig. 6.6)
As we have discussed earlier, there will also be a reflected wave in the first
medium and a transmitted wave in the second, with the forms (see (6.44))
(0)
Erefl = E2 cos(ωt − kr · x),
(0)
Etrans = E3 cos(ωt − kt · x), (6.54)
where k̂r and k̂t denote respectively the directions of propagation for the reflected
as well as the transmitted waves. (The sign of the direction of propagation for
the reflected wave has been absorbed into the definition of k̂r .) The waves need
not a priori be coplanar. However, we know that for them to satisfy Maxwell’s
equations, the magnitudes of the wave vectors must satisfy (see (6.20) and (6.32))
nL ω nR ω
|ki | = |kr | = = nL |k|, |kt | = = nR |k|, (6.55)
c c
where we have defined |k| to represent the wave number in free space.
Given the electric fields, we can also determine the magnetic fields from
(6.30) (or (6.46)) and we have
ki
Binc = × Einc = nL k̂i × Einc ,
|k|
kr
Brefl = × Erefl = nL k̂r × Erefl ,
|k|
kt
Btrans = × Etrans = nR k̂t × Etrans . (6.56)
|k|
With these fields, we can now match the boundary conditions (6.39), (6.42)
and (6.43). Let us note, for example, that the tangential components of the electric
fields have to be continuous across the boundary z = 0, namely,
tang.
(0) (0)
E1 cos(ωt − ki · x) + E2 cos(ωt − kr · x)
z=0
tang.
(0)
= E3 cos(ωt − kt · x) . (6.57)
z=0
Such a condition has two aspects. First, of course, the vector amplitudes have
to satisfy some conditions, but more important is the fact that the phases should
match as well. For example, matching the phases in (6.57) leads to
From the fact that ki lies in the x − z plane, it now follows from (6.58) that all
the three wave vectors must also lie in the x − z plane. This can be seen simply
as follows. The matching condition (6.58) explicitly gives
which leads to the fact that kry = 0 = kty . In other words, all the three plane
waves have to be coplanar (have to lie on the same plane). Therefore, from the
6.3 Boundary conditions 187
geometry of the problem under study (see Fig. 6.6) we note that the unit vectors
have the forms
where θi , θr and θt denote respectively the angles of incidence, reflection and trans-
mission.
In addition to determining that the waves have to be coplanar, the matching
of the phases at the boundary (6.59) also requires that
kix = krx ,
and similarly,
kix = ktx ,
or, nL |k| sin θi = nR |k| sin θt ,
sin θi nR
or, = . (6.63)
sin θt nL
Equation (6.62) describes the familiar law (from optics) that the angle of incidence
is equal to the angle of reflection, while (6.63) is the Snell’s law for refraction.
Let us next come to the vector amplitudes. There are two independent cases
to analyze and let us start with the simple case where the electric field for the
incident wave is normal to the plane of incidence, namely, let us assume that it
lies along the y-axis (so that the electric field has no normal component). Then,
it follows from the boundary condition (6.57) that the electric fields of all the
three waves will lie along the y-axis and the boundary condition for the normal
components of the displacement field will hold automatically. In this case, we
have chosen an incident wave polarized along the y-axis. The vector amplitudes
for the magnetic fields can now be easily calculated from (6.56) and (6.60)
(0) (0) (0)
B1 = nL E1 k̂i × ŷ = nL E1 (− cos θi x̂ + sin θi ẑ),
(0) (0) (0)
B2 = nL E2 k̂r × ŷ = nL E2 (cos θi x̂ + sin θi ẑ),
(0) (0) (0)
B3 = nR E3 k̂t × ŷ = nR E3 (− cos θt x̂ + sin θt ẑ). (6.64)
Furthermore, since for most optically transparent material µ ≈ 1, we will use such
a value. In such a case, matching the tangential components of the electric and
the magnetic fields at z = 0, we obtain (the normal components of the magnetic
fields are automatically continuous which can be seen using Snell’s law)
188 6 Maxwell’s equations
tan θt − tan θi
= ,
tan θt + tan θi
(0)
E3 2nL cos θi 2 cos θi
(0)
= =
E1 nL cos θi + nR cos θt cos θi + sin θi cot θt
2 tan θt
= . (6.66)
tan θt + tan θi
The other case to analyze is when the electric field is polarized parallel to
the plane of incidence. Since the fields have to be perpendicular to the direction of
propagation, we can choose (this choice also makes the normal components of the
(0) (0)
displacement field continuous across the boundary, namely, n2L ẑ · (E1 + E2 ) =
(0)
n2R ẑ · E3 )
(0) (0)
E1 = E1 (cos θi x̂ − sin θi ẑ),
(0) (0)
E2 = −E2 (cos θi x̂ + sin θi ẑ),
(0) (0)
E3 = E3 (cos θt x̂ − sin θt ẑ), (6.67)
which will give rise to the magnetic fields (see (6.56))
(0) (0)
B1 = nL E1 ŷ,
(0) (0)
B2 = nL E2 ŷ,
(0) (0)
B3 = nR E3 ŷ. (6.68)
In this case, matching the tangential components of the electric and the magnetic
fields across the boundary we obtain (we are assuming µ ≈ 1)
(0) (0) (0)
E1 − E2 cos θi = E3 cos θt ,
(0) (0) (0)
nL E1 + E2 = nR E3 . (6.69)
(0) (0)
Once again, we can solve for E2 and E3 using Snell’s law and we obtain
(0) sin θi cos θi
E2 nR cos θi − nL cos θt sin θt
− cos θt
(0)
= = sin θi cos θi
E1 nR cos θi + nL cos θt sin θt
+ cos θt
tan(θi − θt )
= ,
tan(θi + θt )
6.3 Boundary conditions 189
(0)
E3 2nL cos θi 2 cos θi
(0)
= = sin θi cos θi
E1 nR cos θi + nL cos θt sin θt
+ cos θt
2 cos θi sin θt
= . (6.70)
sin(θi + θt ) cos(θi − θt )
nR sin θi
= = tan θi . (6.72)
nL sin θt
The incident angle for which this holds is known as the Brewster’s angle. In
general, of course, a wave can be decomposed into a sum of waves polarized
parallel and perpendicular to the plane of incidence. What this analysis shows is
that at the Brewster’s angle of incidence, the component (of the field) polarized
parallel to the plane will not be reflected and, consequently, the reflected wave
will be polarized perpendicular to the plane of incidence.
The other important observation from all of the above analysis is that if
nL > nR , then, from Snell’s law (6.63) we obtain
sin θi nR
= ,
sin θt nL
or, θt > θi . (6.73)
π
It follows, therefore, that there is some angle of incidence for which θt = 2
. Let
us call this θint so that we have from Snell’s law
nR
sin θint = . (6.74)
nL
π
Let us note that for θi = θint , we have θt = 2
so that at this angle of incidence
(0)
Etrans = E3 cos(ωt − kt · x)
(0) (0)
= E3 cos(ωt − |kt |(sin θt x + cos θt z)) = E3 cos(ωt − |kt |x).
(6.75)
so that, we can identify the energy density stored in the static mag-
netic field as
1 1 µ 2
wmag = H·B= B·H= H . (6.81)
8π 8π 8π
Thus, we can write the total energy density stored in the electromag-
netic field, in the static case, as given by
D·E+B·H
w = welec + wmag = . (6.82)
8π
This also continues to be the energy density stored in the electromag-
netic field in the time dependent case which can be seen as follows
and which also brings out the concept of energy conservation within
the context of electromagnetic phenomena.
Let us consider a volume V containing electromagnetic fields as
well as sources (charges and currents). With time the energy stored
in the electromagnetic fields inside the volume would decrease in two
possible ways. First, there may be some dissipation of the energy
density due to conversion into heat or other forms of mechanical en-
ergy. For example, a wire carrying current may heat up due to the
resistance and in the process lose energy. Second, electromagnetic
waves may leave the volume V carrying with them energy. Let us
call the two kinds of energy losses as mechanical and radiation re-
spectively. The mechanical energy loss is easy to calculate directly
from the Lorentz force law. The rate at which the electromagnetic
field does work on a charged particle moving with velocity v is given
by
dWmech
Pmech =
dt
Z Z
3 3 1
= d xf · v = d xρ E + v × B · v
c
V V
Z Z
3
= d x E · (ρv) = d3 x E · J. (6.83)
V
avoiding in order not to have any confusion with the volume V ) loses
energy at the rate
I Z
Pmech = IE = I dℓ · E = d3 x E · J, (6.84)
V
∇ · (E × H) = −E · (∇ × H) + H · (∇ × E)
4π 1 ∂D 1 ∂B
= −E · J+ +H· −
c c ∂t c ∂t
4π ∂ D·E+B·H
=− E·J+
c ∂t 8π
4π ∂w
=− E·J+ . (6.85)
c ∂t
Thus, defining a vector
c
S= E × H, (6.86)
4π
we see that we can write (6.85) as
∂w
∇·S = − E·J+ ,
∂t
∂w
or, + ∇ · S = −E · J. (6.87)
∂t
This indeed has the structure of a continuity equation describing
conservation of energy (see, for example, (6.3)) with the term on the
right hand side representing the dissipation of energy calculated in
(6.83). This shows that w calculated in (6.82) indeed corresponds to
the energy density for time dependent electromagnetic fields. In fact,
integrating this over the volume V , we obtain
dW
= −(Pmech + Prad ), (6.88)
dt
which shows that the rate at which energy is lost has two parts with
Prad denoting the power carried out by the radiation fields and is
given by
Z Z
3
Prad = d x ∇ · S = ds · S. (6.89)
V S
6.4 Energy and the Poynting vector 193
1 ∂(D × B)
− B × (∇ × H) − , (6.90)
c ∂t
dpmech dprad
+
dt dt
Z
1
= d3 x E(∇ · D) + H(∇ · B) − D × (∇ × E)
4π
V
− B × (∇ × H) , (6.91)
194 6 Maxwell’s equations
∇ · D = 4πρ,
∇ · B = 0,
1 ∂B
∇×E=− ,
c ∂t
4π 1 ∂D
∇×H= J+ . (6.97)
c c ∂t
In the static case, we saw that we can write the electric and the mag-
netic fields in terms of scalar and vector potentials and the question
is whether we can continue to do so for time dependent fields.
To analyze this, let us note that the second of Maxwell’s equa-
tions implies that the magnetic field is divergence free. This can, of
course, be solved as in the static case to give
The only difference is that the vector potential, in the present case,
would be a function of time as well. Putting this back into the third
equation of Maxwell, we obtain
1 ∂B 1 ∂(∇ × A)
∇×E=− =− ,
c ∂t c ∂t
196 6 Maxwell’s equations
1 ∂A
or, ∇× E+ = 0,
c ∂t
1 ∂A
or, E+ = −∇Φ,
c ∂t
1 ∂A(x, t)
or, E(x, t) = −∇Φ(x, t) − . (6.99)
c ∂t
Once again, here the scalar potential Φ depends on space and time.
We note that when Φ, A are independent of time (6.98) and (6.99)
reduce to our earlier discussion of potentials in the static case. As
in the static case, it is clear that the scalar and the vector potentials
have an arbitrariness, namely,
1 ∂Λ(x, t)
Φ′ = Φ + , A′ = A − ∇Λ(x, t), (6.100)
c ∂t
where Λ(x, t) is an arbitrary function give the same electric and mag-
netic fields. Consequently, Maxwell’s equations, written in terms of
the scalar and the vector potentials, will reflect this arbitrariness.
This is known as the gauge invariance of Maxwell’s equations which
we will study in some detail next.
We see that two of Maxwell’s equations (second and third equa-
tions in (6.97)) can be solved to express the electric and the magnetic
fields in terms of scalar and vector potentials. Let us next substitute
the solutions (6.98) and (6.99) into the other two equations. In any
medium, Gauss’ law takes the form
∇ · D = ǫ∇ · E = 4πρ,
1 ∂A
or, ǫ∇ · −∇Φ − = 4πρ,
c ∂t
1 ∂(∇ · A) 4π
or, ∇2 Φ + = − ρ,
c ∂t ǫ
2 ǫµ ∂ 2 Φ 4π 1 ∂ ǫµ ∂Φ
or, ∇ Φ − 2 2
=− ρ− ∇·A+ . (6.101)
c ∂t ǫ c ∂t c ∂t
4π 1 ∂D
∇×H= J+ ,
c c ∂t
1 4π ǫ ∂ 1 ∂A
or, ∇ × (∇ × A) = J+ −∇Φ − ,
µ c c ∂t c ∂t
6.5 Gauge invariance of Maxwell’s equations 197
4πµ ǫµ ∂(∇Φ) ǫµ ∂ 2 A
or, ∇(∇ · A) − ∇2 A = J− − ,
c c ∂t c2 ∂t2
2 ǫµ ∂ 2 A 4πµ ǫµ ∂Φ
or, ∇ A − 2 2
=− J+∇ ∇·A+ . (6.102)
c ∂t c c ∂t
The two equations, (6.101) and (6.102), appear to be coupled
second order equations. However, let us recall that the scalar and the
vector potentials are arbitrary up to gauge transformations (6.100).
This is also reflected in the fact that the two coupled equations are
invariant under a gauge transformation. In such a case, the Cauchy
initial value problem cannot be solved uniquely unless we specify
some further conditions on the potentials. Let us, therefore, choose
the scalar and the vector potentials such that
ǫµ ∂Φ
+ ∇ · A = 0. (6.103)
c ∂t
With such a choice of the potentials, the two equations (6.101) and
(6.102) become decoupled and take the forms
ǫµ ∂ 2 Φ 4π
∇2 Φ − = − ρ,
c2 ∂t2 ǫ
ǫµ ∂ 2 A 4πµ
∇2 A − =− J. (6.104)
c2 ∂t2 c
Namely, with such a choice of the potentials, both the scalar and the
vector potentials satisfy the wave equation with sources.
The choice of a form of the potentials is known as a choice of
gauge. And the particular gauge we have chosen in (6.103) is known
as the Lorenz gauge (named after Ludvig Lorenz) which is manifestly
relativistic invariant as we will see. The choice of a gauge is subject
to the condition that it should be implementable, namely, that we
can always find potentials which would satisfy the gauge condition.
For the case at hand, for example, suppose our potentials did not
satisfy the Lorenz gauge condition, namely, if
ǫµ ∂Φ
∇·A+ 6= 0, (6.105)
c ∂t
then, we can make a gauge transformation
1 ∂Λ
Φ′ = Φ + , A′ = A − ∇Λ, (6.106)
c ∂t
requiring that the new potentials would satisfy the gauge condition
ǫµ ∂Φ′
∇ · A′ + = 0. (6.107)
c ∂t
198 6 Maxwell’s equations
This leads to
ǫµ ∂ 1 ∂Λ
∇ · (A − ∇Λ) + Φ+ = 0,
c ∂t c ∂t
2 ǫµ ∂ 2 ǫµ ∂Φ
or, ∇ − 2 2 Λ=∇·A+ . (6.108)
c ∂t c ∂t
∇ · A = 0. (6.109)
Here, we have used the continuity equation (6.3) as well as the famil-
iar decomposition (also discussed in connection with the Helmholtz
theorem in chapter 1, see (1.53), (1.54) and the discussion there) that
a given vector can be written as a sum of two terms, one longitudinal
and the other transverse with respect to the operation of ∇, namely,
J = Jl + Jt , ∇ · Jt = 0, ∇ × Jl = 0. (6.113)
This is, of course, consistent with the choice of the gauge condition
(6.109). Since the vector potential is transverse in this case, one also
says that this is the transverse gauge (sometimes it is also called the
physical gauge). The Coulomb gauge is quite useful when there are
no sources present, namely, when ρ = 0 = J. In such a case, the
solution of the Poisson equation is trivial and the vector potential
satisfies the free wave equation
ǫµ ∂ 2 A
∇2 A − = 0. (6.117)
c2 ∂t2
It is also worth noting here that even after choosing a gauge,
there is some residual gauge invariance. For example, in the case of
the Lorentz gauge (6.103), even when
ǫµ ∂Φ
+ ∇ · A = 0, (6.118)
c ∂t
holds we can still make a gauge transformation preserving this gauge.
In other words, we can define a new set of potentials
1 ∂ Λ̃
Φ′ = Φ + , A′ = A − ∇Λ̃, (6.119)
c ∂t
which would also satisfy the Lorentz condition provided
2 ǫµ ∂ 2
∇ − 2 2 Λ̃ = 0. (6.120)
c ∂t
Unlike the case of the Laplacian, the D’Alembertian operator allows
for oscillatory solutions to this homogeneous equation.
x′ = x − vt,
t′ = t, (6.121)
One such puzzle came from the plane wave solutions of Maxwell’s
equations which, as we have seen, travel in vacuum with the speed of
light which is a constant. An absolute velocity, however, was against
the spirit of Galilean relativity. Consequently, there was a major con-
flict and to avoid this conflict, Maxwell even assumed that the wave
solutions of his equations travel in a medium called ether, namely,
he tried to promote that the speed with which the waves propagate
is really the speed of light in a specific inertial frame represented by
ether. On the other hand, Michelson and Morley conclusively showed
through their experiments that there is no ether and, therefore, the
speed of light is a constant independent of the inertial frame of ref-
erence. This was indeed the turning point in thinking, for Maxwell’s
equations were not invariant under Galilean transformations. (In
retrospect even a simple experimental measurement such as the life
time of muon, performed decades later, cannot conform to the ideas
of Galilean invariance. Experimentally it is measured that the life
time of the muon decaying at rest in the laboratory is shorter by an
order of magnitude than the life time measured for the muons de-
caying in the cosmic ray showers (τlab ≈ 10−6 sec). This cannot be
explained with Galilean relativity where time is absolute and does
not depend on the choice of the reference frame.)
by
′ β
t =γ t− x ,
c
x′ = γ(x − βct),
y ′ = y,
z ′ = z, (6.122)
All of this led Einstein to propose that space and time together
should be thought of as defining a four dimensional manifold where
events take place and that Lorentz transformations are symmetry
transformations which transform the coordinates of this four dimen-
sional manifold, much like rotations transform the coordinates of the
three dimensional space. A vector in such a manifold would consist
of four components – one time and three space – and is called a four
vector, as opposed to the vectors in three dimensional space that we
are all familiar with. However, space and time components can be
embedded into this four vector in two distinct ways. For example,
let us consider the space-time coordinates themselves which define a
four vector. We note that we can define a four component vector as
x′µ = Λµ ν xν ,
e µ ν xν ,
x′µ = Λ (6.127)
It is clear from the structure of the matrices in (6.128) that they are
inversely related,
e λ ν )T = δµ .
Λµ ν (Λ (6.129)
λ
This is very much like the relation for the rotation matrices in three
dimensional space and suggests that Lorentz transformations can be
thought of as rotations in the four dimensional space-time manifold.
204 6 Maxwell’s equations
c2 t2 − x2 = xµ xµ = xµ xµ = x2 , (6.133)
x2 = ηµν xµ xν = η µν xµ xν , (6.134)
where the covariant and the contravariant metric tensors are deter-
mined from (6.133) and (6.134) to be
1 0 0 0
0 −1 0 0
ηµν = ,
0 0 −1 0
0 0 0 −1
6.6 Lorentz transformation 205
1 0 0 0
0 −1 0 0
η µν
=
0 0 −1 0
.
(6.135)
0 0 0 −1
The metric tensor also allows us to raise and lower the Lorentz
indices,
xµ = η µν xν , xµ = ηµν xν , (6.137)
so that the time component of a vector does not change sign under
raising or lowering while the space components do. Let us note here
that the three dimensional space that we are used to is known as
a Euclidean space where the metric tensor is the trivial Kronecker
delta function δij . Consequently, there is no difference between the
covariant and the contravariant vectors there. The four dimensional
space-time manifold, on the other hand, has a nontrivial metric tensor
giving rise to distinct covariant and contravariant vectors. A manifold
with such a metric (as in (6.135)) is known as a Minkowski space.
Furthermore, from the definition of the invariant length in (6.134),
we see that unlike the three dimensional case, here the length of a
nontrivial (four) vector is not necessarily positive. In fact, it can be
positive, negative, or zero. If
2
x2 = x0 − x2 > 0, (6.138)
lig
ht ke
-li t-li
ke h
ct lig
time-like
space-like space-like
time-like
Figure 6.7: Four dimensional Minkowski space-time projected onto
two dimensions.
A · B = ηµν Aµ B ν = A0 B 0 − A · B = Aµ B µ = Aµ Bµ , (6.141)
e λ λ′ T µ′ ν ′ λ′ .
T µν λ → T ′ µν λ = Λµ µ′ Λν ν ′ Λ (6.143)
Note the change in the relative signs between the time and the space
components of the gradient vectors compared with the components
of the coordinate vectors in (6.125) and (6.126). Since gradients
define vectors we can define a scalar (which would be invariant under
Lorentz transformations) from these as
1 ∂2
= ∂ 2 = ηµν ∂ µ ∂ ν = − ∇2 . (6.145)
c2 ∂t2
As we have seen, this is the wave operator and is known as the
D’Alembertian operator (remember that we have restricted ourselves
to free space for which ǫ = 1 = µ). It is invariant under Lorentz trans-
formations just as the Laplacian is invariant under three dimensional
rotations.
There are other familiar quantities from the study of three di-
mensions which also combine into four vectors. (To be able to com-
bine distinct quantities into a four vector, they must have the right
transformation properties under a Lorentz transformation.) Of course,
the most familiar is the fact that energy and momentum combine into
a four vector such that
E E
pµ = ,p , pµ = , −p . (6.146)
c c
E2
p2 = ηµν pµ pν = − p2 , (6.147)
c2
is Lorentz invariant and we see from this that we can write the Ein-
stein relation as an invariant relation
E2
p2 = − p2 = m2 c2 ,
c2
or, E 2 = c2 p2 + m2 c4 , (6.148)
∂ρ
∂µ J µ = ∂0 J 0 + ∂i J i = + ∇ · J = 0. (6.150)
∂t
F0i = ∂0 Ai − ∂i A0
1 ∂(A)i
=− − ∇i Φ = (E)i ,
c ∂t
6.7 Covariance of Maxwell’s equations 209
Fij = ∂i Aj − ∂j Ai
= − (∇i (A)j − ∇j (A)i ) = −ǫijk (∇ × A)k
= −ǫijk (B)k . (6.153)
e 0 0Λ
= (Λ ei j − Λ
e0 j Λ
e i 0 )F0j + Λ
e0 j Λ
e i k Fjk ,
e i 0Λ
Fij′ = Λ e j k F0k + Λ
ei kΛ
e j 0 Fk0 + Λ
ei kΛ
e j l Fkl
e i 0Λ
= (Λ ej k − Λ
ei kΛ
e j 0 )F0k + Λ
ei kΛ
e j l Fkl . (6.155)
which is, of course, the Gauss’ law. On the other hand, if we choose
ν = j, then (6.157) leads to
4π j
∂0 F 0j + ∂i F ij = −∂0 F0j + ∂i Fij = J ,
c
1 ∂(E)j 4π
or, − + ∇i (−ǫijk (B)k ) = (J)j ,
c ∂t c
4π 1 ∂(E)j
or, (∇ × B)j = (J)j + ,
c c ∂t
4π 1 ∂E
or, ∇×B= J+ . (6.159)
c c ∂t
Clearly, these equations are manifestly Lorentz covariant, since they
are expressed in terms of Lorentz covariant quantities. In fact, we
note that the left hand side of (6.157) behaves like a vector under
a Lorentz transformation (namely, has only one free index) and the
same is true for the right hand side as well. Defining the dual of the
field strength tensor as
1
Feµν = ǫµνλρ Fλρ = −Fe νµ , (6.160)
2
where ǫµνλρ is the completely anti-symmetric (Levi-Civita) tensor in
four dimensions with ǫ0123 = 1 (ǫ0ijk = ǫijk ), we recognize that
1 1
Fe0i = ǫ0ijk Fjk = ǫijk (−ǫjkl (B)l ) = −(B)i ,
2 2
Feij = ǫij0k F0k = ǫ0ijk F0k = ǫijk (E)k , (6.161)
∂µ Feµν = 0. (6.162)
6.7 Covariance of Maxwell’s equations 211
As we have seen in (6.104), in the Lorenz gauge both the scalar and
the vector potentials satisfy the wave equation with sources, namely,
(for simplicity, we are choosing free space)
1 ∂2 2
Φ = − ∇ Φ = 4πρ,
c2 ∂t2
1 ∂2 2 4π
A = 2 2
−∇ A= J. (6.169)
c ∂t c
Of course, we can combine the two equations in (6.169) into a single
covariant equation of the form (which is consistent with the Lorenz
gauge condition)
1 ∂2 2 4π
Aµ = − ∇ Aµ = Jµ . (6.170)
c2 ∂t2 c
It is clear, therefore, that we can solve for the potentials if we know
the solutions to the equation of the form
1 ∂2 2
Ψ = − ∇ Ψ(x, t) = 4πf (x, t). (6.171)
c2 ∂t2
As we have discussed earlier, a simple way to solve such inho-
mogeneous equations is through the method of Green’s functions.
Namely, let us define the Green’s function G(x, t; x′ , t′ ) for the wave
operator to satisfy the equation,
1 ∂2 2
G = − ∇ G(x, t; x′ , t′ ) = −4π δ4 (x − x′ )
c2 ∂t2
= −4πδ3 (x − x′ )δ(c(t − t′ )). (6.172)
Z
4 ′ 1 ′
δ (x − x ) = d4 k e−ik·(x−x ) , (6.175)
(2π)4
4π 1
k 2 G(k) = = 3,
(2π)4 4π
1 1 1 1
or, G(k) = = 3 ω2 . (6.176)
4π 3 k 2 4π ( 2 − k2 )
c
In other words, the cause and the effect are related in a retarded
manner, namely, the effect cannot precede the cause.
Im ω
b b
Re ω
−ck − iǫ ck − iǫ
Figure 6.8: Shifted poles in the complex energy plane (with k = |k|)
for the retarded Green’s function.
both of which are in the lower half of the complex ω plane as in Fig.
6.8. Here we have defined k = |k| for simplicity.
To see that this indeed satisfies the retarded boundary condition,
let us evaluate the ω integral in (6.179) for t < t′ . In such a case, the
6.8 Retarded Green’s function 215
Im ω
b b
Re ω
−ck − iǫ ck − iǫ
b b
Re ω
−ck − iǫ ck − iǫ
G(R) (x, t; x′ , t′ )
Z " ′ ′ ′ ′
#
i 3 e−ikc(t−t )+ik·(x−x ) eikc(t−t )+ik·(x−x )
=− d k − .
(2π)2 k k
(6.182)
R = x − x′ , R = |R|, (6.183)
we can write
Z h
i ′
G(R) (x, t; x′ , t′ ) = − kdkd(cos θ)dφ e−ikc(t−t )+ikR cos θ
(2π)2
′
i
−eikc(t−t )+ikR cos θ
Z∞ h
1 ′ ′
=− dk e−ik(c(t−t )−R) − e−ik(c(t−t )+R)
2πR
0
′ ′
i
−eik(c(t−t )+R) + eik(c(t−t )−R)
Z∞ h i
1 ′ ′
=− dk e−ik(c(t−t )−R) − e−ik(c(t−t )+R)
2πR
−∞
1
=− δ(c(t − t′ ) − R) − δ(c(t − t′ ) + R) . (6.184)
R
It is clear that since t − t′ > 0 (and note that R > 0), the second
delta function does not contribute. Thus, we determine the retarded
Green’s function of the wave equation to be (for t > t′ , a condition
which can be implemented through a step function)
δ(c(t − t′ ) − |x − x′ |)
G(R) (x, t; x′ , t′ ) = − . (6.185)
|x − x′ |
With this, we can now obtain a particular solution of the wave
equation (6.171) satisfying the retarded boundary condition as (see
(6.173))
Z
Ψ(x, t) = − d4 x′ G(R) (x, x′ )f (x′ )
6.8 Retarded Green’s function 217
Z
δ(c(t − t′ ) − |x − x′ |)f (x′ , t′ )
= d3 x′ cdt′
|x − x′ |
Z
f (x′ , t′ )
= d 3 x′ . (6.186)
|x − x′ | t′ =t− |x−x′ |
c
It is clear that this gives a nontrivial solution only for later times,
′|
t = t′ + |x−x c (namely, a retarded solution). Furthermore, from
this we can now write down the retarded solutions for the Maxwell’s
equations (in the Lorenz gauge) (6.170) to be
Z ′ ′
1 3 ′ Jµ (x , t )
Aµ (x, t) = d x , (6.187)
c |x − x′ | t′ =t− |x−x′ |
c
where,
dξ(t)
j µ (t) = (cq, qv) = cq, q . (6.190)
dt
Given this, we can compute the potential that such a moving charge would produce
using (6.188). Namely,
′
Z jµ (t′ )δ 3 (x′ − ξ(t′ ))δ t′ − t + |x−x |
1 ′ 3 ′ c
Aµ (x, t) = dt d x
c |x − x′ |
jµ (t′ )δ t′ − t + |x−ξc(t )|
′
Z
1 ′
= dt . (6.191)
c |x − ξ(t′ )|
Let us next define
|x − ξ(t′ )|
τ = t′ − t + . (6.192)
c
218 6 Maxwell’s equations
Then, the integration over t′ in (6.191) would determine the time coordinate t′ to
be the one for which the argument of the delta function vanishes, namely,
|x − ξ(t′ )|
τ = t′ − t + = 0. (6.193)
c
This can be solved for t′ once we know the trajectory of the particle. Furthermore,
dτ 1 d|x − ξ(t′ )|
′
=1+ , (6.194)
dt c dt′
and using the standard formula for integration with a delta function, namely,
Z
1
dx δ(f (x)) g(x) = df (x) g(x0 ), (6.195)
| dx |x0
where x0 represents the solution of f (x) = 0 (we assume here that there is only
one root x0 of the equation), we obtain
Z ′ ′ |x−ξ (t′ )|
1 jµ (t )δ t − t + c
Aµ (x, t) = dt′
c |x − ξ(t′ )|
1 jµ (t )′
=
c |x − ξ(t′ )| 1 + 1 d|x−ξ(t )| ′
c dt′
τ =0
1 jµ (t′ )
= . (6.196)
c |x − ξ(t′ )| + 1 d|x−ξ(t′ )|2
2c dt′ τ =0
These are known as the Lienard-Wiechert potentials. (We note here that for slow
moving particles it is not necessary to have the magnitude in the Jacobian coming
from the delta function.) From these, we can easily determine the electric and the
magnetic fields that a moving charged particle produces which, in turn, are used
in the study of radiation due to a moving charged particle. We will discuss this
in more detail in a later chapter. ◭
Z Ztf Z
4 ′ ′ ′
′
d x Φ Ψ − Ψ Φ = cdt d3 x′ Φ∂ ′2 Ψ − Ψ∂ ′2 Φ
ti
Ztf Z
′
= cdt d3 x′ ∂µ′ Φ∂ ′µ Ψ − Ψ∂ ′µ Φ
ti
Z
= ds′µ Φ∂ ′µ Ψ − Ψ∂ ′µ Φ
Z t′ =tf
1 ∂Ψ
3 ′ ∂Φ
= d x Φ ′ −Ψ ′
c ∂t ∂t t′ =ti
Ztf Z
−c dt′ ds′ · Φ∇′ Ψ − Ψ∇′ Φ . (6.197)
ti
Here ds′ denotes integration over the two dimensional surface bound-
ing the three dimensional volume. This is the generalization of Green’s
identity to the four dimensional case for any two arbitrary functions
Ψ and Φ.
Let us now specialize to the case where
such that
tf , we obtain,
Z
Ψ(x, t) = − d4 x′ G(x, x′ )f (x′ )
Z t′ =tf
1 ∂Ψ
3 ′ ∂G
+ d x G ′ −Ψ ′
4πc ∂t ∂t t′ =ti
Ztf Z
c ′
− dt ds′ · G∇′ Ψ(x′ , t′ ) − Ψ(x′ , t′ )∇′ G . (6.200)
4π
ti
Ztf Z
c ′
− dt ds′ · G(R) ∇′ Ψ(x′ , t′ ) − Ψ(x′ , t′ )∇′ G(R)
4π
ti
Z
f (x′ , t′ )
3 ′
= d x
|x − x′ | t′ =t− |x−x′ |
c
Z " #
1 ∂Ψ ∂G(R)
− d3 x′ G(R) ′ − Ψ
4πc ∂t ∂t′ ′ t =ti
Ztf Z
c ′
− dt ds′ · G(R) ∇′ Ψ(x′ , t′ ) − Ψ(x′ , t′ )∇′ G(R) .
4π
ti
(6.201)
There are two special cases that we will consider now. First,
let us consider the case where the volume of space is infinite. In
such a case, with the assumptions of asymptotic fall off for the fields
(variables), the surface integral in (6.201) vanishes. The remaining
terms are determined completely in terms of the initial values of Ψ
and ∂Ψ∂t . Namely, we have a solution of the Cauchy initial value
6.9 Kirchhoff’s representation 221
This gives the solution of the initial value problem once we know the
explicit forms of F (x) and H(x).
The second case that we are interested in is when the volume is
finite. Furthermore, let us assume that there are no sources present
in this volume, namely, f (x) = 0 and that the initial values are also
trivial. In such a case, (6.201) leads to
Ztf Z
c
Ψ(x, t) = − dt′ ds′ · G(R) ∇′ Ψ(x′ , t′ ) − Ψ(x′ , t′ )∇′ G(R) .
4π
ti
(6.205)
222 6 Maxwell’s equations
Once again, we can use the form of the retarded Green’s function in
(6.185) and defining, for simplicity R = x − x′ , we have
′
′ (R) ′ δ(c(t − t ) − R)
∇G = −∇
R
′ ∂ δ(c(t − t′ ) − R)
= −(∇ R)
∂R R
δ(c(t − t′ ) − R) 1 ∂δ(c(t − t′ ) − R)
= −R̂ − , (6.206)
R2 cR ∂t′
where we have used (∇′ R) = −R̂. Using this, we can do the time
integral in (6.205) to obtain
Ztf Z
c ′ ′ δ(c(t − t′ ) − R) ′
Ψ(x, t) = dt ds · ∇Ψ
4π R
ti
δ(c(t − t′ ) − R) 1 ∂δ(c(t − t′ ) − R)
− R̂ − Ψ
R2 cR ∂t′
Z " #
1 ′ 1 ′ R̂ R̂ ∂Ψ
= ds · ∇ Ψ − 2Ψ − . (6.207)
4π R R cR ∂t′ ′ R
t =t− c
1. (a) Show that the four Maxwell’s equations (in any medium)
imply the continuity equation.
(b) For a system with well localized charges and currents (namely,
charges and currents which do not extend to infinity), show that
the total charge
Z
Q= d3 x ρ(x, t),
all space
6.10 Selected problems 223
J = σE.
L
(Just to remind you, the resistance R is defined as R = A ρ.)
3. Work out the plane wave solution for a charge neutral (without
any free charge) conducting medium, for which you may assume
J = σE. Compare the present solution with the one for the
dielectrics obtained in this chapter.
Wave guides
225
226 7 Wave guides
∇ · D = 4πρ,
∇ · B = 0,
1 ∂B
∇×E=− ,
c ∂t
4π 1 ∂D
∇×H= J+ . (7.1)
c c ∂t
Various fields satisfy the relations noted earlier in (6.13)
D = ǫ E,
1
H= B, (7.2)
µ
where ǫ, µ represent respectively the permittivity and the permeabil-
ity of the medium. In addition, in a metal, we can relate the con-
duction current to the electric field through Ohm’s law as (see the
second relation in (6.12))
J = σ E, (7.3)
the third equation of Maxwell, we see that electric fields must form
closed loops around time varying magnetic fields. Similarly, the last
equation of Maxwell leads to the fact that magnetic fields must form
closed loops around a conduction current or a “displacement” current
produced by a time varying electric field. This is the general behavior
of electric and magnetic fields when time dependence is present. If we
have a perfect conductor as would be the case for the walls of a wave
guide, then, we realize that the electric and the magnetic fields cannot
penetrate inside the metal. Any change in the external fields would
lead to an instantaneous response whereby charges will move to the
surface of the conductor to prevent any field within. In particular,
we note from (7.3) that, for a perfect conductor with σ → ∞, there
cannot be any tangential component of the electric field present on the
surface of the conductor (for any t), which would otherwise imply an
infinite conduction current that is physically untenable. Similarly, the
normal component of the magnetic field must vanish on the surface of
the conductor simply because there are no magnetic monopoles (on
the surface of the conductor). Mathematically, we can write these
boundary conditions as
n̂ × E| = 0,
n̂ · B| = 0, (7.4)
This follows from the familiar vector identity (if A, B, C involve op-
erators, the order should be maintained)
A × (B × C) = (A · C) B − (A · B) C. (7.6)
228 7 Wave guides
∂
∇ = ∇⊥ + ∇k = ∇⊥ + ẑ . (7.7)
∂z
Exercise. Although ẑ × E gives a component of the E field normal to the z-axis,
show that (7.5) gives the correct decomposition of the fields. Namely, show that
any arbitrary vector V can be written uniquely as
∂Ez
∇⊥ · E⊥ = − ,
∂z
∂Bz
∇⊥ · B⊥ = − . (7.9)
∂z
This, in turn, leads to (taking the dot as well as cross product with
ẑ)
1 ∂Bz
ẑ · (∇⊥ × E⊥ ) = − ,
c ∂t
∂E⊥ 1 ∂B⊥
− ẑ × = ∇⊥ Ez , (7.11)
∂z c ∂t
where we have used the fact that ∇⊥ × E⊥ points along the z-axis as
well as (7.6). Similarly, the last equation in (7.1) can be decomposed
into
ǫµ ∂Ez
ẑ · (∇⊥ × B⊥ ) = ,
c ∂t
∂B⊥ ǫµ ∂E⊥
+ ẑ × = ∇⊥ Bz . (7.12)
∂z c ∂t
7.1 Boundary conditions 229
Let us next assume that the fields are harmonically varying with
time. Furthermore, since we want the wave to be propagating along
the z-axis, we can also extract the z dependence of the fields to write
∇⊥ · E⊥ = −ikEz ,
∇⊥ · B⊥ = −ikBz ,
iω
ẑ · (∇⊥ × E⊥ ) = Bz ,
c
iω
ikE⊥ + ẑ × B⊥ = ∇⊥ Ez ,
c
iǫµω
ẑ · (∇⊥ × B⊥ ) = − Ez ,
c
iǫµω
ikB⊥ − ẑ × E⊥ = ∇⊥ Bz . (7.14)
c
i ω
E⊥ = ǫµω 2
k∇⊥ Ez − ẑ × ∇⊥ Bz ,
− k2 c
c2
i ǫµω
B⊥ = ǫµω 2
k∇⊥ Bz + ẑ × ∇⊥ Ez . (7.15)
− k2 c
c2
kc
B⊥ = ẑ × E⊥ . (7.21)
ω
Furthermore, we note that since the direction of propagation is chosen
to be along the z-axis, the boundary conditions (7.4) for a wave guide
take the simpler form
∂Bz
Etan | = 0 = , (7.22)
∂n
which vanishes at the boundary. Using (7.23) and (7.24) in the last
equation of (7.14) leads to
∂
n̂ · ∇⊥ Bz | = n̂ · ∇ − ẑ Bz = n̂ · ∇Bz |
∂z
∂Bz
= = 0, (7.25)
∂n
Let us consider a rectangular hollow tube along the z-axis and with
transverse dimensions satisfying 0 ≤ x ≤ a, 0 ≤ y ≤ b as shown in
232 7 Wave guides
y=b
y
z
x=a x
Fig. 7.1. As we have noted, the walls of the tube are assumed to be
perfect conductors. In this case, the solutions of Maxwell’s equations
have to satisfy the boundary conditions (7.22). It is clear that, in
this case, for the wall in the x-z plane, n̂ is parallel to ŷ and this
leads to the boundary conditions (see (7.22))
∂Bz
Ex |y=0,b = 0 = Ez |y=0,b , = 0. (7.26)
∂y y=0,b
On the other hand, for the wall in the y-z plane, n̂ is parallel to x̂
leading to the boundary conditions
∂Bz
Ey |x=0,a = 0 = Ez |x=0,a , = 0. (7.27)
∂x x=0,a
Equations (7.26) and (7.27) define all the boundary conditions in this
case. We note that Eq. (7.15) explicitly takes the forms
i ∂Ez ω ∂Bz
Ex = ǫµω2 k + ,
2 ∂x c ∂y
c2 − k
i ∂Ez ω ∂Bz
Ey = ǫµω2 k − ,
2 − k 2 ∂y c ∂x
c
i ∂Bz ǫµω ∂Ez
Bx = ǫµω2 k − ,
− k2 ∂x c ∂y
c2
i ∂Bz ǫµω ∂Ez
By = ǫµω2 k + , (7.28)
2 − k 2 ∂y c ∂x
c
7.2 Rectangular wave guide 233
that the quantity inside the square root can be negative. In this
case, k will become purely imaginary. As a result, there will be no
propagation inside the wave guide. Rather, the electromagnetic signal
will be attenuated along the tube. Defining
r √ p
πc m2 n 2 ǫµ
ωmn = √ 2
+ 2, k = ω 2 − ωmn
2 , (7.35)
ǫµ a b c
we see that propagation of the TMmn wave can take place inside a
rectangular wave guide only if ω > ωmn which is known as the cut-
off frequency below which propagation of the TMmn wave will not
occur. Above this frequency, the TMmn wave will propagate without
any attenuation for a wave guide with perfectly conducting walls.
The wavelength of propagation inside the wave guide for the
TMmn wave is easily obtained to be
2π 2πc 1
λ= =√ p
k ǫµ ω − ωmn
2 2
2πc 1
=√ r . (7.36)
ǫµ π 2 c2 m2 n2
ω2 − ǫµ a2
+ b2
c ω
=√ r . (7.37)
ǫµ 2 2 2 n2
ω 2 − πǫµc m a2
+ b2
Note that, in deriving all these results, we have assumed the inside of
the wave guide to be filled with an arbitrary dielectric. If we assume
that it is empty space inside, then in this case we can identify ǫµ = 1
and the formulae in this section simplify. With this, we note from
Eq. (7.37) that the velocity of propagation of the wave, inside the
wave guide, is larger than its value in free space. We note that it
7.2 Rectangular wave guide 235
∂ω c2 k c2 1
vg = = = ,
∂k ǫµ ω ǫµ v
c2
or, vg v = , (7.38)
ǫµ
so that, when the phase velocity is larger than the speed of light, the
group velocity is smaller, as it should be.
As is clear, the wave guide can support an infinite number of
TM modes. It is for the mode corresponding to m = 1 = n that
the cut-off frequency is the smallest and would correspond to the
dominant TM mode. In this case, with ǫµ = 1, we have
√
πc a2 + b2
ω11 = ,
ab
c
λ = 2π q ,
2 2 2 2
ω 2 − π c a(a2 b2+b )
ω
v=c q . (7.39)
π 2 c2 (a2 +b2 )
ω2 − a2 b2
πmx πny
Bz (x⊥ ) = Bz (x, y) = C cos cos , (7.40)
a b
2 ǫµω 2 π 2 m2 π 2 n 2
k = 2 − + 2 , m, n = 0, 1, 2, . . . . (7.41)
c a2 b
236 7 Wave guides
2π 2πc 1
λ= =√ q ,
k ǫµ ω 2 − π 2 c2
ǫµa2
ω c ω
v= =√ q . (7.46)
k ǫµ ω 2 − π 2 c2
ǫµa2
The other kind of wave guide that is commonly used is the cylindrical
wave guide. Here, we have a hollow cylindrical tube of radius a with
a perfectly conducting wall and length of the tube along the z-axis
as shown in Fig. 7.2. Because of the symmetry in the problem, it is
simpler to study this problem in cylindrical coordinates defined by
r=a
x̂ = r̂ cos φ − φ̂ sin φ,
ŷ = r̂ sin φ + φ̂ cos φ,
ẑ = ẑ.
r̂ = x̂ cos φ + ŷ sin φ,
From these, it follows that the only non-trivial variation of the unit
vectors are given by
∂r̂ ∂ φ̂
= φ̂, = −r̂. (7.48)
∂φ ∂φ
∂ φ̂ ∂
∇⊥ = r̂ + ,
∂r r ∂φ
2 1 ∂ ∂ 1 ∂2
∇⊥ = r + 2 .
r ∂r ∂r r ∂φ2
With these, we can go back and recast all the equations (7.13)–
(7.17) and (7.22) in cylindrical coordinates. In particular, the bound-
ary conditions (7.22) in the case of cylindrical wave guides take the
form (here n̂ is parallel to r̂)
∂Bz
Ez |r=a = 0 = Eφ |r=a , = 0. (7.49)
∂r r=a
The relation for the transverse components in (7.15) take the forms
i ∂Ez ω ∂Bz
Er (x⊥ ) = Er (r, φ) = ǫµω2 k + ,
2 − k 2 ∂r cr ∂φ
c
i k ∂Ez ω ∂Bz
Eφ (x⊥ ) = Eφ (r, φ) = ǫµω2 − ,
2 − k 2 r ∂φ
c
c ∂r
7.3 Cylindrical wave guide 239
i ∂Bz ǫµω ∂Ez
Br (x⊥ ) = Br (r, φ) = ǫµω2 k − ,
− k2 ∂r cr ∂φ
c2
i k ∂Bz ǫµω ∂Ez
Bφ (x⊥ ) = Bφ (r, φ) = ǫµω2 + , (7.50)
2 − k 2 r ∂φ c ∂r
c
while the equations for the Ez , Bz (see Eqs. (7.16)–(7.17)) take the
forms
∂ 2 Ez 1 ∂Ez 1 ∂ 2 Ez ǫµω 2 2
+ + 2 + − k Ez = 0,
∂r 2 r ∂r r ∂φ2 c2
∂ 2 Bz 1 ∂Bz 1 ∂ 2 Bz ǫµω 2 2
+ + 2 + − k Bz = 0. (7.51)
∂r 2 r ∂r r ∂φ2 c2
Physical solutions have to be periodic in the angular variable.
Consequently, we can extract the φ dependence of the z-components
of the fields as
Ez (r, φ) = Ez(0) (r) An ein(φ+φ0 ) + Bn e−in(φ+φ0 ) ,
Bz (r, φ) = Bz(0) (r) Cn ein(φ+φ0 ) + Dn e−in(φ+φ0 ) . (7.52)
0.8
0.6
J0 HxL
0.4
0.2
-0.2
-0.4
0 2 4 6 8 10 12
x
Jn (ha) = 0. (7.57)
There are an infinite number of zeros of the Bessel function for any
n and denoting all such roots as (ha)mn (implying the mth zero of
Jn ), we note that the first few take the values
from (7.50) to be
ian
Er (r, φ) = ǫµω 2
k Jn′ (hr) cos nφ,
c2
− k2
ian kn
Eφ (r, φ) = − ǫµω2 Jn (hr) sin nφ,
2 − k 2 r
c
ian ǫµωn
Br (r, φ) = ǫµω 2
Jn (hr) sin nφ,
− k2 cr
c2
ian ǫµω ′
Bφ (r, φ) = ǫµω 2
Jn (hr) cos nφ. (7.60)
− k2 c
c2
ibn ω ′
Eφ (r, φ) = − ǫµω2 J (hr) cos nφ,
− k 2 c n
c 2
ibn
Br (r, φ) = ǫµω 2
k Jn′ (hr) cos nφ,
c2
− k2
ibn kn
Bφ (r, φ) = − ǫµω2 Jn (hr) sin nφ. (7.68)
c2
−k r
2
and correspondingly, for the propagating modes, ω > ωmn , the wave
length and the velocity of propagation are given as in Eqs. (7.63) and
(7.64). We note from (7.66) that the dominant mode, in a cylindrical
wave guide, is the TE11 mode. Thus, we see that the rectangular and
the cylindrical wave guides have similar qualitative features.
Ez = 0 = Bz .
Bz = 0,
πmC πmy
Ez (y) = sin ,
b b
Ex (y) = 0,
πmy
Ey (y) = ikC cos ,
b
iǫµωC πmy ǫµω
Bx (y) = − cos =− Ey ,
c b kc
By = 0. (7.70)
Here, we have
r
ǫµω 2 π 2 m2
k= − , (7.71)
c2 b2
and m = 0, 1, 2, . . .. In fact, when m = 0, we see that the field
configuration has the form
Bz (y) = 0 = Ez (y),
Ex (y) = 0 = By (y),
Ey (y) = ikC,
iǫµωC
Bx = − , (7.72)
c
with
√
ǫµ ω
k= . (7.73)
c
This field configuration is clearly that of a TEM wave and we see that
waves guided by two infinite parallel conducting planes can support
TEM modes unlike a hollow rectangular wave guide. Furthermore,
in this case, we see from (7.72) and (7.73) that we can write
√
B⊥ (y) = ǫµ ẑ × E⊥ , (7.74)
4πµ ǫµ ∂E
∇×B= J+ ,
c c ∂t
and the closed magnetic loops in the transverse plane can enclose the
conduction current and we do not need a z-component of the electric
field to be present. In fact, from the radial symmetry of the coaxial
cable, we see that the electric field must be radial everywhere. As
a result the “displacement” current must also be along the radial
direction. Since the conduction current is along the z-axis and the
“displacement” current along the radial direction, it follows from the
above equation that the B field cannot have a component along the
z-axis since there is no current along the φ (angular) direction. This
shows that TEM wave is the only wave that can exist in a coaxial
cable. A similar conclusion also follows for two wire transmission
lines. Furthermore, since TEM waves do not have a cut-off frequency,
coaxial cables or two wire transmission lines are used to transmit low
frequency electromagnetic waves.
kc c k 1
ZTM = =√ √
ǫµω ǫµ ω ǫµ
r ω 2
mn
= 1− ZTEM ,
ω
√
ω ǫµ ω 1
ZTE = = √
kc c k ǫµ
1
=q ZTEM . (7.77)
ωmn 2
1− ω
Of course, the values of ωmn are different for the two wave guides.
However, for a given wave guide (for example, rectangular), we see
from the forms of the impedances in (7.77) that if the space inside is
empty (ǫµ = 1, ZTEM = 1), then we can formally write
1
ZTM = . (7.78)
ZTE
Thus far, we have discussed wave guides where the external wall is
assumed to be a perfect conductor with σ → ∞. In reality, however,
the metallic conductor has a finite conductivity, be it very large.
ǫω
Normally, a metal is considered a good conductor if σ ≫ 4π (see
(8.57)). When the conductivity is finite, as in a realistic conductor,
the electromagnetic fields within the wave guide can penetrate inside
the conducting walls. The penetration depth, also known as the skin
depth, is normally very small and this phenomenon induces a surface
current in the metal that plays a very important role. Namely, it
leads to heating and Ohmic losses in the conductor. Even though
this effect is small, it leads to an attenuation of the electromagnetic
fields that are propagated inside the wave guide. This is reflected in
the fact that in the regime of propagation (namely, for ω > ωmn ), the
7.6 Attenuation factor in wave guides 247
ω Power transmitted
Q=
vg Power lost per unit length
ω 1 1 ǫµω
= = v
vg 2α 2α c2
√
ǫµ ω
= q 2 , (7.93)
2αc
1 − ωmn
ω
where we have used Eqs. (7.37) and (7.38). Since, in wave guides,
the attenuation factor α can be very low, it is possible to construct
wave guides with large Q factors. This becomes quite important in
the construction of resonating cavities, which we will study next.
which follows from the first equation of (7.9), since (7.1) requires that
E⊥ = 0 at z = 0, d. This, then, determines the form of Ez from Eq.
(7.16) to be (compare with Eq. (7.30))
πmx πny πℓz
Ez (x, y, z) = A sin sin cos , (7.95)
a b d
with ℓ = 0, 1, 2, . . . as well as the usual restrictions on m, n and
r
πc m2 n 2 ℓ2
ω=√ + + = ωmnℓ , (7.96)
ǫµ a2 b2 d2
which follows from (7.34) with the identification k = πℓ
d . The trans-
verse field components can now be determined from the Maxwell’s
equations, (7.9), (7.10), (7.11) and (7.12), to be
1 ∂Ez
E⊥ (x) = E⊥ (x, y, z) = ǫµω 2
∇⊥ ,
π 2 ℓ2 ∂z
c2 − d2
i ǫµω
B⊥ (x) = B⊥ (x, y, z) = ǫµω 2
ẑ × ∇⊥ Ez , (7.97)
π 2 ℓ2 c
c2
− d2
which lead explicitly to
π 2 mℓ
ad πmx πny πℓz
Ex (x, y, z) = −A ǫµω 2 2 2
cos sin sin ,
− πd2ℓ a b d
c2
π 2 nℓ
bd πmx πny πℓz
Ey (x, y, z) = −A ǫµω 2 2 2
sin cos sin ,
− πd2ℓ a b d
c2
ǫµωπn
cb πmx πny πℓz
Bx (x, y, z) = −iA ǫµω 2
sin cos cos ,
π 2 ℓ2 a b d
c2 − d2
ǫµωπm
ca πmx πny πℓz
By (x, y, z) = iA ǫµω 2 2 2
cos sin cos . (7.98)
− πd2ℓ a b d
c2
7.7 Cavity resonators 251
∂Ez
∇⊥ = −∇⊥ (∇⊥ · E⊥ )
∂z
= −∇2⊥ E⊥ − ∇⊥ × (∇⊥ × E⊥ )
iω
= −∇2⊥ E⊥ − ∇⊥ × ẑBz
c
2 ǫµω 2 π 2 ℓ2
= −∇⊥ E⊥ = − 2 E⊥ , (7.99)
c2 d
which leads to the first relation in (7.97). Here we have used the
harmonic time dependence of the fields as well as Bz = 0 for TM
waves in the intermediate step. We have also used the wave equation
for E⊥ in the last step.
It is clear from Eq. (7.96) that, in such a set up, electromagnetic
fields exist only for a single frequency depending on the given values
of m, n, ℓ. This is the behavior of an undamped resonant system such
as an oscillator. As a result, such a set up is called a cavity resonator
(or a resonant cavity). The cavity can also have resonant TE modes.
For the TE modes, we note that Ez = 0 and the additional boundary
conditions have the form
Bz |z=0,d = 0, (7.100)
1 ∂Bz
B⊥ (x) = B⊥ (x, y, z) = ǫµω 2
∇⊥ . (7.103)
π 2 ℓ2 ∂z
c2 − d2
π 2 mℓ
ad πmx πny πℓz
Bx (x, y, z) = −C ǫµω 2
sin cos cos ,
π 2 ℓ2 a b d
c2 − d2
π 2 nℓ
bd πmx πny πℓz
By (x, y, z) = −C ǫµω 2
cos sin cos .
π 2 ℓ2 a b d
c2 − d2
(7.104)
Bz (r, φ, z) = 0,
πℓz
Ez (r, φ, z) = an Jn (hr) cos nφ cos , (7.105)
d
where
r
ǫµω 2 π 2 ℓ2
h= − 2 . (7.106)
c2 d
Furthermore, for the electric fields to vanish at the cylindrical walls,
we must have (see (7.57))
Jn (ha) = 0, (7.107)
7.7 Cavity resonators 253
Ez (r, φ, z) = 0,
πℓz
Bz (r, φ, z) = bn Jn (hr) cos nφ sin , (7.109)
d
with h still defined as in (7.106). However, for TE waves, as we have
already seen in (7.65), the radial derivative of the magnetic field at
the cylindrical walls must vanish leading to
This is clearly defined only for positive times and the energy inside
the cavity decreases as time evolves due to the losses at the walls of
the cavity (as well as possible losses in the dielectric inside).
It is clear from this that, since there is energy loss in a cavity,
if we want to excite a particular resonant mode in the cavity by, say,
introducing an external electromagnetic wave into the cavity through
a small aperture, the system would behave like a damped oscillator
with a time dependent driving force. From the form of the energy U
in (7.113), we note that we can write the form of the electric fields
inside the cavity to have the form
ωr
−i(ω̄r −i 2Q )t
E(t) = E(0) e , (7.114)
ω̄r = ωr + δω.
The quantity, δω, has been introduced to account for the possible
smearing of the resonant frequency due to other effects. Taking the
7.8 Q factor of a cavity 255
|E(0)|2
4π 2
| |
ω̄r ωr ω
2Q
Figure 7.4: The behavior of the absolute square of the electric field
as a function of the frequency ω.
ωr
ω = ω̄r ± . (7.117)
2Q
Consequently, the width of the curve at half the peak value is obtained
to be
ωr
Γ = ∆ω = ,
Q
ωr
or, Q = . (7.118)
Γ
256 7 Wave guides
This shows that the Q factor, indeed, measures the spread out in the
frequency due to Ohmic losses.
As in the case of the wave guide, the power loss at the walls can
be calculated in a simple manner and, for microwave cavities, one can
obtain a Q factor as large as ten thousand. This implies that one can
construct cavities that can show very sharp resonant behavior.
x=a
x = −a
ik dEz (x)
Ex (x) = ǫµω 2
,
− k2 dx
c2
iω dBz (x)
Ey (x) = − ǫµω2 c ,
−k 2 dx
c2
ik dBz (x)
Bx (x) = ǫµω 2
,
− k2 dx
c2
258 7 Wave guides
i ǫµω
c dEz (x)
By (x) = ǫµω 2
. (7.121)
− k2 dx
c2
These equations must hold in both the regions – inside and out-
side the dielectric slab. Let us assume that ǫ1 , µ1 represent the per-
mittivity and the permeability inside the dielectric (−a < x < a)
while ǫ2 , µ2 represent the same quantities outside (|x| > a). Because
of the symmetry in the problem under x ↔ −x, the solutions can
be classified into even and odd ones. Furthermore, we would like the
solutions to be exponentially damped outside the slab while oscilla-
tory inside. With this, we see that we can have four different kinds of
solutions – TM even, TM odd, TE even and TE odd. Let us simply
work out one of the solutions in detail.
Let us consider the TM even solutions. In this case, we have
Bz = 0. Furthermore defining, in the two regions,
r r
ǫ1 µ 1 ω 2 2
ǫ2 µ 2 ω 2
α= 2
−k , γ = k2 − , (7.122)
c c2
where we assume that both α, γ are real and positive, we see that
ω2 ω2
α2 + γ 2 = (ǫ1 µ1 − ǫ2 µ2 ) = ∆ . (7.123)
c2 c2
Furthermore, we note that because of the symmetry of the solutions,
we can restrict ourselves to the positive x-axis only (x ≥ 0). Using
(7.122) and (7.123), we obtain from Eqs. (7.120) and (7.121) (for
x ≥ 0)
(1) (2)
Ez (x) = A cos αx, Ez (x) = C e−γx ,
(1) (2)
Ex (x) = −iA αk sin αx, Ex (x) = iC k
γ e−γx ,
(1) (2)
Ey (x) = 0, Ey (x) = 0,
(1) (2)
Bx (x) = 0, Bx (x) = 0,
(1) (2)
By (x) = −iA ǫ1cα
µ1 ω
sin αx, By (x) = iC ǫ2 µ2 ω
cγ e−γx ,
(7.124)
where the superscripts (1), (2) denote the two regions 0 ≤ x ≤ a and
x ≥ a respectively (because of our choice x ≥ 0).
Since these are even solutions, we can apply the boundary con-
ditions only at x = a (the conditions will then automatically hold at
the boundary x = −a by symmetry). We have already worked out
7.9 Dielectric wave guides (optical fibers) 259
In the present case, since Ey , Bx (and Bz ) are zero in the two regions,
we obtain from the third relation in (7.125) that
1 (1) 1 (2)
B (a) = B (a)
µ1 y µ2 y
ǫ1 ǫ2
or, A sin αa = −C e−γa . (7.127)
α γ
It is easy to see that the first relation in (7.125) leads to the same
relation as (7.127) and, therefore, (7.126) and (7.127) represent essen-
tially the boundary conditions that need to be satisfied. (The second
relation in (7.125) is trivially satisfied.)
Taking the ratio of Eqs. (7.126) and (7.127), we obtain
α γ
cot αa = −
ǫ1 ǫ2
ǫ1
or, α cot αa = − γ. (7.128)
ǫ2
η
3
0
0 1 2 3 4 5 6
ξ
Figure 7.6: The TM even solutions are obtained from the intersections
of η = − ǫǫ21 ξ cot ξ and η 2 + ξ 2 = constant.
r
ǫ1 ∆ω 2 a2
TM (odd) : ξ tan ξ = − ξ2,
ǫ2 c2
r
µ1 ∆ω 2 a2
TE (even) : ξ cot ξ = − − ξ2,
µ2 c2
r
µ1 ∆ω 2 a2
TE (odd) : ξ tan ξ = − ξ2, (7.130)
µ2 c2
2. Calculate the attenuation factor for the TEM waves in the set
up described in the previous question.
a = 7cm, b = 4cm
∇ · (ǫE) = 4πρ,
∇ · B = 0,
∇ × E = ik B,
1 4π 4πσ
∇× B = J − ikǫ E = E − ikǫ E, (8.1)
µ c c
ω
where we have defined, as before, k = c and have identified
J = σ E, (8.2)
263
264 8 Propagation through a conducting medium
n̂
∆ℓ
∆h
Z Z
ds · (∇ × E) = ik ds · B
I Z
or, dℓ · E = ik ds · B. (8.3)
n̂ × E1 | = n̂ × E2 | , (8.4)
where n̂ represents a unit vector normal to the boundary and the re-
striction represents the boundary. If the second medium corresponds
to a perfect conductor, then there is no electric field in the second
medium. Consequently, in such a case, we obtain the boundary con-
dition to be
n̂ × E| = 0. (8.5)
For non-singular field configurations, the right hand side of (8.6) van-
ishes in the limit of vanishing surface area and the boundary condition
takes the form
1 1
n̂ ×
B1 = n̂ × B2 . (8.7)
µ1 µ2
n̂
t̂
Js
the long arm of the surface (see Fig. 8.2). In this case, in the limit
of vanishing surface area, (8.6) leads to
1 1 4π
n̂ × B1 − B2 ∆ℓ = Js ∆ℓ
µ1 µ2 c
1 1 4π
or, n̂ ×
B1 = n̂ × B2 + Js . (8.8)
µ1 µ2 c
z=0
where the incident and the reflected waves have the coordinate de-
pendence given by
(0)
Einc (x, ω) = Einc eikz ,
(0)
Erefl (x, ω) = Erefl e−ikz , (8.11)
with k = ωc . (We have factored out the time dependence for simplic-
ity.)
The electric and the magnetic fields, in a plane wave, are or-
thogonal to the direction of propagation (as well as to each other)
and, therefore, are along the surface of the boundary (tangential to
the surface). As we see from the boundary condition for a perfect
conductor in (8.5),
E(x, y, z = 0, ω) = 0, (8.12)
which determines
(0) (0)
Erefl = −Einc = −E(0) , ẑ · E(0) = 0. (8.13)
Namely, the incident and the reflected electric fields have the same
amplitudes, but are out of phase. Thus, we can write
Once we have the electric fields, the magnetic fields can be ob-
tained through the use of the relation (see (7.76) in vacuum)
B = ±ẑ × E, (8.16)
There are several things to note from Eqs. (8.14) and (8.18) (or
(8.15) and (8.19)). The unique solution to the problem is obtained
using only the boundary condition (8.5) (namely, we do not need (8.9)
to obtain the solution). Furthermore, the electric and the magnetic
fields define standing waves with the electric field vanishing at (z < 0)
kz = −nπ, n = 0, 1, 2, . . . , (8.20)
(2n + 1)π
kz = − , n = 0, 1, 2, . . . . (8.21)
2
The electric field vanishes at the boundary z = 0 simply because the
incident and the reflected electric fields are out of phase. However, we
note that the magnetic field does not vanish at the boundary. This
8.3 Reflection at oblique incidence 269
= 0. (8.22)
The radiated power vanishes simply because the incident and the
reflected components of the magnetic field are in phase while those
for the electric field are out of phase. Physically, this is clear since
there cannot be any transmitted wave inside a perfect conductor and,
therefore, there cannot be any power loss.
Let us note that the non-vanishing of the magnetic field, at the
boundary, immediately shows that there must be a surface current
in the conductor. In fact, from (8.9), we conclude that (µ = 1 and
ẑ · E(0) = 0)
4π
Js = n̂ × B| = − ẑ × B|
c
= −2ẑ × ẑ × E(0) = 2E(0) ,
c (0)
or, Js = E . (8.23)
2π
This can, therefore, be thought of as the reason for the change in the
phase of the reflected component of the electric field.
x
θi θr
where
where
Therefore, the total electric field in this region has the form
n̂ × E(x, ω)|z=0 = 0,
(0) (0)
or, ẑ × Einc eiki sin θi x + ẑ × Erefl eikr sin θr x = 0. (8.29)
8.3 Reflection at oblique incidence 271
ki = kr = k,
θi = θr = θ,
(0) (0)
ẑ × Einc + Erefl = 0. (8.30)
This shows that the wave numbers for the incident as well as the
reflected waves are the same (they are in the same region) and that
the angle of incidence is equal to the angle of reflection, as is also
true in the case of reflection from a dielectric surface. However, here
we have no transmission.
The last relation in (8.30) allows for two possibilities. First, we
can have the electric fields along the y-axis - perpendicular to the
plane of incidence. In this case, for the last relation in (8.30) to hold,
we must have
(0) (0)
Erefl = −Einc = −E (0) ŷ. (8.31)
Thus, in this case, we see that while the electric field is perpendicular
to the plane of incidence, the magnetic field is parallel to it (lies in
the plane of incidence). Furthermore, we have
so that the tangential component of the magnetic field does not vanish
at the boundary. This, therefore, leads to a surface current (see (8.9))
of the form
c c
Js = n̂ × B| = ( −ẑ × B|)
4π 4π
cE (0)
= ŷ cos θ eikx sin θ . (8.36)
2π
dP c
= Re ẑ · (E × B∗ )
da 8π
c
= Re 2i|E (0) |2 cos θ sin(2kz cos θ)
8π
= 0. (8.37)
(0)
Einc = E (0) (x̂ cos θ − ẑ sin θ),
(0)
Erefl = −E (0) (x̂ cos θ + ẑ sin θ). (8.38)
As a result, we obtain
In this case, we note that the magnetic field is orthogonal to the plane
of incidence while the electric field is parallel to it.
Let us note that
so that the tangential component of the magnetic field does not vanish
on the boundary. Consequently, it leads to a surface current of the
form
c c
Js = n̂ × B| = ( −ẑ × B|)
4π 4π
cE (0)
= x̂ eikx sin θ . (8.43)
2π
We can also calculate the power radiated per unit area in region
z < 0, which takes the form
dP c
= Re ẑ · (E × B∗ )
da 8π
c
= Re 2i|E (0) |2 cos θ sin(2kz cos θ)
8π
= 0. (8.44)
This shows that all the energy that is incident is reflected back and
there is no net loss of energy.
Finally, let us consider, for simplicity, the first solution (where
the electric field is orthogonal to the plane of incidence) and note
some qualitative features which hold for both the solutions. We see,
274 8 Propagation through a conducting medium
from Eq. (8.32) that the electric field forms standing waves along the
z-axis with nodes at (z < 0)
k̄ = k sin θ. (8.46)
λ
λ̄ = , (8.47)
sin θ
corresponding to a propagation velocity
ω c
v̄ = = . (8.48)
k̄ sin θ
Thus, we see that the wave length as well as the velocity of propa-
gation are larger than their corresponding values in vacuum. This is
very similar to the behavior we have seen in wave guides. In fact,
suppose we add a parallel, perfectly conducting surface at z = −b,
then, from the solution for the electric field in (8.32), we see that
this has to satisfy the boundary condition (8.5) at the new boundary
leading to
which gives
p
k̄ = k sin θ = k 2 − k 2 cos2 θ
r
ω 2 π 2 n2
= − 2 . (8.50)
c2 b
This shows that for ω < πcnb , there will be no propagation, while for
ω > πcn
b , there will be propagation of waves. This is, in fact, exactly
what we have seen for a wave guide. However, here we only have a
pair of parallel conducting surfaces. As we have discussed earlier, a
pair of parallel conducting plates can also guide waves and, among
other things, has a TEM mode of propagation.
8.4 Reflection from a good conducting surface 275
∇ · D = 4πρ,
∇ · B = 0,
iω
∇×E= B,
c
1 4π iω 4πσ iωǫ
∇× B = J− D= − E. (8.51)
µ c c c c
We have used here the relation
J = σ E, (8.52)
as well as
D = ǫE, (8.53)
∂ρ 4πσ
= −σ∇ · E = − ρ,
∂t ǫ
4πσt
or, ρ(t) ∼ e− ǫ . (8.55)
This shows that the charge density dissipates with a time scale
ǫ
τ= . (8.56)
4πσ
On the other hand, the only meaningful time scale in a harmonic
problem is ω1 and so, we conclude that if
1
τ≪
ω
4πσ
or, ≫ 1, (8.57)
ǫω
the charge will dissipate quite rapidly and this is what defines a good
conductor. In such a medium, therefore, we can set the charge density
to zero. Note, from (8.55), that for a perfect conductor, σ → ∞ so
that there is no charge density produced.
Returning to the Maxwell’s equations, (8.51), and taking the
curl of the last two equations, we note that they lead to
2 ǫµω 2 4πiσ
∇ E+ 2 1+ E = 0,
c ǫω
ǫµω 2 4πiσ
∇2 B + 1+ B = 0. (8.58)
c2 ǫω
These are complex equations which we can think of as the wave equa-
tions that we have studied earlier, if we allow for a complex permit-
tivity of the form
4πiσ 4πiσ
ǫeff = ǫ 1 + =ǫ+ . (8.59)
ǫω ω
provided
2 ǫµω 2
2 4πiσ
K =K = 2 1+ . (8.61)
c ǫω
This shows that in a conducting medium, the wave number, in gen-
eral, becomes complex with a real and an imaginary part. In partic-
ular, if we have a good conductor, 4πσ
ǫω ≫ 1, we obtain
√ r √
ǫµω 4πiσ 2πσµω
K≈ = (1 + i) = kr + iki , (8.62)
c ǫω c
where we have used
√ (1 + i)
i= √ ,
2
and have identified
√
2πσµω
kr = ki = . (8.63)
c
Thus, in a good conductor, the solutions of the Maxwell’s equa-
tions traveling along the z-axis, for example, lead to electric and
magnetic fields of the forms
Ez = 0 = Bz , (8.66)
iσ E(0) iσ
= = E| , (8.71)
K K
where the restriction refers to the surface assumed to be at z = 0. In
analogy with electric circuits, the ratio of the tangential component
of the electric field at the surface to the surface current density is
defined as the surface impedance of the medium, namely,
E| = Zs Js ,
r
iK 2πµω 1
or, Zs = − = (1 − i) = (1 − i) , (8.72)
σ σc2 σδ
where δ represents the skin depth defined in (8.65). The real part of
the surface impedance can be thought of as the surface resistance of
the medium,
r
1 2πµω
Rs = = . (8.73)
σδ σc2
8.4 Reflection from a good conducting surface 279
Note that when σ → ∞, the surface resistance vanishes (as does the
surface impedance). It is now straightforward to see from (8.68) and
(8.72) that
c c
Z= Zs = (1 − i) . (8.74)
4πµ 4πµσδ
With these, we are now ready to analyze reflection from a good
conducting surface. Let us assume that the conducting surface is at
z = 0 and that an electromagnetic wave is incident from vacuum in
z < 0 at an angle of incidence θi in the x − z plane as shown in Fig.
8.5. In this case, we do expect a reflected wave in the region z < 0
with angle of reflection θr as well as a transmitted wave in the region
z > 0 with the angle of transmission θt . Of course, the transmitted
wave will be highly attenuated and we assume that the thickness of
the conducting medium is much larger than the skin depth so that
the transmitted wave will practically be a surface wave. We can
parameterize the electric fields associated with the three components
as
θt
x
θi θr
where we have used the results from our earlier analysis that the
280 8 Propagation through a conducting medium
(ẑ × Ei )eiki x sin θi + (ẑ × Er )eikr x sin θr = (ẑ × Et )eikt x sin θt . (8.77)
θi = θr = θ,
ki = kr = k,
k
sin θt = sin θ
kt
ẑ × Et = ẑ × (Ei + Er ). (8.78)
We note that the first of these relations tells us the familiar result
that the angle of reflection is the same as the angle of incidence while
the third gives Snell’s law, namely
sin θt k 1
= = = Zt , (8.79)
sin θ kt nt
where nt represents the index of refraction of the conducting medium
and we have used k = ωc as well as (8.70). The important thing to note
here is that, in this case, the index of refraction is complex signifying
dissipation (absorption) of energy in the medium. Furthermore, for
a good conductor (see (8.69)),
Zt ≪ 1, (8.80)
so that we expect
θt ≈ 0. (8.81)
conditions, that all the electric fields will have the same polarization,
namely,
Ei + Er = Et ,
1
(Ei − Er ) cos θ = Et cos θt . (8.84)
Zt
These can be solved to obtain the coefficients of reflection and refrac-
tion in a straightforward manner from
The magnetic fields are obtained from this to have the forms
also discussed earlier along general lines when we studied the Poynt-
ing vector). Let us recall that the momentum density associated with
electromagnetic fields is related to the Poynting vector as (see (6.92))
ǫµ
p= Re S. (8.90)
c2
For harmonic fields, therefore, we can define a time averaged momen-
tum density as
ǫµ c ǫ
p= 2 Re (E × H∗ ) = Re (E × B∗ ). (8.91)
c 8π 8πc
The total momentum exerted by the EM waves on a surface of area
“a” in a time interval ∆t, therefore, follows to be
ǫa(c∆t)
∆p = p a(c∆t) = Re (E × B∗ ). (8.92)
8πc
This leads to a pressure exerted by an electromagnetic wave on a
surface of the form
F 1 ∆p ǫ
P= = = pc = Re (E × B∗ ). (8.93)
a a ∆t 8π
Let us now analyze all of this in the case of an electromagnetic
wave (in vacuum) at normal incidence on a perfectly conducting sur-
face at z = 0 that we have already studied. We have seen that, in
this case, the incident electric and the magnetic fields have the forms
(see (8.11) and (8.17))
It follows, therefore, from (8.93) that such a wave will exert a pressure
on the surface at z = 0 of the form (in vacuum ǫ = 1)
1
P= Re E(0) × (ẑ × (E(0) )∗ )
8π
1
= ẑ |E(0) |2 . (8.95)
8π
This shows that the incident wave exerts a pressure on the surface
along the z-axis. Furthermore, if we have a perfectly conducting
surface, the incident wave is totally reflected doubling the pressure
so that, for a perfectly conducting surface, we have
1
P total = 2P = ẑ |E(0) |2 . (8.96)
4π
284 8 Propagation through a conducting medium
MHz and 1 GHz. Use ǫ = 80, σ = 3.6 × 1010 /sec for sea water,
ǫ = 80, σ = 4.5 × 107 /sec for fresh water and ǫ = 15, σ =
9 × 107 /sec for “good” earth.
with
i ∂Ez i ∂Ez
Hx = , Hy = − ,
k ∂y k ∂x
∂ 2 Ez ∂ 2 Ez
+ + k 2 Ez = 0,
∂x2 ∂y 2
and
ii) E = (Ex , Ey , 0), H = (0, 0, Hz ),
with
i ∂Hz i ∂Hz
Ex = − , Ey = ,
k ∂y k ∂x
∂ 2 Hz ∂ 2 Hz
+ + k 2 Hz = 0.
∂x2 ∂y 2
Radiation
1 ∂Φ
∂µ Aµ = + ∇ · A = 0, (9.2)
c ∂t
the scalar potential is related to the vector potential and it is sufficient
for us to study only the (three dimensional) vector potential. From
(9.1) we have
′
Z J(x′ , t′ )δ t′ − t + |x−x |
1 c
A(x, t) = d3 x′ dt′ ′
. (9.3)
c |x − x |
The space integral here is over the volume V which contains (charges)
currents and if we are interested in the fields at points which are far
away from the volume containing the (charges) currents, we can make
an expansion much like in the static case. This would give rise to the
287
288 9 Radiation
we have
Z iω|x−x′ |
1 e c
A(x) = d3 x′ J(x′ ) . (9.8)
c |x − x′ |
It is clear that all the variables (Φ, A) and (E, B) will have the same
harmonic time dependence which can be factored out. The space
dependent magnetic field can, of course, be determined from (9.8) as
while the electric field in the region outside the volume containing
charges and currents (in vacuum) can be obtained from the last equa-
tion of Maxwell in (7.1) as
ic i
E(x) = ∇ × B(x) = ∇ × B(x). (9.10)
ω k
Substituting this into the expression for the vector potential in (9.8),
we have
Z ′ ′
1 ei|k|(r−r cos θ )
A(x) ≈ d3 x′ J(x′ )
c (r − r ′ cos θ ′ )
Z ′ ′
!
eikr 3 ′ ′ e−ikr cos θ
= d x J(x ) ′ θ′
. (9.12)
cr 1 − r cos
r
kd << 1, (9.13)
parenthesis in (9.12) as
′ ′
!
e−ikr cos θ i
r ′ cos θ ′
= 1 + (−ik)(1 + )r ′ cos θ ′
1− r kr
(−ik)2 2i 2
+ (1 + − )(r ′ cos θ ′ )2 + · · ·
2! kr (kr)2
∞
X
= Xn (r, k)(r ′ cos θ ′ )n , (9.14)
n=0
(m)
and an ’s (m ≤ n) are numerical constants. Since Xn (r, k) is inde-
pendent of the variables of integration, it can be taken outside the
integral in (9.12) leading to an expansion of the vector potential of
the form
∞ Z
eikr X 3 ′ ′ ′ ′ n
A(x) = Xn (r, k) d x J(x )(r cos θ ) . (9.16)
cr n=0
In other words, in the near zone, the vector potential oscillates har-
monically with time. Otherwise, it has a purely static character (no
291
propagation). (Here we have used the fact that in the near zone,
eikr ≈ 1.) We note that, since r ′ ≪ r, the dominant term in (9.18) is
the zeroth order term leading to
Z
1
A(x) = d3 x′ J(x′ ). (9.19)
cr
(0)
Here, we have used the fact that ã0 = 1 (see (9.15)).
In contrast, in the far zone (kr >> 1), we see from Eq. (9.15)
that we have
(−ik)n
Xn (r, k) ≈ , (9.20)
n!
so that the vector potential in (9.16) takes the form
∞ Z
eikr X (−ik)n
A(x) = d3 x′ J(x′ )(r ′ cos θ ′ )n
cr n=0 n!
Z ∞
X
eikr 3 ′ ′ (−ikr ′ cos θ ′ )n
= d x J(x )
cr n=0
n!
Z
eikr ′ cos θ ′
= d3 x′ J(x′ ) e−ikr . (9.21)
cr
We note that, since kd ≪ 1 (see Eq. (9.19)), the dominant term in
(9.21) leads to
Z
eikr
A(x) = d3 x′ J(x′ ). (9.22)
cr
We see that in the far zone, the vector potential is represented
by a spherically outgoing wave. This is because the phase of the
complete vector potential is given by (ωt − kr). As we have learnt
earlier, wave fronts are described by surfaces of constant phase. Thus,
at any given time, the surfaces of constant phase are given by
ωt − kr = constant
or, kr = constant, (9.23)
which are spherical surfaces of radius r. This is like the plane wave
solutions of radiation that we studied earlier, but the waves, in the
present case, are spherically outgoing. The other thing to note is
that, in the far zone, the higher order terms in the expansion (of the
exponential) fall off rapidly, simply because kd << 1. Consequently,
292 9 Radiation
in the study of radiation, only the first term in the series contributes
significantly. We can calculate the electric and the magnetic fields
from the potentials (see (9.9), (9.10)). However, we will not go into
it now except for the observation that, asymptotically, the electric
and the magnetic fields fall off as 1r which is precisely the behavior of
radiation fields.
q b
d
−q b
Let us further assume that the size of the dipole, d, is very small
and that the time dependence of the charge is harmonic as before.
Namely,
Z
= −iω d3 x′ x′ ρ(x′ )
= −iωp, (9.27)
where we have used the fact that the currents are contained within
the volume and, consequently, n̂ · J = 0 on the surface. Furthermore,
we have used the continuity equation
∂ρ
+ ∇ · J = 0, (9.28)
∂t
as well as the definition of the electric dipole moment p for an arbi-
trary charge distribution (see (2.49))
Z
p = d3 x x ρ(x), (9.29)
eikr eikr
A(x) = (−iωp) = −ikp . (9.30)
cr r
This allows us to calculate the magnetic field directly as
∂ eikr
B(x) = ∇ × A(x) = −ik (r̂ × p)
∂r r
ikr
i e
= k 2 (r̂ × p) 1 + . (9.31)
kr r
294 9 Radiation
On the other hand, the electric field can be calculated from the Am-
pere’s law, namely, (since we are interested in points far away from
the sources, J = 0, see also (9.10))
1 ∂E
∇×B= , (9.32)
c ∂t
which gives
i
E(x) = (∇ × B(x))
k
2 2i 2
= −k (r̂ × (r̂ × p)) 1 + − (9.33)
kr (kr)2
ikr
i 1 e
−(θ̂ × (θ̂ × p) + φ̂ × (φ̂ × p)) − ,
kr (kr)2 r
where we have used the fact that the unit vectors in spherical coor-
∂r̂ ∂r̂ ∂r̂
dinates are not fixed. In fact, while ∂r = 0, ∂θ = θ̂ and ∂φ = φ̂ sin θ.
Furthermore, recalling that p is along the z-axis, we can simplify this
and write (recall that, in spherical coordinates, ẑ = r̂ cos θ − θ̂ sin θ)
2
E(x) = −k (r̂ × (r̂ × p))
i 1 eikr
+ (3r̂(r̂ · p) − p) − . (9.34)
kr (kr)2 r
r̂ · B = 0. (9.35)
However, it follows from (9.34) that the electric field is not transverse
to the radial vector in general. We note that the magnetic field
in (9.31) has two terms – one behaving as 1r which dominates for
large r, while the second depends on the radial coordinate as r12 and,
therefore, contributes significantly for small r. The second is known
as the static (or induction) term while the first is called the radiation
term for reasons that will become clear shortly. Similarly, we see from
(9.34) that the electric field has three terms out of which the r13 term
gives the most contribution for small r (and is called the static field)
while the 1r term dominates at large distances and is known as the
radiation field.
9.1 Electric dipole radiation 295
We note that in the near zone, kr << 1 and we can write the
electric and the magnetic fields as (eikr ≈ 1)
ik(r̂ × p)
B(x) = ,
r2
3r̂(r̂ · p) − p
E(x) = . (9.36)
r3
This shows that in the near zone, the magnetic field is what would be
obtained from the Biot-Savart law for a current element (except for
the trivial time dependence that has been factored out). Similarly,
in this region, the electric field is that of a static dipole (except for
the trivial harmonic time dependence which we have factored out).
Furthermore, since in this region kr << 1, the electric field dominates
over the magnetic field. In the far (radiation) zone, on the other hand,
we note that kr >> 1 and we can approximate the electric and the
magnetic fields in (9.31) and (9.34) as
eikr
B(x) = k 2 (r̂ × p) ,
r
eikr
E(x) = −k 2 (r̂ × (r̂ × p)) = −r̂ × B(x). (9.37)
r
The last relation in (9.37) can also be written as
which is the relation for traveling EM waves that we have seen earlier,
for example, in (6.30). From (9.37) we see that both the E and the B
fields fall off as 1r and are transverse to the direction of propagation
(as well as to each other) as is expected of radiation fields. We see
that the radiation terms are new compared to the behavior of static
distributions. We note that, in the static limit (ω = 0 or k = 0),
the magnetic field identically vanishes everywhere. Furthermore, the
radiation component of the electric field also vanishes in this limit.
Thus, we see that radiation is an essential feature associated with
time varying charges and currents.
Incidentally, although we have discussed a very simple system, it
is behind many physical systems such as antennas. In this discussion,
we have only retained the lowest order term which, as we see, leads to
an electric dipole description (in the static zone). The higher order
terms in the expansion, similarly, can be shown to give rise to the
description of a magnetic dipole, electric quadrupole etc.
296 9 Radiation
da
R
dΩ
these components of the fields are known as the radiation fields. They
lead to a radiation of power to infinity (which cannot come back to
the system).
If we look at the forms of the radiation components of the elec-
tric and the magnetic fields in Eq. (9.37) and recall that the dipole
moment is along the z-axis (by assumption) with ẑ = r̂ cos θ − θ̂ sin θ,
it follows that in the limit R → ∞,
dP c 4 2 2
= k |p| sin θ. (9.43)
dΩ 8π
This shows that the radiated power is highly directional (not uniform,
but depends on θ), peaking at θ = π2 . This also leads to a total
integrated average radiated power of the form
Z Z
dP c 4 2
Ptotal = dΩ = k |p| dΩ sin2 θ
dΩ 8π
c 4 2 8π ck 4 |p|2
= k |p| = . (9.44)
8π 3 3
Using k = ωc and the fact that for this simple system we can think of
(see also (9.24), (9.25))
|I0 |2 d2
|p|2 = |q0 |2 d2 = , (9.45)
ω2
we can also write the (time averaged) total power radiated to infinity,
(9.44), as
ω 2 d2 2 2ω 2 d2 2
Ptotal = |I 0 | = I . (9.46)
3c3 3c3 rms
Here, I0 represents the peak current while Irms denotes the effective
current (root mean square). This suggests that we can associate a
radiation resistance with the dipole from the standard definitions as
2ω 2 d2
Rrad = . (9.47)
3c3
a
Figure 9.3: An alternating current loop of radius a in the x − y plane.
I(t) = I0 e−iωt ,
π
then, it follows that (note that z ′ = 0 or equivalently θ ′ = 2 because
of the delta function constraint)
1
|x − x′ | = (r 2 + ρ′2 − 2rρ′ sin θ cos(φ − φ′ )) 2
≈ r − ρ′ sin θ cos(φ − φ′ ). (9.52)
Here, we have assumed that the size of the current loop is extremely
small compared with the distance of the point of observation (r ≫ ρ′ )
which allows us to keep only the lowest order terms in the expansion.
Furthermore, let us assume that we are interested only in the
far field approximation (namely, in the radiation zone). In that case,
r is very large (kr ≫ 1) and we can approximate the denominator
by the lowest order term. However, we have to be careful with the
exponent. Let us also note that we can write φ̂′ = −x̂ sin φ′ + ŷ cos φ′ .
With these, the expression (9.50) for large values of r takes the form
(and we assume that a is very small, ka ≪ 1)
Z2π
I0 aeikr ′
A(x) ≈ dφ′ (−x̂ sin φ′ + ŷ cos φ′ ) e−ika sin θ cos(φ−φ )
cr
0
Z2π
I0 aeikr
≈ dφ′ (−x̂ sin φ′ + ŷ cos φ′ )
cr
0
× 1 − ika sin θ cos(φ − φ′ )
Z2π
ikI0 a2 sin θeikr
=− dφ′ −x̂ sin φ + sin(2φ′ − φ)
2cr
0
+ŷ cos φ + cos(2φ′ − φ)
We note that the form of (9.53) is, in fact, quite similar to the case
of the electric dipole radiator studied earlier, with the electric dipole
moment replaced by the magnetic dipole moment (along with various
constants needed for dimensional reasons).
We note that, unlike the electric dipole case, here the vector
potential in the radiation zone is transverse to the direction of prop-
agation, namely,
r̂ · A(x) = 0. (9.54)
B(x) = ∇ × A
k 2 I0 πa2 sin θeikr 1
= −θ̂ +O . (9.55)
cr r2
Similarly, the electric field takes the form
i
E(x) = ∇×B
k
k 2 I0 πa2 sin θeikr 1
= φ̂ +O . (9.56)
cr r2
This allows us to calculate the (time averaged) power radiated per
unit solid angle in a given direction through the surface of a large
sphere to be (see (9.42))
dP
= R2 r̂ · S
dΩ
(k 2 I0 πa2 )2 k 4 |m0 |2
= sin2 θ = sin2 θ. (9.57)
8πc 8πc
We see that the angular dependence of the radiated power, in this
case, is the same as that in the case of the electric dipole radiation
in (9.43). The total power radiated through the surface of a large
sphere, averaged over a cycle, is then obtained to be
π 2 k 4 a4 2 2π 2 k 4 a4 2
Ptotal = I0 = Irms , (9.58)
3c 3c
so that we can identify the radiation resistance associated with this
system to be
2π 2 k 4 a4
Rrad = . (9.59)
3c
9.3 Center-fed antennas 301
d
2
d
Figure 9.4: A dipole antenna (on the left) and a monopole antenna
(on the right).
With this, we can repeat the calculation of the earlier sections. There
is a simpler method for obtaining the relevant results by noting that
the average current in the antenna, in this case, follows to be
I0
Iavg = , (9.61)
2
where we are assuming that I0 represents the peak current at the
center of the antenna.
With this, we can now extend our previous analysis simply by
letting I0 → I20 or Irms → Irms
2 . Thus, for example, for the center-fed
short dipole antenna we obtain (see (9.46)),
ω 2 d2 2
Ptotal = I . (9.62)
6c3 rms
Correspondingly, the radiation resistance of the center-fed dipole an-
tenna is smaller by a factor of 4,
ω 2 d2
Rrad = . (9.63)
6c3
On the other hand, for a short center-fed monopole antenna,
since only one half of the antenna really radiates, we have
1 ω 2 d2 2 ω 2 d2 2
Ptotal = × I = I , (9.64)
2 6c3 rms 12c3 rms
and correspondingly, the radiation resistance
ω 2 d2
Rrad = , (9.65)
12c3
is even smaller.
These theoretical predictions work quite well and can be checked
experimentally to hold for short antennas satisfying d ≪ λ where
λ denotes the wavelength of the signal. In fact, they hold up to
d ≤ λ4 . However, in transmitting radio waves (for radio waves ν =
300Hz −3000GHz, λ = 100Km −1mm), where antennas are primarily
used, it is found that the transmission is better if the dimensions of
the antenna were of the order of the wave length, d ∼ λ (see Fig. 9.5).
Of course, these are no longer short antennas and the analysis has
to be carried out more carefully. The difficulty really lies in knowing
the distribution of the current along the length of the antenna. If
this distribution is known, the calculation of the fields can be carried
out in principle. Following studies of the transmission lines where
9.3 Center-fed antennas 303
d z
z= 2
z = − d2
Namely, we are assuming the antenna to lie along the z-axis and the
current to be sinusoidally varying with z such that the maximum is
at the center and the ends have vanishing current. This translates to
a current density of the form
where we identify
d d
J(x) = ẑ δ(x) δ(y) I0 sin k − |z| , |z| ≤ . (9.68)
2 2
1 1
′
≈ . (9.70)
|x − z ẑ| r
Using these in (9.69), we obtain for large r (the other term sin(kz ′ cos θ)
coming from the exponential vanishes by anti-symmetry in z ′ )
d
Z2
2I0 eikr
A(x) ≈ ẑ dz ′ cos kz ′ cos(kz ′ cos θ)
cr
0
d
Z2
I0eikr
= ẑ dz ′ cos kz ′ (1 − cos θ) + cos kz ′ (1 + cos θ)
cr
0
d
I0 eikr sin kz ′ (1 − cos θ) sin kz ′ (1 + cos θ) 2
= ẑ +
cr k(1 − cos θ) k(1 + cos θ) 0
π
I0 eikr 1 1
= ẑ + cos cos θ
ckr 1 − cos θ 1 + cos θ 2
2I0 eikr cos π2 cos θ
= ẑ . (9.73)
ckr sin2 θ
Exercise. Show that, for an antenna of arbitrary length d, the vector potential has
the large distance behavior given by
kd
2I0 eikr cos 2
− cos kd2
cos θ
A(x) = −ẑ .
ckr sin2 θ
Recalling that ẑ = r̂ cos θ−θ̂ sin θ, we can now calculate the magnetic
9.3 Center-fed antennas 305
B(x) = ∇ × A(x)
!
∂ θ̂ ∂ φ̂ ∂
= r̂ + +
∂r r ∂θ r sin θ ∂φ
2I0 eikr cos π2 cos θ
× (r̂ cos θ − θ̂ sin θ)
ckr sin2 θ
2iI0 eikr cos π2 cos θ 1
= −φ̂ +O . (9.74)
cr sin θ r2
This shows that, at large distances, the dominant term in the mag-
netic field is the radiation term and is along the φ̂ direction. The
electric field can also be calculated similarly and for large distances
has the form
i
E(x) = ∇×B
k
i ∂ 2iI0 eikr cos π2 cos θ 1
= − r̂ × φ̂ +O
k ∂r cr sin θ r2
2iI0 eikr cos π2 cos θ 1
= −θ̂ +O . (9.75)
cr sin θ r2
dP
= R2 r̂ · S
dΩ
I02 cos2 π2 cos θ
= . (9.76)
2πc sin2 θ
cycle, to be
π
Z2 π
2I02 cos2 2 cos θ
Ptotal = dθ
c sin θ
0
π
Z 2
I2 1 + cos(π cos θ)
= 0 dθ . (9.77)
c sin θ
0
Z1
I2 1 1
= 0 dx (1 + cos πx) +
2c 1+x 1−x
0
Z1
I2 1 + cos πx
= 0 dx . (9.78)
2c 1+x
−1
Z2π
I2 1 − cos t
Ptotal = 0 dt
2c t
0
I02
= (C + ln 2π − Ci(2π)) . (9.79)
2c
Here, C ≈ 0.577 is the Euler constant and Ci(x) is known as the
cosine integral defined as (see, for example, Gradshteyn and Ryzhik)
Z∞ Zx
cos t 1 − cos t
Ci(x) = − dt = C + ln x − dt . (9.80)
t t
x 0
Gd = 10 log10 gd . (9.84)
There are also other measures for the antenna gain, but we will not
get into these details. Let us simply note here that, by a clever choice
of an array of antennas, the antenna gain can be enhanced quite a
bit.
308 9 Radiation
1 ∂2ψ
∇2 ψ − = 0, (9.85)
c2 ∂t2
where, for simplicity, we are assuming wave propagation in vacuum.
The conventional way one solves this equation is by Fourier trans-
forming the solution in the time variable
Z
ψ(x, t) = dω e−iωt ψ(x, ω). (9.86)
into (9.87), we find that the Helmholtz equation separates into two
equations of the forms
2
d 2 d 2 ℓ(ℓ + 1)
+ +k − Rℓ (r, ω) = 0,
dr 2 r dr r2
(9.89)
1 ∂ ∂1 ∂2
− sin θ 2 +Yℓ,m = ℓ(ℓ + 1)Yℓ,m .
sin θ ∂θ sin θ ∂φ2
∂θ
Here, ℓ takes positive integer values including zero while, for a given
ℓ, we have m = −ℓ, −ℓ + 1, . . . , ℓ (as we know from the study of
angular momentum in quantum mechanics). The angular functions
Yℓ,m , the spherical harmonics, are eigenstates of L2 and Lz .
The radial equation in (9.89) is the equation for the spherical
Bessel functions and the two independent solutions can be written as
the spherical Bessel functions and the spherical Neumann functions
defined as
π 1
2
jℓ (x) = Jℓ+ 1 (x),
2x 2
π 1
2
ηℓ (x) = Nℓ+ 1 (x). (9.90)
2x 2
The spherical Bessel functions jℓ (x) are regular at the origin while
the spherical Neumann functions ηℓ (x) diverge. An alternative way to
write the solutions is in terms of the spherical Hankel functions, which
are defined as linear combinations of the spherical Bessel functions
and the spherical Neumann functions, namely,
(1)
hℓ (x) = jℓ (x) + iηℓ (x),
(2)
hℓ (x) = jℓ (x) − iηℓ (x). (9.91)
From the fact that the spherical Bessel functions (as well as the spher-
ical Neumann functions) are real, it follows that the two spherical
Hankel functions are complex conjugates of each other. We note that
either of the sets in (9.90) or (9.91) can be thought of as an inde-
pendent set of solutions for the spherical Bessel equation in (9.89).
Thus, we can write the most general radial solution of the form
(1) (1) (2) (2)
Rℓ (r) = aℓ hℓ (kr) + aℓ hℓ (kr), (9.92)
(1,2)
where aℓ are coordinate independent constants. The full solution
for the Helmholtz equation can now be written in the form
X (1) (1) (2) (2)
ψ(x, ω) = aℓ hℓ (kr) + aℓ hℓ (kr) Yℓ,m (θ, φ). (9.93)
ℓ,m
310 9 Radiation
There are several things to note here. First, the spherical har-
monics are normalized so that
Z
∗
sin θdθ dφ Yℓ,m (θ, φ)Yℓ′ ,m′ (θ, φ) = δℓℓ′ δmm′ . (9.94)
r̂ · L = 0. (9.98)
L × L = iL. (9.99)
r̂ · (L × L) = 0. (9.100)
∇ · E = 0, ∇ · B = 0,
(9.101)
∇ × E = − 1c ∂B
∂t , ∇×B= 1 ∂E
c ∂t .
9.4 Multipole expansion 311
∇ · B = 0,
∇2 + k 2 B = 0,
i
E= ∇ × B. (9.103)
k
Here, we treat B as the independent field. Alternatively, if we elimi-
nate the magnetic field from (9.102), then we obtain
∇ · E = 0,
∇2 + k 2 E = 0,
i
B = − ∇ × E, (9.104)
k
where E represents the independent field.
Both the sets of equations in (9.103) or (9.104) are equivalent
and also equivalent to the Maxwell’s equations in (9.102) and we
note that the independent field in (9.103) or (9.104) can be solved
by solving a Helmholtz equation. However, in the present case, the
dynamical variable (the electric or the magnetic field) is a vector,
as opposed to the earlier case where ψ was a scalar function. Cor-
respondingly, the solutions can be expressed as before in terms of
spherical Hankel functions and spherical harmonics, but with vector
coefficients. Thus, for example, the solutions for B in (9.103) can be
written as
X (1) (1) (2) (2)
B(x) = aℓ hℓ (kr) + aℓ hℓ (kr) Yℓ,m (θ, φ), (9.105)
ℓ,m
(1,2)
where aℓ now represent arbitrary vector coefficients. These coeffi-
cients are arbitrary except that the magnetic field has to be transverse
312 9 Radiation
= 0. (9.106)
(1,2)
Since hℓ represent independent solutions of the spherical Bessel
equation, it follows that, for (9.106) to hold, we must have indepen-
dently,
X (i) (i)
∇· aℓ hℓ (kr)Yℓ,m (θ, φ) = 0, i = 1, 2. (9.107)
ℓ,m
Decomposing the gradient into its radial and angular parts (see (9.95)),
∂ i
∇ = r̂ − r̂ × L, (9.108)
∂r r
we obtain from (9.107)
(i)
!
X (i) dhℓ (kr) i (i) (i)
r̂ · aℓ Yℓ,m − hℓ (kr) r̂ · (L × aℓ )Yℓ,m = 0,
dr r
ℓ,m
(9.109)
The vanishing of the first term in (9.109) is, of course, obvious from
(9.98). The vanishing of the second term follows from (9.100), namely,
r̂ · (L × L) = 0. (9.111)
i
E(x) = ∇ × B. (9.112)
k
9.4 Multipole expansion 313
i
B(x) = − ∇ × E, (9.113)
k
which leads to the magnetic multipole fields. The reason for this
nomenclature will become clear shortly.
Since the combination (LYℓ,m ) arises frequently in the study of
electrodynamics, it is given a special name, vector spherical harmon-
ics, and is defined such that it is normalized, namely,
(
√ 1 (LYℓ,m (θ, φ)), ℓ 6= 0,
ℓ(ℓ+1)
Yℓ,m (θ, φ) = (9.114)
0, ℓ = 0.
The fact that it is normalized follows from
Z
∗
sin θdθ dφ Yℓ,m (θ, φ) · Yℓ′ ,m′ (θ, φ)
Z
1
=p sin θdθ dφ (L† Yℓ,m
∗
) · (LYℓ′ ,m′ )
ℓ(ℓ + 1)ℓ′ (ℓ′ + 1)
Z
1 ∗
=p sin θdθ dφ Yℓ,m (θ, φ)(L2 Yℓ′ ,m′ (θ, φ))
ℓ(ℓ + 1)ℓ′ (ℓ′ + 1)
s Z
ℓ′ (ℓ′ + 1) ∗
= sin θdθ dφ Yℓ,m (θ, φ)Yℓ′ ,m′ (θ, φ)
ℓ(ℓ + 1)
i
E(x) = ∇ × B, (9.116)
k
314 9 Radiation
i
B(x) = − ∇ × E. (9.117)
k
An arbitrary electric field can, of course, be written as a linear su-
perposition of the electric and the magnetic multipole fields.
Xk
Yℓ,m
=− L . (9.119)
ℓ r ℓ+1
ℓ,m
Here, we have used the fact that L is an angular operator which does
not act on the radial coordinate as is clear from (9.95). The behavior
of the electric field in the near zone now follows from (9.116) to be
Xi
i Yℓ,m
E(x) = ∇ × B → − ∇×L . (9.120)
k ℓ r ℓ+1
ℓ,m
(∇ × L)i = −i (∇ × (r × ∇))i
= −iǫijk ǫkst ∇j rs ∇t
= −i (δis δjt − δit δjs ) ∇j rs ∇t
= −i (∇j ri ∇j − ∇j rj ∇i )
= −i δij ∇j + ri ∇2 − 3∇i − rj ∇j ∇i
= −i ri ∇2 − 2∇i − ∇i rj ∇j + δij ∇j
= −i ri ∇2 − ∇i (1 + r · ∇) . (9.121)
ℓ,m Y
where we have identified Φℓ,m = rℓ+1 , the factor arising in a multipole
expansion. We recognize each term in the sum, namely Eℓ,m , to co-
incide precisely with the static electric multipole moments. For small
1
distances, these behave as ∼ rℓ+2 as opposed to the corresponding
1
terms in the magnetic fields (Bℓ,m ∼ rℓ+1 ). Therefore, the electric
fields dominate over the corresponding magnetic fields. In fact, in
the static limit, k = 0, the magnetic field vanishes and we are simply
left with the electric field which has the correct multipole expansion.
It is for these reasons that these solutions are known as the electric
multipole fields. (We have already seen this behavior in the case of
electric dipole radiation.)
Let us next analyze the behavior of the electric multipole fields in
the far away region (radiation zone) where kr ≫ 1. Asymptotically,
for large values of the argument, we know that the spherical Hankel
functions behave as
(1) eix
hℓ (x) → (−i)ℓ+1 ,
x
(2) (1) e−ix
hℓ (x) = (hℓ (x))∗ → (i)ℓ+1 . (9.127)
x
On the other hand, from our earlier discussion on fields produced by
an arbitrary distribution of charges, we know that in the radiation
zone the fields have the forms of outgoing spherical waves (see (9.22)
(2)
and (9.23)). Correspondingly, we conclude that the coefficient aℓ =
0 which leads to (for large values of kr)
X (−i)ℓ+1 ikr
e Yℓ,m
B(x) → aℓ L ,
k r
ℓ,m
X (−i)ℓ
eikr Yℓ,m
E(x) → aℓ ∇ × L (9.128)
k2 r
ℓ,m
X (−i)ℓ+1 ikr
2 ∂ e Yℓ,m
= aℓ r∇ − ∇ 1 + r ,
k2 ∂r r
ℓ,m
2eikr Yℓ,m
r∇
r
1 ∂ ikr
eikr 2
=r 2 (ikr − 1)e Yℓ,m − 3 L Yℓ,m
r ∂r r
ℓ(ℓ + 1)
= −r̂ k 2 + eikr Yℓ,m . (9.129)
r2
On the other hand, using (9.108) we have
ikr
∂ e Yℓ,m
∇ 1+r
∂r r
= ik∇ eikr Yℓ,m
keikr
= −r̂k 2 eikr Yℓ,m + r̂ × LYℓ,m . (9.130)
r
Substituting (9.129) and (9.130) into the expression for the electric
field in (9.128), we obtain, for large distances, (namely, we are ne-
glecting terms of order r12 compared to those of order 1r )
X (−i)ℓ+1 ikr
e Yℓ,m
E(x) → − aℓ r̂×L = −r̂×B(x). (9.131)
k r
ℓ,m
There are several things to note from the structure of the radia-
tion fields in Eqs. (9.128) and (9.131). We note that the fields in the
far off zone do fall off as 1r . Furthermore, using (9.98) it is clear that
the magnetic field is perpendicular to the direction of propagation r̂.
Similarly, the electric field is also perpendicular to the direction of
propagation and the electric and the magnetic fields are orthogonal
to each other. This is the general characteristic of radiation fields.
We can also obtain the small distance as well as the large distance
behaviors for the magnetic multipole fields in a similar manner. In
fact, looking at Eqs. (9.103) and (9.104), we see that the electric
multipole fields go over to the magnetic multipole fields under the
transformations (known as duality transformations)
E → B, B → −E. (9.132)
Given the electric and the magnetic fields for either of the mul-
tipole expansion, we note that we can write
X
B(x) = Bℓ,m (x),
ℓ,m
X
E(x) = Eℓ,m (x). (9.133)
ℓ,m
Here, Bℓ,m (x) and Eℓ,m (x) are the multipole fields of order (ℓ, m).
For the electric multipole fields, in the radiation zone, for example,
as we have seen
ikr
(−i)ℓ+1 aℓ e Yℓ,m
Bℓ,m (x) = L ,
k r
Eℓ,m (x) = −r̂ × Bℓ,m (x)
(−i)ℓ+1 aℓ eikr Yℓ,m
=− r̂ × L . (9.134)
k r
Using (9.134) we can now calculate the time averaged radiated power
for the multipole component fields through the surface of a large
sphere per unit solid angle along a given direction as
dPℓ,m cR2
= R2 r̂ · Sℓ,m = Re r̂ · (Eℓ,m × B∗ℓ,m )
dΩ 8π
cR2
= (B∗ℓ,m · Bℓ,m )
8π
c|aℓ |2
= (LYℓ,m )∗ · (LYℓ,m ). (9.135)
8πk 2
The right hand side can be simplified by noting that
(LYℓ,m )∗ · (LYℓ,m )
1
= |L+ Yℓ,m |2 + |L− Yℓ,m|2 + |Lz Yℓ,m|2
2
h1
= (ℓ − m)(ℓ + m + 1)|Yℓ,m+1 |2
2
i
+(ℓ + m)(ℓ − m + 1)|Yℓ,m−1 |2 + m2 |Yℓ,m |2 , (9.136)
Let us next work out explicitly the radiation pattern for a few
low order multipoles. For dipoles, ℓ = 1 and we obtain, from Eqs.
(9.137)–(9.139), that
dP1,0 c|a1 |2 3
= sin2 θ,
dΩ 8πk 2 4π
dP1,±1 c|a1 |2 3
= (1 + cos2 θ). (9.141)
dΩ 8πk 2 8π
We note that for the dipole case, the two distinct possible radiation
patterns can be represented as shown in Figs. 9.6 and 9.7. The first
describes the polar plot for the case m = 0 while the second denotes
the case for m = ±1. For m = 0, we see that the maximum power is
radiated along θ = π2 while for m = ±1, it is along θ = 0, as is also
clear from (9.141).
θ
z
θ
z
dP2,0 c|a2 |2 45
= sin2 θ cos2 θ,
dΩ 8πk 2 4π
dP2,±1 c|a2 |2 15
= 2
1 − 3 cos2 θ + 4 cos4 θ ,
dΩ 8πk 8π
dP2,±2 c|a2 |2 15
= 2
1 − cos4 θ . (9.142)
dΩ 8πk 8π
Similarly, the angular distributions of radiated power for higher mul-
tipole fields can also be calculated.
then, so is (r × ∇)ψ.
322 9 Radiation
(a) (a)
4. A time harmonic source “a” produces the fields Eω , Bω and
(b) (b)
another independent source “b” produces the fields Eω , Bω .
Show that, at any point outside the sources,
1 ∂H
A= , Φ = −∇ · H.
c ∂t
b) Show from this that, at large distances from the dipole, the
fields can be written as (up to possible multiplicative constants)
1 d2 −iω(t− r )
E= r × r × 2 pe c ,
4πcr 3 dt
1 d2 −iω(t− r )
B=− r × 2 pe c .
4πcr 2 dt
Chapter 10
323
324 10 Electromagnetic fields of currents
R(t′ ) = x − ξ(t′ ),
|x − ξ(t′ )| R(t′ )
τ = t′ − t + = t′ − t + , (10.5)
c c
where we have identified R(t′ ) = |R(t′ )|. Then, the integration over
t′ in (10.4) would have contribution only from that value of the time
coordinate for which the argument of the delta function vanishes,
namely,
R(t′ )
τ = t′ − t + = 0. (10.6)
c
This can, of course, be solved for t′ in principle, once we know the
trajectory of the particle. Furthermore, we note that
dτ 1 dR(t′ )
s(t′ ) = = 1 +
dt′ c dt′
1 ′ dR(t′ ) ′
b ′ ) · v(t ) ,
=1+ R(t ) · = 1 − R(t (10.7)
cR(t′ ) dt′ c
b = R . We note
where we have used the definition of R in (10.5) and R R
′
that s(t ) is a positive quantity whenever the speed of the particle
is less than the speed of light. (This does not hold for Čerenkov
radiation which we will discuss later.) We note the standard formula
for the delta function integral,
Z
g(x)
dx δ(f (x)) g(x) = , (10.8)
df
dx
x=x0
0 has only one solution. (We are also assuming that the speed of
the particle is less than the speed of light which is always true in
vacuum.) If there are more than one solution, we must sum over all
the contributions. Furthermore, the electric and the magnetic fields
associated with a moving charge can be calculated from the vector
potentials in the following manner.
We note that we can rewrite the scalar and the vector potentials
explicitly in a form convenient for our purpose as
Z R(t′ )
δ t′ − t + c
Φ(x, t) = q dt′ ,
R(t′ )
′
Z ′ δ t ′ − t + R(t )
v(t ) c
A(x, t) = q dt′ ′
. (10.10)
c R(t )
1 ∂A
E(x, t) = −∇Φ − . (10.11)
c ∂t
There are several things to note here. First of all, since Φ and A
depend on x only through their dependence on R, it can be easily
checked that acting on these functions, the effect of the gradient inside
the integral can be represented as ((∇R) = R) b
∂f (R) b ∂f (R) .
∇f (R) = (∇R) =R (10.12)
∂R ∂R
Z R
!
′b ∂ δ t′ − t + c
E(x, t) = −q dt R
∂R R
Z ′ R
q ∂ ′ v δ t −t+ c
− dt
c ∂t c R
Z " #
′ b δ t′ − t + Rc 1 ∂ ′ R
= −q dt R − − δ t −t+
R2 cR ∂t c
Z R
q ∂ v δ t′ − t +
′ c
− dt
c ∂t c R
326 10 Electromagnetic fields of currents
Z ′−t+ R
b δ
′ t c
= q dt R
R2
Z Rb − v δ t′ − t + R
q ∂ c c
+ dt′
c ∂t R
b b v
qR q ∂ R− c
= + . (10.13)
sR2 c ∂t sR
τ =0 τ =0
R(t′ )
t = t′ + = A(t′ )
c
dt dA(t′ ) 1 dR(t′ )
or, = = 1 + = s(t′ ), (10.14)
dt′ dt′ c dt′
t′ = B(t),
dt′ dB(t) 1
or, = = . (10.15)
dt dt s(t′ ) t′ =B(t)
∂ ′ ∂
f (t ) t′ =B(t) = f (B(t))
∂t ∂t
dB(t) ′
= f (B(t))
dt
df (t′ )
′
= dt ′ , (10.16)
s(t ) ′
t =B(t)
fied as follows.
b Rb−v
qR q ∂ c
E(x, t) = +
sR2 ′ c ∂t sR
t =B(t) t′ =B(t)
b Rḃ − v̇ Rb − v ṡ Rb − v Ṙ
R 1 c c c
= q 2 + 2
− 3
− 2 2 .
sR c s R s R s R
t′ =B(t)
(10.17)
Ṙ = −v,
Ṙ = −Rb · v,
ḃ d R(t′ ) 1 b b
R= ′ = R (R · v) − v ,
dt R(t′ ) R
ḃ · v − R
ṡ = −R b · v̇
c c
b ′) · v v
1 dR(t 1 b b b · v̇ .
=− = − R(R · v) − v · + R
c dt′ R c c
(10.18)
B(x, t) = ∇ × A(x, t)
Z
q ′ v ′ R
= dt ∇ × δ t − t +
c R c
328 10 Electromagnetic fields of currents
Z ′−t+ R
q b × v)
′ ∂ δ t c
= dt (R
c ∂R R
"Z
b ′ R
q ′ (R × v) δ t − t + c
=− dt
c R2
Z Rb × v δ t′ − t + R
∂ c
+ dt′
∂t cR
Rb × v b ×v
R
q ∂
=− + . (10.20)
c sR2 ∂t csR
′ t =B(t) t′ =B(t)
There are several things to note from the forms of the electric
and the magnetic fields in Eqs. (10.19) and (10.21) respectively. We
see that we need to solve for t′ = B(t) to determine the electromag-
netic fields. (t′ is known as the retarded time.) This can, in principle,
be done once we know the trajectory of the particle. We note from
(10.19) that, for a charged particle at rest, v = 0 = v̇, in which case
(s = 1 in such a case)
b
qR
E(rest) (x, t) = ,
R2
b × Erest (x, t) = 0,
B(rest) (x, t) = R (10.22)
R = x − ξ, (10.23)
will lead to power loss through radiation. However, we note that the
term corresponding to the radiation field ( R1 ) is proportional to v̇.
Correspondingly, we see that a particle moving in vacuum will radi-
ate power only if it is being accelerated. Conversely, an accelerated
charged particle will lose energy through radiation. On the other
hand, there is no radiation of power if the particle is not being accel-
erated (in vacuum). Let us note, from (10.19), that the amplitude of
the radiation term dominates if
R|v̇|
≫ 1. (10.24)
c2 ′
t =B(t)
We can solve for the trajectory of the particle easily in this case,
x
R
vt v
R(t′ )
t′ = t − ,
c
or, c(t′ − t) = −R(t′ ) = −|R − v(t′ − t)|,
2
or, (c2 − v 2 )(t′ − t)2 + 2Rv cos θ(t′ − t) − R = 0. (10.29)
Since R and θ are actually functions of time t, Eq. (10.31) gives the
necessary relation t′ = B(t).
Next, let us note from (10.28) and (10.31) that
R(t′ ) = R − v (t′ − t)
s
v 2
Rc
=R+ v cos θ + 1 − v sin θ , (10.32)
v2 c c
1− 2 c
b
b ′) − v = R R .
or, R(t (10.34)
c R(t′ )
Similarly, using (10.34), it is easily calculated that
2
v cos θ
b ′ ) · v = 1 − v 1 −
s(t′ ) = 1 − R(t q c
c c2 v cos θ
v sin θ 2
c + 1− c
q
v2 θ 2
1 − c2 1 − v sinc
= q . (10.35)
v cos θ v sin θ 2
c + 1 − c
Using (10.34) and (10.36) in the expression for the electric field
(10.25), we obtain
q Rb − v 1 − v22
c c
E(uniform) (x, t) =
3
s R 2
t′ =B(t)
b− v v2
q R c R 1− c 2
=
(sR)3
′
t =B(t)
b R (1 − v )
qR
2
c2
=
3 3
θ 2 2
R 1 − v sinc
qRb 1 − v2
c 2
= . (10.37)
2 3
θ 2 2
R 1 − v sinc
The magnetic field can also be calculated for this case using (10.21)
and (10.34) and it turns out to be
b ′ )
B(uniform) (x, t) = R(t × E(uniform) (x, t)
′ t =B(t)
v b 1 − v2
× qR
c c2
=
2 3
θ 2 2
R 1 − v sin
c
v
= × E(uniform) (x, t). (10.38)
c
This is, in fact, what we would expect from our earlier studies
in electrostatics. To see that, let us specialize to the case of particle
motion along the x-axis and the observation point x lying in the x−y
plane for simplicity, as shown in Fig. 10.2.
Introducing the conventional notations,
v
β= ,
c
1 1
γ=p =q , (10.39)
1− β2 1− v2
c2
we note that we can write the components of the electric and the
10.2 Uniform linear motion 333
y
x y
R
θ
vt x
Figure 10.2: A charged particle moving along the x-axis with the
observation point in the x − y plane.
Bx′ = Bx = 0,
x′ = γ(x − βct),
y ′ = y,
z ′ = z,
′ βx
t =γ t− . (10.43)
c
x′
(x − vt) = (x − βct) = ,
γ
1 r′
R(1 − β 2 sin2 θ) 2 = , (10.44)
γ
qx′
Ex′ = ,
r ′3
qy ′
Ey′ = ′3 ,
r
Ez′ = 0 = Bx′ = By′ = Bz′ . (10.45)
This shows that, in the rest frame of the charged particle, the mag-
netic field is zero and the electric field has the form that we will
expect for a point charge at rest. (There is no z-component of the
10.2 Uniform linear motion 335
electric field at the observation point since it lies in the x-y plane.)
For a charged particle moving with a uniform velocity in vacuum, we
do not expect any power loss due to radiation, since the radiation
components vanish in this case.
Let us now analyze Eqs. (10.37) and (10.38) in some detail.
First, we see that, for very small velocities, the electric and the mag-
netic fields have the forms following from Coulomb’s law and Biot-
Savart’s law respectively. However, as the velocity increases and ap-
proaches the speed of light, we see, from Eq. (10.37), that the electric
field becomes negligible along the direction of motion (along θ = 0).
Its magnitude is larger as θ increases from zero and peaks at θ = π2 ,
namely, at a direction perpendicular to the motion as shown in Fig.
10.3. (Namely, the electric field is dominantly transverse to the di-
rection of motion.)
E
v= 0
v^
Z∞
qγ (x − vt) eiωt
= dt 3
2π (γ 2 (x − vt)2 + y 2 ) 2
−∞
iωx Z∞ − iωyξ
qe v ξe γv
= dξ 3
2πγvy (1 + ξ 2 ) 2
−∞
iωx Z∞ − iωyξ
qe v iγv d e γv
= dξ 3
2πγvy ω dy (1 + ξ 2 ) 2
−∞
iωx
q e v πωy (1) iωy
= H
2πγvy γv 0 γv
iωx
qω e v (1) iωy
= H 0 , (10.46)
2γ 2 v 2 γv
γ(x−vt)
where we have used (10.41) and defined ξ = y in the interme-
(1)
diate steps. We note that Hn
represents the nth Hankel function
of the first kind and we have used some standard relations from the
table of integrals (see, for example, Gradshteyn and Ryzhik 8.407.1,
8.432.5, 8.472.1). Similarly, the Fourier transform of the y component
of the electric field leads to
Z∞
1
Ey (x, ω) = dt Ey (x, t) eiωt
2π
−∞
iωx Z∞ − iωyξ
qe v e γv
= dξ 3
2πvy (1 + ξ 2 ) 2
−∞
iωx
qe v πωy (1) iωy
= − H1
2πvy γv γv
iωx
qω e v (1) iωy
=− H1 . (10.47)
2γv 2 γv
10.3 Method of virtual photons 337
r
iωy 2γv − ωy π
Hn(1) → e γv −i(2n+1) 4 , (10.49)
γv iπωy
Ex (x, ω) → 0,
q iωx q iωx
Ey (x, ω) → e v → e c . (10.51)
πvy πcy
The non-vanishing magnetic field along the z-axis also has the same
form and magnitude as Ey . Thus, we see that every frequency com-
ponent of the fields behaves like a plane wave when v → c.
and write
γv
Ey (x, ω) = 0, ω> y ,
q iωx γv
(10.52)
Ey (x, ω) = πvy e v , ω< y .
This shows that the original field associated with the particle can be
thought of as an electromagnetic pulse.
It is, of course, much easier to solve for the fields for a given fre-
quency and once this is known, the original field can be reconstructed
by taking a linear superposition of the form in (10.53). An electro-
magnetic field configuration corresponding to a given frequency can,
of course, be associated with a photon of energy ~ω. The electromag-
netic fields associated with a moving charged particle can, therefore,
be thought of as resulting from emission and reabsorption of “vir-
tual” photons by the charged particle. The word “virtual” (also in
the method of virtual photons) comes from the following quantum
mechanical correspondence. Quantum mechanics allows a charged
particle to emit and reabsorb photons while it travels as shown in
Fig. 10.4.
k, ω
p, E p′ , E ′
Here, we have used the fact that the particle is moving along the
x-axis with a uniform velocity v, to convert the x integral to a time
integral. Furthermore, from the definitions of the Fourier transform
as well as the delta function,
Z∞
E(x, t) = dω E(x, ω) e−iωt ,
−∞
Z∞ Z∞
1 1
δ(ω) = dt e iωt
= dt e−iωt , (10.59)
2π 2π
−∞ −∞
we obtain
Z Z
v ′
U= dy y dt dω dω ′ Ey (x, ω)Ey∗ (x, ω ′ ) e−i(ω−ω )t
2
Z Z
= πv dy y dω dω ′ Ey (x, ω)Ey∗ (x, ω ′ ) δ(ω − ω ′ )
Z Z∞
= πv dy y dω |Ey (x, ω)|2
−∞
Z Z∞
= 2πv dy y dω |Ey (x, ω)|2
0
γv
Z∞ Zω
q2
= 2πv dω dy
π2v2 y
0 b
γv
Z∞ Zω
2q 2 dy
= dω . (10.60)
πv y
0 b
2q 2 γv
~ω N (ω) = U (ω) = ln ,
πv ωb
2q 2 γv
or, N (ω) = ln . (10.63)
πv~ω ωb
Let us assume that the charged particle is an electron with
charge q = −e. Then, we note that the fine structure constant has
the form
e2
α= . (10.64)
~c
Furthermore, identifying the minimum impact parameter from the
uncertainty relation as
~ ~
b∼ = , (10.65)
|∆p| |p − p′ |
2α γβc|p − p′ |
N (ω) ∼ ln
πβω (E − E ′ )
2α~ E βc|p − p′ |
= ln , (10.66)
πβ(E − E ′ ) mc2 (E − E ′ )
where we have used (10.56). This gives the flux of transverse “vir-
tual” photons. Note that, since the fine structure constant is a small
number, the number of photons associated with an electron is also
small.
R(t′ ) ≈ r ≫ d. (10.67)
Under this approximation, we see that only the second term in (10.19)
would dominate and the electric field would have the form (we are
now preparing to consider nonuniform motion)
b × ((R
qR b − β) × β̇)
E(x, t) = . (10.68)
cs3 R ′
t =B(t)
|x − ξ(t′ )| r
t′ = t − ≈t− ,
c c
r
or, (t′ − t) ≈ − , (10.69)
c
R(t′ ) = x − ξ(t′ ) ≈ x,
R(t′ ) ≈ r = |x|,
′
b ′ ) = R(t ) ≈ x̂,
R(t ′
R(t )
r¯
β(t′ ) ≈ β(t) + (t′ − t)β̇(t) ≈ β̄ − β̇ ≈ β̄,
c
b ′ ) · β(t′ )
s(t′ ) = 1 − R(t
r ¯
≈ 1 − x̂ · β̄ + x̂ · β̇
c
≈ 1 − x̂ · β̄, (10.70)
R(t′ ) ≈ R ≈ x,
R(t′ ) ≈ R ≈ r,
b ≈ x̂,
b ′) ≈ R
R(t
|β(t′ )| ≈ |β̄| ≪ 1,
s(t′ ) ≈ 1. (10.71)
344 10 Electromagnetic fields of currents
¯
q x̂ × (x̂ × β̇)
≈
cr
¯ ¯
q (x̂(x̂ · β̇) − β̇)
=
cr
qa sin θ
= θ̂, (10.72)
c2 r
¯
where we have defined a = c|β̇| to be the magnitude of the acceler-
ation. Here, θ is the angle between x̂ and v̇ (see Fig. 10.5) and θ̂
represents the unit vector along this direction (which is not fixed) so
¯ ¯ ¯
that β̇ = x̂(x̂ · β̇) − θ̂|β̇| sin θ. It follows now that
b ′ )
B(dipole) (x, t) = R(t × E(dipole) (x, t)
′
t =B(t)
qa sin θ
≈ (x̂ × θ̂). (10.73)
c2 r
θ̂
x
θ
v̇
therefore, the factor differs from the earlier formula by two), in this
case, has the form
c
S= Re E(dipole) × H(dipole) ∗
4π
q 2 a2 sin2 θ
= θ̂ × (x̂ × θ̂)
4πc3 r 2
q 2 a2 sin2 θ
= x̂. (10.74)
4πc3 r 2
Here, we have used the identities involving vector products as well as
the fact that x̂ and θ̂ are orthogonal. We have also assumed that we
are in vacuum where H = B. Thus, we see that the radiated power
along x̂, through a sphere of large radius r, is given by
dP q 2 a2
= r 2 x̂ · S = sin2 θ. (10.75)
dΩ 4πc3
This is exactly the angular distribution of power radiated from a
dipole (see (9.43) as well as Fig. 10.6) and for this reason, this ap-
proximation is conventionally called the dipole approximation.
θ
z
dP |p̈|2
= sin2 θ. (10.77)
dΩ 4πc3
346 10 Electromagnetic fields of currents
10.4.2 Linear acceleration. Let us next consider the case where the
particle is extremely relativistic and is subjected to an acceleration
that is small during the time scale that observations are made. We
also assume that the particle is moving in a linear trajectory so that
the acceleration is parallel to the velocity. Therefore, we have
β̇ k β, |β̇| ≪ 1. (10.79)
In this case, therefore, from (10.70) we have
R(t′ ) ≈ r,
b ′ ) ≈ x̂,
R(t
β(t′ ) ≈ β̄,
s(t′ ) ≈ 1 − x̂ · β̄. (10.80)
Furthermore, since the acceleration is parallel to the velocity, we have
β × β̇ = 0. (10.81)
With these, we see that at large distances, the electric field takes
the form (see (10.68))
q R b − β) × β̇)
b × ((R
E(linear) (x, t) =
cs3 R ′
t =B(t)
¯
q x̂ × (x̂ × β̇)
≈
cr(1 − x̂ · β̄)3
qa sin θ
= θ̂, (10.82)
c2 r(1 − β̄ cos θ)3
10.4 Asymptotic values of the fields 347
qa sin θ
≈ (x̂ × θ̂). (10.83)
c2 r(1 − β̄ cos θ)3
dP cr 2
= r 2 x̂ · S = Re x̂ · E(linear) × H(linear) ∗
dΩ 4π
q 2 a2
≈ sin2 θ. (10.84)
4πc3 (1 − β̄ cos θ)6
Thus, we see that the radiated power has an angular distribution very
much like the dipole approximation except that it is modulated by
1
the relativistic correction 6 as shown in Fig. 10.7.
(1−β̄ cos θ)
which gives
Zπ
q 2 a2 sin3 θ
Ptotal = 2π dθ
4πc3 (1 − β̄ cos θ)6
0
Z1
q 2 a2 1 − x2
= 2π dx
4πc3 (1 − β̄x)6
−1
q 2 a2 8π(5 + β̄ 2 ) 2q 2 a2 (5 + β̄ 2 )
= = . (10.85)
4πc3 15(1 − β̄ 2 )4 3c3 5(1 − β̄ 2 )4
where θ is the angle x makes with the z-axis (see Fig. 10.8, it is
different from the θ in the earlier examples) and ê{x,y,z} represent
the Cartesian unit vectors along the three axes.
xθ
ξ y
x
ξ̇ ωa
β= = (−êx sin ωt + êy cos ωt)
c c
= β (−êx sin ωt + êy cos ωt) ,
ω2a
β̇ = − (êx cos ωt + êy sin ωt)
c
cβ 2
=− (êx cos ωt + êy sin ωt)
a
= −β̇ (êx cos ωt + êy sin ωt) . (10.87)
ωa cβ 2
β = |β| = , β̇ = |β̇| = . (10.88)
c a
β · β̇ = 0. (10.89)
R(t′ ) ≈ r,
b ′ ) ≈ x̂,
R(t
β(t′ ) = β,
s(t′ ) ≈ 1 − x̂ · β = 1 + β sin θ sin ωt. (10.90)
It is now easy to calculate the electric field in (10.19) for large dis-
350 10 Electromagnetic fields of currents
dP
= r 2 x̂ · S
dΩ
cq 2 β 4 (1 − β 2 ) cos2 θ + (β + sin θ sin ωt)2
= . (10.94)
4πa2 (1 + β sin θ sin ωt)6
This is, of course, time dependent and since the motion is harmonic,
we can average over one cycle of the motion. In doing so, however,
we have to be careful and note from (10.14) that
Z2π
cq 2 β 4 (1 − β 2 ) cos2 θ + (β + sin θ sin φ)2
= 2 2 dφ
8π a (1 + β sin θ sin φ)5
0
Z2π
cq 2 β 2 (1 − β 2 )(1 − β 2 sin2 θ)
= 2 2 dφ
8π a (1 + β sin θ sin φ)5
0
2(1 − β 2 ) 1
− + .
(1 + β sin θ sin φ)4 (1 + β sin θ sin φ)3
(10.96)
Z2π
1
In = dφ , (10.97)
(1 + β sin θ sin φ)n
0
z = eiφ . (10.98)
where the integration is over a unit circle. The integrand has n-th
order poles at
q
i 2 2
z= −1 ± 1 − β sin θ . (10.100)
β sin θ
We note that only the first root lies within the unit circle and, con-
sequently, the integral can be evaluated in a straightforward (but
tedious manner since it is a higher order pole) using the method of
352 10 Electromagnetic fields of currents
#
35
+ 9 . (10.101)
(1 − β 2 sin2 θ) 2
where we have used (10.88). This shows that the power radiated
peaks along the z-axis (θ = 0). In contrast, in the relativistic case we
see from (10.102) that the radiated power peaks at θ = π2 . The study
of this system is particularly useful in the analysis of synchrotron
radiation in accelerators.
of the traveling wave (this is also true of the speed of light in the
medium). A charged particle traveling in a medium can, therefore,
travel faster than the speed of light in the medium without violating
causality provided c′ < v < c. In such a case, radiation is produced
even when the particle is not being accelerated and this effect, known
as the Čerenkov effect, provides an important tool in detecting high
energy particles.
The Čerenkov effect is an interesting phenomenon where the
charged particle does not lead to radiation directly. Rather, it is a col-
lective phenomenon where the radiation is produced by the medium
through which the charged particle moves. Quantum mechanically,
the phenomenon can be understood as follows. The charged particle
moving through the medium excites the electrons in the atoms which,
upon return to their original state, emit a coherent radiation. Macro-
scopically, the phenomenon is analogous to the production of sound
waves (shock waves) in the case of supersonic motion in a medium,
where the fluctuations in the density of the medium produce a sound
wave. In fact, geometrically, it is easy to see that if the particle trav-
els with v < c′ , then the information spheres (spherical wave fronts
traveling with the speed of light in the medium), originating at later
times, are contained inside the earlier ones. In this case, it is easy to
see geometrically that there is only one unique retarded time for every
point as shown in Fig. 10.9. (Basically, what this means is that any
observation time can lie only on the surface of a single information
sphere and, therefore, would correspond to a unique retarded time.)
123
b b b
c′
cos θc = . (10.104)
v
c ′t θc
b b b
1 2 3 vt
which also implies β ′ > 1. This is the condition for the emission of
Čerenkov radiation. Geometrically, one can see that, in this case,
there can be more than one retarded times associated with any given
point.
To understand the Čerenkov radiation more quantitatively, let
us note that, in an arbitrary medium, the vector potential, in the
10.5 Čerenkov effect 355
R h p i
c′ (t′ − t) = − β ′
cos θ ∓ (β ′ cos θ)2 + (1 − β ′2 ) .
(1 − β ′2 )
(10.111)
Here θ is the angle between R = R(t) and v (see Fig. 10.2). For t′ to
represent a retarded time, as we have argued before, only the second
root is allowed when β ′ < 1. As we will see now, for β ′ > 1, both the
roots are allowed leading to two retarded times for a given t.
For β ′ > 1, we note from (10.111) that real roots will exist only
if
(β ′ cos θ)2 + (1 − β ′2 ) > 0
2 1
or, cos θ > 1 − ′2 , (10.112)
β
which leads to the two solutions
1 1
1 2 1 2
cos θ > 1 − ′2 , or, cos θ < − 1 − ′2 . (10.113)
β β
On the other hand, for t′ to represent a retarded time, we see from
(10.111) that cos θ must be negative (because of the factor in the
denominator). Therefore, we choose θ to be an obtuse angle and
note from (10.113) that the allowed root corresponds to
1 !
1 2
arccos − 1 − ′2 < θ < π. (10.114)
β
In this case, it is easy to see from (10.111) that both the solutions
lead to retarded times and provide the two retarded times for this
problem.
Even though the retarded times are determined (and now the
contributions from both these solutions must be added in (10.110)),
evaluating the fields is extremely complicated. In what follows, we
will present an alternate derivation of the fields as well as the energy
radiated, using the method of Fourier transforms, which also brings
out some other interesting features. To keep the discussion parallel
to the earlier discussion of uniform linear motion in vacuum, let us
choose the charged particle to be traveling along the x-axis. In this
case, therefore, the current has a non-vanishing component only along
the x-axis given by
Jx (x, t) = qv δ(x − vt)δ(y)δ(z). (10.115)
10.5 Čerenkov effect 357
B(x, ω) = ∇ × A(x, ω)
Z
4πi k × J(k, ω) ik·x
= ′ d3 k 2 e . (10.120)
c k2 − cω′2
Putting in the form of the current in (10.116), we see that the mag-
netic field is orthogonal to the direction of motion of the charged
particle (namely, Bx = 0) and has the form
358 10 Electromagnetic fields of currents
iωx Z
iq e v (kz êy − ky êz )
B(x, ω) = dky dkz ei(ky y+kz z) .
2π 2 c′ ky2 2 ω2
+ kz − c′2 1 − β ′2 1
(10.121)
If β ′ < 1 (v < c′ ), we note that the quantity inside the square root is
negative and, therefore, the poles are on the imaginary axis. Enclos-
ing the contour in the upper half plane would pick up the contribution
of the pole on the positive imaginary axis and it is easy to check that
the integrals, in this case, lead to fields of the forms discussed in
(10.47) and (10.48).
On the other hand, if β ′ > 1 (v > c′ ), then, the quantity inside
the square root in (10.122) can be positive for some values of ky ,
while it will be negative for other values. When ky is such that the
quantity is negative, the poles will lie on the imaginary axis and
the integration can be carried out, much like in the earlier case, by
enclosing the contour in the upper half plane. For values of ky for
which the quantity inside the square root is positive, the poles will
lie on the real axis. However, since we are interested in retarded
solutions, as discussed in (6.179), the proper prescription is obtained
by letting (this gives the pole prescription for the retarded Green’s
function or solution)
ω → ω + iǫ,
whose effect, in the present case, is to push the pole at the positive
value of kz to the upper half plane while pushing the other pole at the
negative value to the lower half plane. Thus, enclosing the contour
in the upper half plane would pick up two contributions – one from
the pole on the positive imaginary axis and the other from the pole
slightly above the positive real axis. This is the basic difference from
the case β ′ < 1 and is a reflection of the existence of two retarded
times in this language. Denoting by k∗ the generic location of the
10.5 Čerenkov effect 359
a medium with a speed larger than the speed of light in the medium.
To see that there is radiation, in such a case, let us calculate the
energy radiated through the conical surface along the ẑ direction as
the particle travels a distance d along the trajectory. Let us consider
a surface parallel to the x − y plane defined by z = z0 . Further-
more, let us consider a strip of the surface of width d along the x-axis
(x0 ≤ x ≤ x0 + d). Then, the total energy radiated through this
strip (considering contributions from both the surfaces at ±z0 , as we
will see shortly, the contribution is independent of the value of z0 ) is
given by
Z∞ Z∞ xZ0 +d
c′
E = 2 Re dt dy dx ẑ · (E(x, t) × H∗ (x, t))
4π
−∞ −∞ x0
Z∞ Z∞ xZ0 +d
c′
= Re dt dy dx Ex (x, t)By∗ (x, t)
2π
−∞ −∞ x0
Z∞ Z∞ xZ0 +d
c′
= 4π Re dω dy dx Ex (x, ω)By∗ (x, ω)
2π
0 −∞ x0
Z∞ Z∞ xZ0 +d
Z∞ Z∞ Z∞ Z∞
q 2
= 2d − Re dω dy dky dky′
2π
0 −∞ −∞ −∞
′ ′
ω c′2 ei(ky −ky )y+i(k∗ −k∗ )z0
× ′2 1− 2
c v k∗
Z∞ Z∞ Z∞
q2d ω c′2
= Re dω dky dky′ 1− 2
π c′2 v
0 −∞ −∞
′
δ ky − ky ei(k∗ −k∗ )z0
′
×
k∗
Z∞ Z∞
q2d ω c′2 1
= Re dω ′2 1− 2 dky . (10.129)
π c v k∗
0 −∞
where we have used standard results from the table of integrals (see,
for example, Gradshteyn and Ryzhik, 2.274 or this can also be done
by elementary methods). With this, we obtain
Z∞
2 ω c′2
E = q d dω ′2 1 − 2
c v
0
Z∞
2 1 1
=q d dω ω ′2
− 2 . (10.132)
c v
0
We can also define the energy radiated per unit length as
Z∞
E 2 1 1
W = =q dω ω ′2 − 2 . (10.133)
d c v
0
362 10 Electromagnetic fields of currents
This can be thought of as the energy radiated per unit length of the
trajectory of the charged particle. Thus, we see that, even though the
charged particle is not being accelerated, by virtue of the fact that
it travels faster than the speed of light in the medium, it leads to
emission of radiation. This effect is known as the Čerenkov radiation
(effect). (Equation (10.133) shows that if v > c′ , energy is radiated
since E > 0. On the other hand, if v < c′ , no energy is radiated since
E < 0.)
10.6 Self-force
(t′ − t)2 ¯
= R + (t′ − t)(−v̄) + (−v̇)
2
(t′ − t)3 ¯ + ··· ,
+ (−v̈) (10.134)
6
where R(t′ ) = x−ξ(t′ ). Let us recall that the retarded time is defined
as
R(t′ )
t′ = t − . (10.135)
c
Using this as well as the fact that the particle is instantaneously at
rest (v̄ = 0), we obtain, from (10.134),
R2 (t′ ) ¯ R3 (t′ ) ¯
R(t′ ) = R − v̇ + v̈ + · · · . (10.136)
2c2 6c3
We are assuming here as before that Rc (and, therefore, t′ − t) is small
and, consequently, a Taylor expansion as in (10.136) is meaningful.
Furthermore, we keep expansions up to order R3 because, as we will
see shortly, that is sufficient for our purposes.
Equation (10.136) leads to
2 R2 (t′ ) 3 ′
¯ + R (t ) R · v̈
4 ′
¯ + R (t ) v̇
¯ · v̇
¯ + ··· ,
R2 (t′ ) = R − R · v̇
c2 3c3 4c4
(10.137)
which can be rearranged to the form
¯ ¯ ¯ · v̇
¯
2 ′ R · v̇ 2 R3 (t′ )R · v̈ R4 (t′ )v̇
R (t ) 1 + 2 ≈R + 3
+ 4
,
c 3c 4c
¯ − 2 2 R3 (t′ )R · v̈
1
¯ ¯ 2
¯ · v̇
1
′ R · v̇ R4 (t′ )v̇
or, R(t ) ≈ 1+ 2 R + + .
c 3c3 4c4
(10.138)
364 10 Electromagnetic fields of currents
3
Iterating (10.138) and keeping terms to order R , we obtain
2 ¯ ¯
!
¯
R · v̇ ¯ 2 R(R · v̈)
3(R · v̇) ¯ R (v̇ · v̇)
′
R(t ) ≈ R 1 − 2
+ 4
+ 3
+ 4
.
2c 8c 6c 8c
(10.139)
Using this, we can rewrite (10.136) to leading orders as
2¯ 2 ¯ v̇
¯ 3¯
′ R v̇ R (R · v̇) R v̈
R(t ) = R − 2 + + 3 + ··· . (10.140)
2c 2c4 6c
In a similar manner, we can also Taylor expand (note that v̄ = 0)
′ 2
¯ + (t − t) v̈
v(t′ ) = v̄ + (t′ − t) v̇ ¯ + ···
2!
¯ ¯ v̇
¯ 2¯
Rv̇ R(R · v̇) R v̈
≈− + +
c 2c3 2c2
¯ v̇¯ ¯
R ¯ (R · v̇) Rv̈
= −v̇ + + ,
c 2c2 2c
¯ + (t′ − t)v̈
v̇(t′ ) ≈ v̇ ¯ = v̇¯ − R v̈.¯ (10.141)
c
As we have seen in (10.19), the electric field of a point charge
can be written as
" v̇ #
R − Rvc 1 − ( v 2
c ) R × R − Rv
c × c
E(x, t) = q 3
+ 3
.
(sR) c(sR) ′t =B(t)
(10.142)
Thus, using (10.139)-(10.141), we can expand
R(t′ ) · v(t′ )
s(t′ )R(t′ ) = R(t′ ) −
c
" 2 ¯ ¯
#
¯
R · v̇ ¯ 2 R(R · v̈)
(R · v̇) ¯ 3R (v̇ · v̇)
=R 1+ − − − + ··· ,
2c2 8c4 3c3 8c4
(10.143)
and, similarly,
2¯ 3¯ 2 ¯ v̇¯
R(t′ )v(t′ ) R v̇ R v̈ R (R · v̇)
R(t′ ) − =R+ 2 − 3 − 4
+ ··· ,
c 2c 3c 2c
2 ¯ ¯
v(t′ ) 2 R (v̇ · v̇)
1−( ) =1− 4
+ ··· . (10.144)
c c
10.6 Self-force 365
3
The reason for keeping terms up to order R in the expansion is now
clear. Since sR ∼ R and (sR)3 occurs in the denominator (in both
the terms), the dominant terms would come from terms up to order
3
R in the numerator.
Using (10.143) and (10.144), we now obtain
" 2 ¯ ¯
R− Rv
1 − ( vc )2 1 3(R · v̇) ¯ R (v̇ · v̇)
c
= 3 R 1− +
(sR)3 R 2c 2 8c 4
¯ 2 R(R · v̈)
15(R · v̇) ¯
+ +
8c4 c3
2 2
! 3
#
R 5R (R · ¯
v̇) R
¯
+v̇ − − 3 v̈ ¯ + ··· ,
2c2 4c4 3c
v̇
R× R − Rv × c 1 ¯
c
= R (R · ¯ − R(R · v̈)
v̇)
c(sR)3 cR3 c
2 ¯ ¯
!
R (v̇ · v̇) 3(R · v̇) ¯ 2
− −
2c2 2c2
2
! 3
#
¯
¯ −R + 2R (R · v̇)
+v̇
2
+
R ¯
v̈ + · · · .
c2 c
(10.145)
Z
Rv
v 2
R− c 1− c
E(x, t) = d3 ξ ρ(ξ)
(sR)3
v̇ #
R× R − Rv
c × c
+ 3
. (10.146)
c(sR) ′ t =B(t)
Z
Rv
v 2
R− c 1− c
= d3 x d3 ξ ρ(x)ρ(ξ)
(sR)3
v̇ #
R× R − Rv
c × c
+ 3 . (10.147)
c(sR)
We can now use the expansions in (10.145) and note that, for spher-
ically symmetric charge distributions, the terms in the integrand
which are odd in R would vanish (recall that R = x − ξ). As a
result, we obtain
Z ¯ ¯ ¯
3 3 (R · v̇)R v̇ 2v̈
Fself = d x d ξ ρ(x) ρ(ξ) − 3 − +
2c2 R 2c2 R 3c3
¯ Z
2v̇ ρ(x) ρ(ξ) 2q 2 ¯
= − 2 d3 x d3 ξ + 3 v̈
3c R 3c
4U ¯ 2q 2 ¯
=− v̇ + 3 v̈. (10.148)
3c2 3c
Here, we have used symmetric integration in the intermediate steps
and have identified the self-energy of the system as
Z Z
1 3 3 ρ(x) ρ(ξ) 1 ρ(x) ρ(ξ)
U= d xd ξ = d3 x d3 ξ . (10.149)
2 R 2 |x − ξ|
¯ = v̇, v̈
where we have identified, for simplicity, v̇ ¯ = v̈ and have
defined
4U
mem = , m = mI + mem . (10.151)
3c2
We can think of mI as the inertial mass of the electron while mem can
be thought of as the electromagnetic mass of the electron (charged
particle). Neither is individually observable. Experimentally, one can
measure only the observable mass m.
In deriving Eq. (10.148), we have neglected higher order terms
which would vanish in the limit that the particle has no structure.
However, in that limit, the self-energy of the electron (particle) in
(10.149) and, therefore, the electromagnetic mass diverges. In a quan-
tum theory, such a phenomenon is handled through renormalization.
Classically we can think of the electron not as a point particle, rather
as one with a structure of the size of about (experiments put an upper
bound of 10−17 cm on the size of the electron)
e2
re = ≈ 3 × 10−13 cm. (10.152)
mc2
This is also known as the Lorentz radius or the Thomson scattering
length. Consequently, the self-energy is finite. Furthermore, this
distance scale also defines a time scale
2re 2e2
τ= = ≈ 10−24 sec, (10.153)
3c 3mc3
where we have put in the factor of 23 in the definition of the time scale
for later convenience (see also the second term on the right hand side
of (10.150)). This time scale is tiny showing that the expansion used
is convergent. Moreover, since such a time scale is in the domain
of quantum mechanics, we recognize that we can, at best, think of
the classical equation in (10.150) as an approximate equation. If not,
Eq. (10.150) leads to conceptual problems. For, suppose there is no
external force present, namely, Fext = 0, then we see that
v̇ = τ v̈
t
or, v(t) = v(0) e τ + constant. (10.154)
forces and that the radiation loss due to the self-force is small com-
pared with the energy of the particle. We note that these assumptions
are quite reasonable. First of all, the self-force as we see from (10.148)
is proportional to the acceleration (as well as higher derivative terms)
and, consequently, cannot exist in the absence of an external force.
We also note that a charged particle of finite size cannot be in stable
equilibrium unless external forces are applied. Therefore, the first
assumption is quite reasonable. Since τ is very small, the energy loss
due to radiation in any finite amount of time can only be small. To
see this quantitatively, let us consider an one dimensional charged
oscillator in the presence of the self-force.
mẍ + mω02 x = mτ v̈,
or, ẍ + ω02 x = τ v̈, (10.155)
where ω0 is the natural frequency of the oscillator. This is a third
order equation in the time derivatives (recall that v = ẋ). Choosing
a solution of the form
x(t) = x(0) e−iωt , (10.156)
we obtain, from (10.155)
ω 3 τ − iω 2 + iω02 = 0. (10.157)
This can also be rewritten as
(ωτ )3 − i(ωτ )2 + i(ω0 τ )2 = 0. (10.158)
We note that the natural dimensionless variables in this equation are
ωτ and ω0 τ . Assuming that ω0 τ ≪ 1, we can obtain the solution to
the cubic equation in (10.158) perturbatively as
i
ωτ = ±ω0 τ − (ω0 τ )2 + O (ω0 τ )3 ,
2
ωτ = i + i(ω0 τ )2 + O (ω0 τ )3 . (10.159)
We can discard the last solution as unphysical since it leads to an
exponentially growing solution. The other two solutions, on the other
hand, lead to
1 2
x(t) = x(0) e∓iω0 t e− 2 ω0 τ t , t > 0. (10.160)
Both these solutions are exponentially damped. We recall that the
radiation component of the electric field is proportional to the accel-
eration (ẍ) which is proportional to ω02 and, consequently we expect
10.7 Selected problems 369
Z∞
1
x(ω) = dt eiωt x(t)
2π
0
Z∞
x(0) i
dt ei(ω∓ω0 + 2 ω0 τ )t
2
=
2π
0
ix(0)
= . (10.161)
2π (ω ∓ ω0 ) + 2i ω02 τ
This shows that the oscillation of the charged particle no longer con-
sists of a sharp single frequency. The intensity of the oscillations is
obtained to be
|x(0)|2
|x(ω)|2 =
ω04 τ 2
. (10.162)
4π 2 (ω ∓ ω0 )2 + 4
z = a cos ωt,
dP q 2 cβ 4 (4 + β 2 cos2 θ) sin2 θ
h i= 7 .
dΩ 32π 2 a2 (1 − β 2 cos2 θ) 2
370 10 Electromagnetic fields of currents
2q 2 a2 1
P = .
3c (1 − β 2 )3
3
(What this means is that you should correct for the retarded
time and define
dP dP dt′
→ ,
dΩ dΩ dt
and integrate this to obtain the total “time corrected” power
radiated.)
4. For a real function, f (t), show from the definition of the Fourier
transformation
Z ∞
1
f (ω) = dt eiωt f (t),
2π −∞
that
f ∗ (−ω) = f (ω).
Using this, derive the relation used in this chapter that
Z ∞ Z ∞
2
dt |f (t)| = 4π dω |f (ω)|2 .
−∞ 0
dpµ µ µ
= fext + fself ,
dτ
where pµ = muµ with uµ representing the four velocity of the
particle. m, τ denote respectively the rest mass and the proper
µ µ
time of the electron. Similarly, fext , fself denote respectively the
relativistic generalization of the external force and the self-force.
10.7 Selected problems 371
Plasma
373
374 11 Plasma
+±±±±±±±±−
+±±±±±±±±−
+±±±±±±±±−
+±±±±±±±±−
+±±±±±±±±−
+±±±±±±±±− x
Figure 11.1: The two charged surfaces which develop when the elec-
trons are displaced to the right along the x-axis.
The motion of the electrons (within the two charged surfaces) along
the x-axis will now be subjected to a force leading to (e > 0)
4πN e2
ωp2 = . (11.3)
m
This is known as the plasma frequency and this analysis shows that,
because of this displacement (disturbance), the plasma will begin to
oscillate as (this is the complex notation and the coordinate of the
particle will be given by the real part)
Thus, we see that, in this simple case, the plasma will oscillate only
with the plasma frequency. Although this is not exactly the “plasma
oscillation” that one talks about in connection with a plasma (which
11.1 General features of a plasma 375
we will discuss next), this simple example illustrates how the plasma
tries to maintain its neutrality when disturbed slightly.
If the plasma is subject to a driving force with a given angu-
lar frequency, then it can oscillate with a frequency different from
the plasma frequency (11.3). For example, suppose we have a har-
monic electric field of the form (we are suppressing the coordinate
dependence for simplicity)
ie2 iωp2
j = −ev = E= E, (11.9)
mω 4πN ω
where we have used the definition of the plasma frequency in (11.3).
Relation (11.9) is very interesting in that it is reminiscent of the
Ohm’s law,
iN e2 iωp2
J = N j = −N ev = E= E = σ E, (11.10)
mω 4πω
except for the factor of “i” in the proportionality constant. This fac-
tor simply implies that the electric field and the current are not in
phase. In this sense, a plasma is somewhat like a dielectric medium
and, as we will see shortly, this phase difference has important con-
sequences.
376 11 Plasma
∇2 Φ = −4πρ,
1 d 2 d
or, r Φ = −4πe(Ni − Ne ),
r 2 dr dr
d2 Φ 2 dΦ eΦ
or, 2
+ = 8πN e sinh . (11.12)
dr r dr kT
11.2 Plasma oscillation 377
d2 Φ 2 dΦ 8πN e2
+ − Φ = 0. (11.13)
dr 2 r dr kT
We recognize this as the spherical Bessel equation of order zero and
the solution that vanishes asymptotically has the form
C − rr
Φ(r) = e D, (11.14)
r
where C is a constant and we have defined the Debye length as
r s
kT kT
rD = = . (11.15)
8πN e2 2mωp2
C(r + rD ) − rr
E = −∇ Φ = r̂ e D. (11.16)
r 2 rD
Thus, we see that the scalar potential as well as the electric field fall
off rapidly for r > rD . Namely, the charged particles of the plasma
will reorganize themselves so as to screen the external charge beyond
the Debye length. For this reason, the Debye length is also sometimes
referred to as the screening length. These examples illustrate that
when a plasma is disturbed, it tries to restore its charge neutrality.
where the fluctuation from the equilibrium value, n(x, t), is consid-
ered to be small. As we have already seen, there will be oscillations
in the plasma leading to a current density (see (11.9))
iNeq e2 iωp2
J = Ne (x, t) j ≈ E= E, (11.18)
mω 4πω
where we are neglecting terms quadratic in the fluctuations, an ap-
proximation which is also known as the linearized approximation
(namely, since the velocity v is already a fluctuation, n(x, t)v is
quadratic in the fluctuations).
The first of the Maxwell’s equations, in this case, takes the form
∇ · B = 0,
iω
∇×E= B,
c
!
4π iω iω ωp2
∇×B= J− E=− 1− 2 E, (11.20)
c c c ω
ωp2
ǫp = 1 − . (11.21)
ω2
It is worth noting that ǫp ≤ 1 as opposed to the case of a dielectric
for which ǫ ≥ 1.
There are now several interesting cases to be discussed. If the
frequency of the harmonic field coincides exactly with that of the
plasma, namely, if ω = ωp , then we have
ǫp = 0. (11.22)
In this case, the second and the fourth of Maxwell’s equations give
∇ · B = 0,
∇ × B = 0. (11.23)
11.2 Plasma oscillation 379
Namely, the electron current, in this case, exactly cancels the dis-
placement current so that the curl of the magnetic field vanishes. In
the absence of sources, these two equations imply that the magnetic
field vanishes identically, namely,
B = 0. (11.24)
∇ · E = 4πρ(x, t),
∇ × E = 0. (11.25)
∂2ρ
+ ωp2 ρ = 0. (11.27)
∂t2
Namely, the density of electrons fluctuates in time with the plasma
frequency. This is a cooperative phenomenon in the sense that,
not one electron, but the plasma of electrons as a whole oscillates.
However, there is no traveling disturbance that is generated. This
should be contrasted with the transverse traveling wave solutions of
Maxwell’s equations for which
∇ · E = 0,
This is, in fact, the basic principle used in transmitting low frequency
radio waves. Let us recall that the ionosphere consists of a plasma
of electrons and positive ions where the ionization is a consequence
of radiations coming from the sun. As a result, low frequency radio
waves cannot penetrate the ionosphere and are reflected back, leading
to a transmission of such waves around the globe. The theory of
propagation of electromagnetic waves in the ionosphere is, however,
slightly more involved owing to the fact that the density of electrons
in the ionosphere is not uniform. Rather, it changes with the height
from the surface of the earth.
me v̇e = −eE,
ie
ve = − E. (11.36)
me ω
mi v̇i = eE,
ie
vi = E. (11.37)
mi ω
We have used the subscripts “e” and “i” to represent the respective
quantities associated with electrons and ions. We note now that
the total current density associated with the plasma (including the
contribution due to positive ions) takes the form
iN e2 1 1
J = N e (vi − ve ) = + E
ω mi me
i 2 2
= ωp,i + ωp,e E = σ E, (11.38)
4πω
11.4 Effect of a background magnetic field 383
2 4πN e2
ωp,e = ,
me
2 4πN e2 2
ωp,i = ≪ ωp,e . (11.39)
mi
Correspondingly, when the motion of the positive ions is taken into
account, the conductivity can be written as
i 2 2
σ= ωp,e + ωp,i . (11.40)
4πω
The fact that the conductivity σ is a scalar simply signifies that
the plasma behaves like an isotropic medium. If we now substitute
(11.40) into the Maxwell’s equations, we can derive, in this case, that
the permittivity of the plasma takes the form
1 2 2
ǫp = 1 − 2
ωp,e + ωp,i . (11.41)
ω
Namely, the permittivity also continues to be a scalar signifying that
the refractive index of the isotropic medium has the form
r
√ 1 2 2 .
np = ǫp = 1 − 2 ωp,e + ωp,i (11.42)
ω
We note now from Eq. (11.39) that
2 2
ωp,e ≫ ωp,i , (11.43)
eB̄
Ω= . (11.45)
mc
In terms of this, the velocity can be expressed as
Ω ie
v + iv × =− E, (11.46)
ω mω
ie 1 i
or, v = − E − 2 (Ω · E) Ω − (E × Ω) ,
mω(1 − ( Ω 2
ω) )
ω ω
ie
Pij vj = − Ei , (11.47)
mω
where
i
Pij = δij + ǫijk Ωk . (11.48)
ω
11.4 Effect of a background magnetic field 385
where
iN e2 1 i
σij = δij − 2 Ωi Ωj − ǫijk Ωk
mω(1 − ( Ω 2
ω) )
ω ω
iωp2 1 i
= δij − 2 Ωi Ωj − ǫijk Ωk . (11.52)
4πω(1 − ( Ω 2
ω) )
ω ω
Substituting (11.51) and (11.52) into the Maxwell’s equations (as well
as using the continuity equation), we find that in the present case,
they take the forms (repeated indices are summed)
∇i (ǫp, ij Ej ) = 0,
∇i Bi = 0,
iω
(∇ × E)i = Bi ,
c
iω
(∇ × B)i = − ǫp, ij Ej , (11.53)
c
386 11 Plasma
Then, taking the curl of the third equation in (11.53) (and using the
fourth), we obtain
ω2
δij k 2 − ki kj Ej = 2 ǫp ij Ej ,
c
2 ki kj
or, ǫp ij − n δij − 2 Ej = 0, (11.56)
k
We note that without any loss of generality, we can choose our coor-
dinate axes (x and y) such that the direction of propagation lies in
the y − z plane, namely,
X XY X(1 ∓ Y )
n 2 = ǫ1 ± ǫ2 = 1 − ± =1−
1−Y2 1−Y2 1−Y2
X ωp2
=1− =1− , (11.64)
1±Y ω(ω ± Ω)
where we have used (11.58). Note that, in this case, it follows from
(11.56) that Ez = 0 and since the electric field is perpendicular to
the direction of propagation (and the magnetic field), the electrons
feel the effect of the magnetic field.
The other special case is when the direction of propagation is
perpendicular to the direction of the magnetic field, namely, θ = π2 .
In this case, we see from (11.62) that the vanishing of the determinant
leads to
(ǫ0 − n2 ) ǫ1 (ǫ1 − n2 ) − ǫ22 = n2 − ǫ0 n2 ǫ1 − ǫ21 + ǫ22 = 0.
(11.65)
Let us go back to Eq. (11.64) and analyze the result in some more
detail. We see that when the direction of propagation is along the
direction of the magnetic field (which we have chosen to correspond
to the z-axis), Ez = 0 if ω 6= ωp . Let us denote the two eigenvalues
for the index of refraction, in this case, as
ωp2
n2± = ǫ1 ± ǫ2 = 1 − . (11.67)
ω(ω ± Ω)
Let us next recall that right and left circularly polarized waves
traveling along the z-axis can be represented respectively as (see, for
example, (6.37))
It follows from this as well as Eqs. (11.69) and (11.71) that traveling
along the direction of the magnetic field in a plasma, the right and
the left circularly polarized waves will suffer different rotations of
phase – the right circularly polarized wave rotating with the index
of refraction n+ while the left circularly polarized wave rotates with
the index of refraction n− . The medium responds differently to right
and left circularly polarized waves. This phenomenon is commonly
known as the Faraday rotation.
We note that in vacuum, a linearly polarized wave can always
be written as a sum of a right and a left circularly polarized wave. If
such a wave, initially linearly polarized along the x-axis, is incident
along the z-axis (the direction of the magnetic field) on a plasma (at
z = 0), then traveling through the plasma, the planes of polarization
of the right and the left circularly polarized waves will rotate as
1
E= (ER + EL )
2
E0 −iωt h i
= e (x̂ − iŷ) eikR z + (x̂ + iŷ) eikL z , (11.73)
2
where
n+ ω n− ω
kR = , kL = . (11.74)
c c
As a result, in traveling through a certain distance z, the tilt in the
polarization will be given by
1 ω
ψ = (kR − kL )z = (n+ − n− )z. (11.75)
2 2c
If the harmonic frequency is high compared to the plasma frequency
as well as the cyclotron frequency, ω ≫ ωp , Ω, then we obtain from
Eqs. (11.67) that
ωp2 Ω
n± ≈ 1 − 2 1∓ . (11.76)
2ω ω
11.6 Alfvén waves 391
This leads to
ω ωp2 Ω
ψ= (n+ − n− )z ≈ z
2c 2cω 2
2πe3 B̄
= N z. (11.77)
m2 c2 ω 2
This shows that the tilt in the polarization planes is proportional to
the distance traveled in the plasma as well as to the number density
of electrons in the plasma. This has, of course, been derived assuming
that the electron density is a constant in the plasma. If the density
changes with distance, as is the case in the ionosphere, then the tilt
is obtained to be
Z z
2πe3 B̄
ψ= 2 2 2 dz ′ N (z ′ ). (11.78)
m c ω 0
In either case, it is clear that the number density of electrons in a
plasma (or the total number of electrons in a volume of unit height)
can be determined from a study of the tilt in the polarization planes.
We recognize that the positive ions will also contribute to the conduc-
tivity tensor as well as to the permittivity. The qualitative structures
of these tensors will be the same and, if we choose the magnetic field
to be along the z-axis, we can write the permittivity tensor as in
(11.59)
ǫ1 iǫ2 0
ǫp,ij = −iǫ2 ǫ1 0 , (11.81)
0 0 ǫ0
where
ǫ0 = 1 − Xe − Xi = 1 − (1 + η)Xe
≈ 1 − Xe ,
Xe Xi
ǫ1 = 1 − −
1 − Ye2 1 − Yi2
(1 + η)Xe (1 − ηYe2 )
=1−
(1 − Ye2 )(1 − η 2 Ye2 )
Xe (1 − ηYe2 )
≈1− ,
(1 − Ye2 )(1 − η 2 Ye2 )
Xe Ye Xi Yi
ǫ2 = −
1 − Ye2 1 − Yi 2
(1 − η 2 )Xe Ye
=
(1 − Ye2 )(1 − η 2 Ye2 )
Xe Ye
≈ . (11.82)
(1 − Ye )(1 − η 2 Ye2 )
2
ǫ0 ≈ 1 − Xe ,
Xe (1 − ηYe2 )
ǫ1 ≈ 1 −
(1 − Ye2 )(1 − η 2 Ye2 )
11.6 Alfvén waves 393
Xe (−ηYe2 ) Xe
≈1− 2 2 2
=1+ ,
(−Ye )(−η Ye ) ηYe2
Xe Ye
ǫ2 ≈ 2
(1 − Ye )(1 − η 2 Ye2 )
Xe Ye Xe
≈ = 2 3. (11.83)
(−Ye2 )(−η 2 Ye2 ) η Ye
Since Ye ≫ 1, it follows that
ǫ1 ≫ ǫ2 , (11.84)
and, therefore, ǫ2 can be neglected in all our manipulations.
The relation (11.84) is significant in simplifying all the expres-
sions. For example, we can now write the matrix (11.62) as
2 ki kj
ǫp,ij − n δij − 2
k
ǫ1 − n 2 0 0
= 0 ǫ1 − n2 cos2 θ n2 sin θ cos θ . (11.85)
0 n2 sin θ cos θ ǫ0 − n2 sin2 θ
The vanishing of the determinant of this matrix is now easily seen to
give
2
ωp,e
Xe
n 2 = ǫ1 ≈ 1 + = 1 + ,
ηYe2 ηΩ2e
ǫ1 ǫ0
n2 =
ǫ0 cos2 θ + ǫ1 sin2 θ
!
2
ωp,e 2 − ω2
ωp,e
≈ 1+ 2 . (11.86)
ηΩ2e 2 − (1 +
ωp,e
ωp,e
sin2 θ)
ηΩ2e
component. Let us note from (11.50) that in this low frequency limit,
we can write for the first root (remember that the magnetic field is
along the z-axis),
ie ω2 iΩe e
ve,y ≈ − − 2 Ex = − Ex ,
me ω Ωe ω me Ω e
ie ω2 iΩi e
vi,y ≈ − 2 − Ex = − Ex . (11.88)
mi ω Ωi ω me Ω e
11.7 Collisions
iN e2
J = −N ev = E = σE, (11.94)
m(ω + iν)
where the four component vector potentials are defined as (see (6.151))
1 ∂Ai
F0i = ∂0 Ai − ∂i A0 = − ∇i Φ = Ei ,
c ∂t
Fij = ∂i Aj − ∂j Ai = ∇i Aj − ∇j Ai
= −ǫijk (∇ × A)k = −ǫijk Bk , (12.3)
397
398 12 Electromagnetic interactions
∂L
∂µ = 0,
∂∂µ Aν
or, ∂µ F µν = 0. (12.7)
As we have seen earlier (see (6.157)), Eq. (12.7) gives only two of
Maxwell’s equations (in vacuum in the absence of sources), namely,
∇ · E = 0,
1 ∂E
∇×B= .
c ∂t
The other two equations are contained in the Bianchi identity satisfied
by the field strength tensor
which follows from the definition of the field strength tensors in (12.1).
We have just described Maxwell’s equations in the absence of
sources. Since Maxwell’s equations (with sources) are also Lorentz
covariant, we can try to introduce sources in a covariant manner as
well. The simplest case is, of course, to consider the interaction of a
charged particle with electromagnetic fields. Thus, we first need to
give a relativistic description of the motion of a free particle. This
is done in a simple manner as follows. First, let us consider a free
particle moving along a trajectory in the four dimensional space-time
manifold as shown in Fig. 12.1. Unlike the non-relativistic case, here
we cannot parameterize the trajectory with t which is not Lorentz
12.1 Relativistic Lagrangian description 399
τ)
x µ(
b
d2 xµ (τ )
m = 0, (12.10)
dτ 2
where m denotes the rest mass of the particle. This is manifestly
Lorentz covariant since m and τ are Lorentz invariant scalars and xµ
is a Lorentz vector. Introducing a relativistic four velocity associated
with the particle as
dxµ (τ )
uµ = = γ(c, v), (12.11)
dτ
where
dx 1 dt
v= , γ=q = , (12.12)
dt 1− v2 dτ
c2
we see that the four velocity transforms like a vector under a Lorentz
transformation and that
η µν uµ uν = γ 2 c2 − v2 = c2 , (12.13)
400 12 Electromagnetic interactions
which follows from (12.9) (or the definitions in (12.11) and (12.12)).
Thus, the components of the four velocity are not all independent,
rather they are constrained by (12.13).
The free particle equation in (12.10) can now be written as
duµ
m = 0. (12.14)
dτ
We see that the relation (12.13) is consistent with the equations of
motion (12.14). Furthermore, the form of the equation in (12.14)
allows us to define a relativistic four momentum associated with the
particle as
v
pµ = muµ = γmc 1, = γmc (1, β) , (12.15)
c
so that the equation of motion, (12.14), can also be written as
dpµ
= 0. (12.16)
dτ
It follows now from (12.13) that
p2 = η µν pµ pν = m2 η µν uµ uν = m2 c2 . (12.17)
Eβ
E = γmc2 , p= , (12.18)
c
which we have used earlier in connection with the method of virtual
photons (see (10.56)). In the non-relativistic limit, β ≪ 1 (or v ≪ c)
and we have
γ ≈ 1,
τ ≈ t,
uµ ≈ (c, v),
pµ ≈ m(c, v), (12.19)
dp
= 0.
dt
12.1 Relativistic Lagrangian description 401
λ → ξ = ξ(λ), (12.21)
we have
dxµ dξ dxµ
→ . (12.22)
dλ dλ dξ
Consequently, under such a reparameterization, the action transforms
as
Z 1
dxµ dxν 2
S = mc dλ ηµν
dλ dλ
Z 2 !1
2
dξ dxµ dxν
→ mc dλ ηµν
dλ dξ dξ
Z 1
dξ dxµ dxν 2
= mc dλ ηµν
dλ dξ dξ
Z 1
dxµ dxν 2
= mc dξ ηµν = S. (12.23)
dξ dξ
This is an invariance under a local transformation much like the gauge
transformation in the case of Maxwell’s theory. Consequently, we can
choose a gauge and, in particular, we can chose the trajectory to be
parameterized by the proper time through the identification (which
is also known as a gauge choice)
λ = τ, (12.24)
402 12 Electromagnetic interactions
which leads to
1
dxµ dxν 2 dτ
ηµν =c = c,
dλ dλ dλ λ=τ
λ=τ
µ ν 2
or, ηµν u u = c . (12.25)
p2 = η µν pµ pν = m2 c2 , (12.28)
which is the Einstein relation (12.18). The fact that not all the com-
ponents of the momenta are independent is a consequence of the
gauge invariance which the system possesses. (Let us note here par-
enthetically that the action in (12.20) is meaningful only for time-like
trajectories. For massless particles, an alternative form of the action
is more useful.)
So far, we have talked about a free relativistic particle. To
describe a relativistic particle subjected to a force, we can generalize
the dynamical equation (12.10) as
d 2 xµ
m = f µ,
dτ 2
dpµ
or, = f µ, (12.29)
dτ
where f µ represents the force four vector. From the fact that
dpµ duµ m d(uµ uµ )
uµ = muµ = = 0, (12.30)
dτ dτ 2 dτ
12.1 Relativistic Lagrangian description 403
uµ f µ = 0. (12.31)
Namely, much like the four velocity and the four momentum, the
components of the relativistic force are not all independent. In fact,
explicitly we have from (12.31) that
uµ f µ = γ(cf 0 − v · f ) = 0,
v·f
or, f 0 = = β · f. (12.32)
c
The physical meaning of Eq. (12.34) is quite clear. Only the compo-
nent of the coordinates along the direction of the velocity of the frame
(and, of course, the time coordinate) transforms, while the compo-
nents perpendicular to the velocity of the frame do not. (Note that
the first parenthesis on the right hand side of the second relation in
(12.34) denotes the orthogonal component while the first term in the
second parenthesis describes the longitudinal component of the coor-
dinate with respect to β.) Using this, then, the general form of the
force in a frame with an instantaneous velocity v can be determined
from its form in the rest frame to be
f 0 = γβ · F,
β(β · F)
f = F + (γ − 1) . (12.35)
β2
404 12 Electromagnetic interactions
The fields, on the right hand side of (12.45), are the rest frame vari-
ables. They can be transformed to the new variables by noting that
the components of the electric fields parallel to the velocity of the
frame do not transform while the orthogonal components do so that
E′k = Ek ,
1
E′⊥ = γ E⊥ + v × B . (12.46)
c
where we have used the standard representation for the current den-
sity associated with a charged particle, namely,
Z
µ dxµ (λ)
j (x) = q dλ δ(x − x(λ)). (12.51)
dλ
The action in (12.48) is invariant under reparameterization and, if we
choose the gauge (12.24) (see also (12.26)), then we obtain
∂L q q
Πµ = µ
= mẋµ + Aµ = pµ + Aµ , (12.52)
∂ ẋ c c
which shows that the interaction is indeed that of minimal coupling.
Here we have identified
dxµ
pµ = mẋµ = m , (12.53)
dτ
which is also known as the kinematic momentum (or sometimes also
as the mechanical momentum) of the particle. The Euler-Lagrange
equation, in this case, leads to
d ∂L ∂L
− = 0,
dτ ∂ ẋµ ∂xµ
d q q
or, pµ + Aµ − (∂µ Aν )ẋν = 0. (12.54)
dτ c c
Noting that
dAµ
= (∂ ν Aµ )ẋν , (12.55)
dτ
the Euler-Lagrange equation in (12.54) takes the form
dpµ q q
= (∂ µ Aν − ∂ ν Aµ ) ẋν = F µν uν . (12.56)
dτ c c
This is indeed the dynamical equation that we have discussed in
(12.41).
We have, of course, neglected the dynamics of the electromag-
netic fields in this discussion. If we add the Maxwell term (12.4)
to the action, then, the complete set of coupled equations of motion
takes the form
dxµ
m = pµ ,
dτ
dpµ q
= F µν uν ,
dτ c
4π ν
∂µ F µν = j , (12.57)
c
408 12 Electromagnetic interactions
where the current density has the form in (12.51) (with the identifi-
cation λ = τ ). We note here that sometimes the second equation in
(12.57) is also written as
dpµ q
m = F µν pν , (12.58)
dτ c
with the identification in (12.15).
Let us next work out the solutions for a few interacting systems. Let
us recall that the equations of motion (12.41) can be written as (recall
dt
that pµ = muµ = γm(c, v) and γ = dτ )
d mc2
q = qv · E,
dt 1− v2
c2
d mv 1
q =q E+ v×B , (12.59)
dt 1− v
2 c
c2
E = −∇Φ,
mc2
or, q + qΦ = constant. (12.60)
2
1 − vc2
E = (E, 0, 0), B = 0,
Φ = −Ex, A = 0. (12.61)
mc2
q = qEx,
2
1 − vc2
mv
q = qEt, (12.62)
v2
1− c2
The first form of the relation in (12.65) clearly shows that the motion
is hyperbolic in the presence of a constant acceleration. This has to
be contrasted with the non-relativistic case where
1
x = x0 + at2 ,
2
dx ct
v= =q 22 . (12.66)
dt m c
+ t 2
q2 E 2
qE
≪ 1, (12.67)
mc
mc
then, (in a time interval t ≪ qE ) we have from Eqs. (12.65) and
(12.66),
qE
v≈ t = at,
m
mc2 1 qE 2
x≈ + t
qE 2 m
1 2
= x0 + at , (12.68)
2
v2 m2 c2
or, 1 − = 2 . (12.70)
c2 m2 c2 + p(0) + qEt
Here, we have used the standard formula from the table of integrals
(see, for example, Gradshteyn and Ryzhik, 2.261). We note from Eq.
(12.70) that as t → ±∞, the speed of the particle approaches the
speed of light as before. However, in the present case for the particle
moving in a plane, the motion cannot be characterized as hyperbolic.
E = 0, B = Bẑ, (12.73)
where E represents the energy of the system (recall that the first
equation gives the energy relation and we are denoting energy by E
to avoid any confusion with the electric field). Furthermore, from
the second equation we see that vz is also a constant. It follows,
therefore, that
p
v⊥ = v 2 − vz2 , (12.76)
z = z0 + vz t. (12.77)
12.3 Motion in a uniform magnetic field 413
The second equation can now be written out explicitly for the
x, y components of the velocity to give
dvx qB
= vy ,
dt γmc
dvy qB
=− vx . (12.78)
dt γmc
Recalling that vx2 + vy2 = v⊥2 is a constant, we can write the solutions
of Eq. (12.78) as
qB
vx = −v⊥ sin t ,
γmc
qB
vy = −v⊥ cos t . (12.79)
γmc
v⊥ γmcv⊥ Ev⊥
R= = = , (12.83)
ω qB qBc
so that
E · B = 0. (12.86)
1
Fµν F µν = η µλ η νρ Fµν Fλρ , Fµν F̃ µν = ǫµνλρ Fµν Fλρ , (12.87)
2
Furthermore, since (F̃ µν = 12 ǫµνλρ Fλρ is also known as the dual field
strength tensor, ǫ0ijk = ǫijk )
1 0ijk
F̃ 0i = ǫ Fjk = −Bi ,
2
F̃ ij = ǫij0k F0k = ǫijk Ek , (12.89)
it follows that
E′k = Ek ,
u
E′⊥ = γ(u)(E⊥ + × B⊥ ),
c
B′k = Bk ,
u
B′⊥ = γ(u)(B⊥ − × E⊥ ), (12.92)
c
where the “parallel” and the “perpendicular” decompositions are
with respect to the velocity u.
Let us first consider the case when |E| = E > B = |B| and
consider a frame moving with a velocity u such that
u E×B
= . (12.93)
c E2
416 12 Electromagnetic interactions
u2 E 2 B 2 − (E · B)2 B2
= = < 1, (12.94)
c2 E4 E2
where we have used the fact that the electric and the magnetic fields
are orthogonal. In this case, therefore, we have
1 1
γ(u) = q =q . (12.95)
u2 B2
1− c2
1− E2
Ek = 0 = Bk , ⇒ E′k = 0 = B′k
= γ(u)(B − B) = 0. (12.96)
Here, we have used the orthogonality of the electric and the magnetic
fields in the intermediate steps.
This shows that for orthogonal electric and magnetic fields, when
E > B, we can find a Lorentz frame where the magnetic field iden-
tically vanishes. Furthermore, the electric field, in this new frame, is
along the same direction as the original field, only scaled by a Lorentz
factor. The solution to this problem is, as before, unbounded motion
and is not very interesting.
12.4 Motion in crossed fields 417
Let us next consider the case when E < B and choose a Lorentz
transformation with
u E×B
= . (12.97)
c B2
In this case, we have
u2 E 2 B 2 − (E · B)2 E2
= = < 1, (12.98)
c2 B4 B2
so that it is an allowed transformation. For the present case,
1 1
γ(u) = q =q . (12.99)
u2 E2
1− c2 1− B2
Ek = 0 = Bk .
Furthermore, we have
u (E × B) × B
E′ = γ(u) E + × B = γ(u) E +
c B2
(E · B)B − (B2 )E
= γ(u) E +
B2
= γ(u)(E − E) = 0,
′ u (E × B) × E
B = γ(u) B − × E = γ(u) B −
c B2
(E2 )B − (E · B)E
= γ(u) B −
B2
r
E2
= 1 − 2 B = γ −1 (u) B. (12.100)
B
Thus, in this case, we see that the electric field vanishes in the
new frame and the magnetic field is along the same direction as the
original field, but scaled by a field dependent Lorentz factor. For
418 12 Electromagnetic interactions
where, as before,
γ(v)mcv⊥
R= .
qB
Here we have also used the fact that, under an inverse Lorentz trans-
formation,
R′ → R, v ′ → v, ′
v⊥ → v⊥ ,
and so on. In this case, we see that the motion corresponds to an el-
lipsoidal motion superimposed with a constant “drift” velocity along
the y axis. This is known as a trochoidal motion.
We note from the equations of motion (12.59), that when
v E×B u
= = , (12.106)
c B2 c
the Lorentz force identically vanishes (see also the first relation in
(12.100)), namely,
(E × B) × B
q E+ = q (E − E) = 0. (12.107)
B2
In this case, the particle would move along the initial trajectory com-
pletely undeflected by the presence of the fields, independent of its
mass. (Namely, f = 0 also implies that f 0 = v·f c = 0, see (12.32).)
This is a very important feature which is utilized in creating a veloc-
ity filter. Namely, if a number of particles are incident on a region
with crossed electric and magnetic fields (with E < B), then only
those particles that have the initial velocity coinciding with (12.106)
would travel undeflected. Correspondingly, one can choose different
electric and magnetic fields to select the desired particles with a given
velocity.
∇ · B = 0,
that B = B(x, y) and that, since in general, the curl of the magnetic
field will not vanish, such a configuration must have an associated
steady current and we see that, in this case, the magnetic field can-
not depend on the (longitudinal) z coordinate. As we noted earlier,
in some physical situations, we do need a dependence of the magnetic
field on the longitudinal coordinate which can arise only if the mag-
netic field is not parallel everywhere. For confinement of plasma to a
small region in space, for example, we would expect the plasma not
to extend beyond a certain vertical and horizontal dimension. Under
the action of a magnetic field, as we have seen, the particle moves
in an orbit whose radius is determined by the magnetic field. Thus,
we see that with a suitable choice of a magnetic field, the plasma
can be easily confined to a given vertical dimension. Let us note
also that if we have a magnetic field that is converging along the z-
axis in some region, then the magnetic force acting on the charged
particles will be so as to force it into the interior of the region (see
Fig. 12.2). Therefore, there will be a component of the force along
the z-axis which would decelerate the “drift” velocity. As a result,
at some point along the horizontal (z) axis, the “drift” velocity will
vanish (the “guiding center” will come to rest) and then, the direc-
tion of the drift will change. This would lead to a containment of
the charged particle along the z-axis as well, as shown in Fig. 12.3.
This process is known as “mirroring” and is used to trap plasma by a
12.5 Motion in a slowly varying magnetic field 421
v×B
v×B
Figure 12.2: The Lorentz force directing the particle into the interior
of the region.
1 ∂Bz
ρBρ = − ρ2 ,
2 ∂z
1 ∂Bz
or, Bρ = − ρ . (12.110)
2 ∂z
The constant of integration is easily seen to vanish (with the assump-
tion that Bρ = 0 at ρ = 0). It is clear now from (12.110), that for
R ∂B
slowly varying fields (| B ∂z | ≪ 1), Bρ is much smaller than Bz within
a radius of the size of R. Consequently, we can approximate
B = |B| ≈ Bz , (12.111)
to write
1 ∂B
Bρ ≈ − ρ . (12.112)
2 ∂z
From the time component of the equations of motion, (12.59),
we note that in the absence of an electric field, v is constant. How-
ever, unlike the earlier case of motion in a uniform magnetic field,
here a non-vanishing Bρ leads to a magnetic force along the z-axis.
As a result, vz and, therefore, v⊥ will no longer be constant. In
fact, looking at the equation for the z component of the velocity (see
(12.59)) for a particle moving in an orbit of radius R, we have (γ is
a constant since v is)
d q
(γmvz ) = v⊥ Bρ ,
dt c
dvz 1 qRv⊥ ∂B 1 v 2 ∂B
or, =− =− ⊥
dt 2 γmc ∂z 2 B ∂z
2 (0)
1 v⊥ ∂B
≈− . (12.113)
2 B(0) ∂z
Here, we have used the fact that the velocity of the particle, in the x−
y plane, is given by v⊥ = −v⊥ φ̂ (see (12.79)), v⊥ = qRB
γmc (see (12.83))
as well as the relation in (12.112) for a particle moving in an orbit of
radius R. Furthermore, since the fields are slowly varying, we have
approximated the coefficient multiplying ∂B ∂z (which is already small)
by its value at z = 0 which can be thought of as the leading order
approximation. It now follows from (12.113) that (the assumption
here is that at t = 0, the particle is at z = 0 and that at t its
12.5 Motion in a slowly varying magnetic field 423
coordinate is z)
1 2 v 2 (0)
or, vz − vz2 (0) = − ⊥ (B(z) − B(0)),
2 2B(0)
2 (0)B(z)
v⊥
or, vz2 = v 2 − , (12.114)
B(0)
vz2 (0) + v⊥
2
(0) = v 2 (0) = v 2 , (12.115)
2 (0)B(z)
v⊥
v 2 − vz2 = v⊥
2
= ,
B(0)
2
v⊥ v 2 (0)
or, = ⊥ . (12.116)
B(z) B(0)
2
v⊥ qRv⊥
= , (12.117)
B(z) γmc
γ 2 m2 c2 v⊥
2
πR2 B = π B
q2B 2
2
πγ 2 m2 c2 v⊥
=
q2 B(z)
2 (0)
πγ 2 m2 c2 v⊥
= , (12.118)
q2 B(0)
qω qωR2
µ = πR2 =
2π 2
2 2 (0)
γmc v⊥ γmc v⊥
= = . (12.119)
2 B(z) 2 B(0)
∂L q q
Πµ = = m∆ẋµ + Aµ = ∆pµ + Aµ . (12.127)
∂ ẋµ c c
Let us first consider the case when Mµν is a constant tensor.
In this case, the Euler-Lagrange equations following from the La-
grangian, in the gauge (12.24), lead to
dxµ
m = pµ = muµ ,
dτ
d ∂L ∂L
− = 0,
dτ ∂ ẋµ ∂xµ
dpµ 1 h q µν i
or, = F uν + m(∂ν ∆) c2 η µν − uµ uν
dτ ∆ c
1 q µν 1 λρ 2 µν µ ν
= F uν + 2 M Fλρ,ν c η − u u .
∆ c 2c
(12.128)
∆ ≈ 1, (12.130)
in which case, the dynamical equation for the momentum takes the
form
dpµ q µν 1 λρ 2 µν µ ν
≈ F uν + 2 M Fλρ,ν c η − u u . (12.131)
dτ c 2c
1 µνλρ
sµ = ǫ pν Mλρ . (12.132)
2q
For completeness, let us also note that we can invert the relation
(12.132) to write
q
Mµν = ǫµνλρ pλ sρ . (12.136)
2
It follows from this that
pµ Mµν = muµ Mµν = 0 = sµ Mµν . (12.137)
Let us note that in the rest frame of the particle, we can identify
the space components of the spin variable as the spin of the particle
(up to multiplicative factors). Furthermore, in the rest frame of the
particle, we know the equation for the spin in the presence of external
electromagnetic fields to correspond to
dS q g f
= S×B+ S×E . (12.138)
dt mc 2 2
From this, we can transform to any other Lorentz frame in a simple
manner. First, let us note that we can write
(S × B)i = −Fij Sj = Fij Sj = Fiν srest
ν ,
which is consistent with (12.133) and holds only because of the pres-
ence of the last term in (12.141).
Equation (12.141) is still not in a simple form. It can be further
simplified using the dynamical equations for the particle in (12.128).
In fact, for a uniform field (and for ∆ ≈ 1), we can use (12.128) to
write the equation for the spin in any Lorentz frame as
dsµ q g µν 1 g
= F sν + 2 − 1 sλ F λρ uρ uµ
dτ mc 2 c 2
f µν 1 λρ µ
+ F̃ sν + 2 sλ F̃ uρ u . (12.143)
2 c
There are several things to observe from this equation. We note
that for a particle without an electric dipole moment (f = 0) moving
in a uniform magnetic field (F0i = 0), the equation for the spin takes
the form
dsµ q µν (g − 2) λρ µ
= gF sν + s λ F uρ u . (12.144)
dτ 2mc c2
Writing out the equations explicitly, we obtain,
ds gq γ 2 q(g − 2)
= s×B+ (v · (s × B)) v,
dτ 2mc 2mc3
ds0 1 d(v · s) γq(g − 2)
= = v · (s × B), (12.145)
dτ c dτ 2mc2
where we have made the identification in (12.134). The time rate
of variation of the longitudinal component of the spin can now be
calculated using (12.128) and the second relation in (12.145).
d(v̂ · s) 1 d(v · s) (v · s) dv
= − 3
v·
dτ v dτ v dτ
1 d(v · s)
=
v dτ
γq(g − 2)
= v̂ · (s × B),
2mc2
d(v̂ · s) q(g − 2)
or, = v̂ · (s × B). (12.146)
dt 2mc2
Here the second term on the right hand side in the first line van-
ishes because v · dv
dτ = 0 which follows from the second equation in
(12.128) for a uniform (constant) magnetic field. Equations (12.145)
430 12 Electromagnetic interactions
431
432 13 Scattering and diffraction
where jℓ (kr) denote spherical Bessel functions and we have used the
definition
r
(2ℓ + 1)
Yℓ,0 (θ) = Pℓ (cos θ).
4π
This shows that a scalar plane wave contains only waves with angular
momentum projection m = 0.
Let us next recall that a circularly polarized harmonic electro-
magnetic wave traveling along the z-axis in vacuum will have electric
and magnetic fields of the forms (factoring out the time dependence,
see also (11.72))
1
ER,L (x) = √ (x̂ ∓ iŷ) eikz ,
2
BR,L (x) = ẑ × ER,L (x) = ±i ER,L (x), (13.3)
where the first sign corresponds to a right circularly polarized wave
and the second to a left circularly polarized wave. (For simplicity, we
13.1 Scattering from a perfectly conducting sphere 433
are assuming fields of unit intensity.) Of course, the right and left
circularly polarized fields in (13.3) can be expressed as superpositions
of electric and magnetic multipole fields which we have discussed
earlier. Therefore, for the fields in (13.3), we can write in general
(see the discussion in sections 9.4 and 9.5 on multipole expansions,
in particular, see (9.116) and (9.117))
Xh
ER,L (x) = a∓ (ℓ, m)jℓ (kr)Yℓ,m
ℓ,m
i i
+ b∓ (ℓ, m)∇ × (jℓ (kr)Yℓ,m ) ,
k
Xh i
BR,L (x) = − a∓ (ℓ, m)∇ × (jℓ (kr)Yℓ,m )
k
ℓ,m
i
+ b∓ (ℓ, m)jℓ (kr)Yℓ,m . (13.4)
Here, we have used the well behaved spherical Bessel functions for the
expansion of the plane waves (and not the spherical Neumann func-
tion) since the plane wave is well behaved everywhere. The constants
a∓ (ℓ, m) and b∓ (ℓ, m) specify respectively the amounts of magnetic
multipole terms and the electric multipole terms present in the wave
corresponding to a given (ℓ, m). Furthermore, Yℓ,m (θ, φ) represent
the vector spherical harmonics defined in (9.114).
To determine the expansion coefficients a∓ (ℓ, m), b∓ (ℓ, m), we
need to understand some of the properties of the vector spherical
harmonics. We have already seen in (9.115) that
Z
dΩ Yℓ∗′ ,m′ (θ, φ) · Yℓ,m (θ, φ) = δℓℓ′ δmm′ . (13.5)
Let us next note that for any arbitrary radial function fℓ (r), we have
′
i
(1)
± i(e2iδℓ − 1)hℓ (kr)Yℓ,∓1 . (13.30)
Let us next consider the asymptotic forms of these fields for
large r, far away from the scattering source. In this case, we know
that
(1) eikr
hℓ (kr) → (−i)ℓ+1 . (13.31)
kr
It follows, therefore, from Eq. (13.9) as well as (13.31) that for large
r
(1) (−i)ℓ eikr
∇ × (hℓ (kr)Yℓ,∓1 ) → r̂ × Yℓ,∓1 . (13.32)
r
Substituting these into the expression for the scattered electric and
magnetic fields, we obtain that, for large r,
(sc) eikr
ER,L → f∓ (θ, φ),
r
(sc) (sc)
BR,L → r̂ × ER,L , (13.33)
where we have defined
∞ p
X 2π(2ℓ + 1)
f∓ (θ, φ) =
k
ℓ=1
h ′
i
× eiδℓ sin δℓ Yℓ,∓1 ∓ ieiδℓ sin δℓ′ r̂ × Yℓ,∓1 . (13.34)
440 13 Scattering and diffraction
∞
2π X
= 2
(2ℓ + 1) sin2 δℓ + sin2 δℓ′ . (13.38)
k
ℓ=1
This shows that the magnetic and the electric multipole fields con-
tribute incoherently to the total scattering cross section (there is no
cross term because of (13.24)) although, as we will see, there can be
interference terms present in the differential cross section.
13.1 Scattering from a perfectly conducting sphere 441
(kd)2ℓ+1
or, δℓ ≈ − ,
(2ℓ + 1)[(2ℓ − 1)!!]2
∂
(rj (kr)) ℓ+1 (kd)2ℓ+1
ℓ
tan δℓ′ = ∂r →
∂
∂r (rηℓ (kr))
ℓ (2ℓ + 1)[(2ℓ − 1)!!]2
r=d
ℓ+1
or, δℓ′ ≈ − δℓ . (13.39)
ℓ
This shows that the higher angular momentum components of the
phase shifts fall off rapidly with ℓ in the long wave length limit. There-
fore, we can approximate the expression for the scattering amplitude
by keeping only the lowest order term corresponding to ℓ = 1. We
note from (13.39) that
(kd)3
δ1 ≈ − , δ1′ ≈ −2δ1 . (13.40)
3
Using this, the scattering amplitude in this approximation becomes
√
6π
f∓ (θ, φ) ≈ δ1 (Y1,∓1 ± 2i r̂ × Y1,∓1 ) . (13.41)
k
Furthermore, using the properties of the angular momentum opera-
tors and the spherical harmonics, we obtain
We note that the differential cross section is the same for both the
right and the left circularly polarized waves. Furthermore, it is pro-
portional to the fourth power of the frequency and has a peak in the
backward direction θ = π. Integrating this, we obtain
Z
10πd2
σR,L = dΩ |f∓ (θ, φ)|2 =
total
(kd)4 . (13.43)
3
The interference term does not contribute to the total cross section.
The dependence of the scattering cross section on the fourth power
of the frequency is a characteristic of dipole fields and is known as
Rayleigh’s law.
∇ × (∇ × V) − k 2 V = 4πf ,
or, ∂i ∂j − δij ∇2 + k 2 Vj = 4πfi , (13.47)
where Vi stands for the three components of either the electric or the
magnetic fields and fi the sources. (Repeated indices are assumed
to be summed.) To solve such an inhomogeneous equation, we will
make use of the method of Green’s functions. In this case, we note
that the Green’s function will be a tensor and from (13.47) it follows
that the equation satisfied by the Green’s function would have the
form
∂i ∂m − δim ∇2 + k 2 Gmj (x − x′ ) = 4πδij δ3 (x − x′ ). (13.48)
The form of the Green’s function satisfying (13.48) can now be de-
termined along the lines discussed in section 6.8. First, we note that
444 13 Scattering and diffraction
the Green’s function for the scalar Helmholtz equation (9.87) satisfy-
ing the boundary conditions that at large distances it represents an
outgoing wave has the form
eikx
g(x) = , (13.51)
x
leads to
Z
1
Ei (x, ω) = − ds′ n′j Em − n′m Ej ∂j′ Gmi
4π
ik
− √ ǫjmℓ Gmi n′j Bℓ
ǫµ
Z
1 ′ ′
′ ik ′
=− ds ǫjℓm n × E m ∂j Gℓi − √ Bm ǫmjℓ nj Gℓi .
4π ǫµ
(13.55)
The combination
Similarly, identifying
Vi (x′ ) = Bi (x′ ), Aij (x, x′ ) = Gij (x − x′ ), (13.60)
in a source free region (within the bounding surface), we obtain from
(13.53),
Z
1
Bi (x, ω) = − ds′ n′j Bm − n′m Bj ∂j′ Gmi
4π
√
+i ǫµk ǫjmℓ Gmi n′j Eℓ
Z
1 √
=− ds′ ǫjℓm n′ × B m ∂j′ Gℓi + i ǫµk Em ǫmjℓ n′j Gℓi .
4π
(13.61)
We see once again that if we have the electric tensor Green’s function,
then the magnetic field can be solved uniquely in terms of the bound-
ary values of its tangential components specified on a given surface
as
Z
1 (e)
Bi (x, ω) = − ds′ n′ × B(x′ ) m ǫmjℓ ∂j′ Gℓi (x − x′ ).
4π
(13.62)
The electric field then follows from (13.44) to be
i
Ei (x, ω) = √ (∇ × B(x, ω))i
ǫµk
Z
i
= √ ds′ n′ × B(x′ ) j ǫjℓm ǫipq ∂ℓ′ ∂p′ G(e) ′
mq (x − x ).
4π ǫµk
(13.63)
This analysis makes it clear that the electric and the magnetic fields
can be determined uniquely if the values of the tangential compo-
nents of either the electric or the magnetic field are given on a given
boundary surface. However, specifying the tangential components of
both the electric as well as the magnetic fields on a boundary over-
specifies the system unless the tangent components of the electric
and the magnetic fields on the boundary are consistent. In such a
case, the electric field at any point within the region bounded by the
surface is given by (13.55), namely,
Z h
1
Ei (x, ω) = − ds′ ǫjℓm n′ × E m ∂j′ Gℓi
4π
ik i
− √ Bm ǫmjℓ n′j Gℓi . (13.64)
ǫµ
13.2 Kirchhoff’s approximation 447
The relations obtained above are exact in the sense that there
has been no approximation used so far. However, determining the
boundary conditions in a given problem, (namely, determining the
tangential components of the electric and the magnetic fields on the
boundary) is in general difficult and it is here that approximations
creep in. In particular, Kirchhoff’s approximation uses the notions
from geometrical optics in estimating the tangential components of
the fields on a boundary surface. For example, let us assume that we
are considering diffraction of electromagnetic waves from a spherical
conducting surface. In this case, the surface of the sphere can be
divided into two parts, one that is illuminated by the incident wave
and the other that is in the shadow. Kirchhoff’s approximation con-
sists of assuming that in the illuminated part of the spherical surface,
we have already seen from our studies on reflection from a perfectly
conducting surface (see the discussion in sections 8.2 and 8.3) that
n × E = n × E(inc) + E(refl) = 0,
n × B = n × B(inc) + B(refl) = 2n × B(inc) . (13.65)
On the other hand, in the shadow part of the spherical surface, Kirch-
hoff’s approximation assumes that there is no total field so that
(sc) (sc)
n × E(inc) = −n × EII , n × B(inc) = −n × BII . (13.67)
Here the subscripts I and II refer to the two regions of the surface of
the sphere.
Without going into too much technical details, let us indicate
how Kirchhoff’s approximation can be used to calculate the diffrac-
tion of a plane wave from a conducting sphere. Let us assume that
the plane wave is incident in vacuum (ǫ = 1 = µ) along the z-axis on a
perfectly conducting sphere of radius a. Thus, we identify k̂ = ẑ = n′ .
We assume that the origin of the coordinate system coincides with
the center of the sphere. Then, we can consider a point outside the
sphere to be contained in a region bounded by the surface of the con-
ducting sphere as well as the large spherical surface at infinity. With
a little bit of analysis, it can be shown that the surface integral over
the large sphere at infinity vanishes. Thus, we can write the scattered
448 13 Scattering and diffraction
eikx −ik·x′ ′
g(x − x′ ) ≈ e = g(x) e−ik·x ,
x
ki kj ′
Gij (x − x ) ≈ − 2 + δij g(x) e−ik·x ,
′
(13.70)
k
451
452