Book Introduction To Electrostatic
Book Introduction To Electrostatic
Book Introduction To Electrostatic
Introduction to Electrostatics
Coulomb’s Law
• r21 = r2 − r1
1
Chapter 1 2
F 21
q2
r2 F 12
r1 q1
In SI units :
• The charges q1, q2 are measured in Coulombs (C), and defined via the
magnetic effects of currents.
• The forces on the two charges are equal and opposite, obeying Newton’s
third law: F12 = −F21.
Electric Field: The electric field E at r is defined as the force acting on a unit
test charge at that point. More strictly,
F (r)
E(r) = lim ,
q→0 q
so that the electric field due to the test charge can be ignored.
Chapter 1 3
In the above we have looked at the fields due to single, isolated point-like
charges. In this section, we will explore the second emperical ingredient necessary
for our understanding of electrostatic fields, the linear superposition principle.
Therefore, the electrostatic field at the point r due to the element of charge dq at
r′ is
ρ(r′ ) (r − r′ )∆V ′
∆E(r) = k
| r − r′ |3
where we take ∆E(r) −→ 0 as r −→ ∞. We now use the principle of linear
superposition to write that the resultant field at r as a sum over the elements
∆V ′ in V
X ρ(r ′ ) (r − r ′ )∆V ′
E(r) = k
∆V ′ | r − r′ |3
In the limit that ∆V ′ becomes infinitessimal, we have
Much of the rest of this course is centred on methods for obtaining the electrostatic
field, and we begin with one of the simplest - Gauss’ Law.
Suppose that the charge density ρ(r) is the sole source of the electrostatic field
E(r). Gauss’ Law relates the flux of E out of a closed surface S bounding a
volume V to the total charge Q contained within V
Chapter 1 5
dS
Gauss’ Law provides a powerful way to compute the electrostatic field for the
case where there is spherical, or even cylindrical, symmetry. It will also form the
starting point for our derivation of Laplace’s equation later in the course.
Consider a point charge q placed at the origin (not necessarily inside V ), and the
electrostatic flux across an area dS.
Chapter 1 6
dS
q
θ
dΩ
Then we have
dS cos θ
E · dS = kq
r2
= kq dΩ
• If the charge q is outside the volume, then the total flux V E ·dS is zero; the
R
dS dS
q
dS
Outside Inside
Though this provides an intuitive interpretation of Gauss’ Law, we will now pro-
ceed to a more formal proof.
We will begin by proving Gauss’ Law for a single, pointlike charge q at the origin.
Origin inside V :
E(r) is undefined at r = 0. Therefore define V to be the region between the
closed surfaces S ′ and S, where S ′ is a small sphere of radius ǫ centred at the
origin:
S
dS
dS’
O S’
ε V
Introduce spherical polar coordinates (r, θ, ψ). Then on the sphere S ′ we have:
dS = −ǫ2 sin θ dθ dψ er ,
Chapter 1 9
where the outward normal for S ′ points towards the origin. Therefore
er
Z Z 2π Z π !
E · dS = kq 2 · (−ǫ2 sin θ dθ dψ er )
S′ 0 0 r r=ǫ
= −4πkq independent of ǫ
We now let ǫ → 0, so that V → total volume within S, and we have
Z q
E · dS = 4πkq = in SI units
S ǫ0
so that the theorem is proved.
If the point charge is at the point r1 , then we have
r − r1
E(r) = kq .
| r − r1 |3
By changing variables to ρ = r − r1 it is easy to show
Z
4πkq = q/ǫ0 in SI units if r1 ∈ V
E · dS =
S 0 otherwise
We can extend the proof of Gauss’ Law for a single charge distribution to a set of
N point charges {qi} at {ri } using the linear-superposition principle:
N
X
E(r) = Ei(r)
i=1
where E(r) is the total electrostatic field at the point r, and Ei (r) is the electro-
static field at the point r due to the charge qi at the point ri. Applying Gauss’
Law for point charges proved above, we have
Z
4πkqi = qi/ǫ0 in SI units if ri ∈ V
Ei · dS =
S 0 otherwise
Hence
Z XZ X
E · dS = Ei · dS = 4πk qi
S i S i,ri ∈V
Q
= 4πkQ = (in SI units)
ǫ0
Chapter 1 10
This we prove by exact analogy with derivation of the electrostatic field for a
continuous distribution: we divide up the volume V into elements of volume ∆V ′ ,
centred at r′ , and obtain
Z
ρ(r′ ) ∆V ′
X
E · dS = 4πk
S ∆V ′ ∈V
′
Z
∆V →0
−→ 4πk ρ(r′ ) dV ′ = 4πkQ,
V
not included
included
Chapter 1 11
Spherical Symmetry
where Q(r) = ρ(r′ ) dV is the total charge contained within the sphere of radius
R
V
r.
Thus we have
kQ(r) Q(r)
E(r) = er = er in SI units.
r2 4πǫ0 r2
Note that outside a spherically symmetric charge distribution, the field is the
same as if we had a point-like charge Q(r) at the origin.
Cylindrical Symmetry
0 L
Eρ
The field will depend solely on ρ, and therefore must be in the eρ direction,
E(r) = E(ρ)eρ. Applying Gauss’ Law to the cylinder we have
Z
E · dS = 4πkQ(ρ, L)
S
Thus
2kQ(ρ, L) 2Q(ρ, L)
E(ρ) = = in SI units
ρL 4πǫ0 ρL
Example: Infinitely long, thin rod carrying charge λ per unit length. Thus,
Q(ρ, L) = λL and we have
λ
E(ρ) = .
2πǫ0 ρ
Chapter 1 13
w < ρ << l
where w and l and the width and the length of the rod respectively.
and thus Z
{∇ · E − 4πkρ} dV ′ = 0.
V
This applies for any volume V , and therefore the integrand itself must vanish:
ρ
∇ · E = 4πkρ = . (1.1)
ǫ0
This is Maxwell’s First Equation (ME1). ME1 is essentially an expression of
Gauss’ law in differential form.
Chapter 1 14
ME1 has provided us with a differential equation to describe the electric field,
E(r), but it would be easier were we able to work with a scalar quantity. The
scalar potential provides a means of so doing.
Scalar Potential
• Given a vector field A(r), under what conditions can we write A as the
gradient of a scalar field φ, viz. A(r) = −∇ φ(r), where the minus sign
is conventional?
R R
Let ∇ × A(r) = 0 in R, and consider any two curves, C1 and C2 from the point
r0 to the point r in R. Introduce the closed curve C = C1 − C2, and let S be a
surface spanning C.
r
C1
r0
C2
Chapter 1 16
Thus the scalar potential φ(r) of the vector field A(r) defined by
Z r
φ(r) = − A(r′ ) · dr′
r0
is independent of the path of integration joining r0 and r.
Proof that (2) implies (3)
Consider two neighbouring points r and r + dr. Define the scalar potential as
before: Z r
φ(r) = − A(r′ ) · dr′
r0
r
dr
r + dr
r0
Chapter 1 17
1.5.2 Terminology
• The field φ(r) is the scalar potential for the vector field A(r).
1.5.3 Uniqueness
∇φ − ∇ψ = A − A = 0
Therefore
∇ (ψ − φ) = 0
Integration of this equation wrt any of x, y, or z gives
ψ − φ = constant
Therefore
ψ = φ + constant
The absolute value of a scalar potential has no meaning, only potential dif-
ferences are significant.
Chapter 1 19
After the digression on subject of the scalar potentials, it is time to show that the
electrostatic field is, indeed, irrotational.
The central result of this chapter was the expression for the electrostatic field due
to a continuous charge distribution
Z ρ(r′ )(r − r′ )dV ′
E(r) = k .
V | r − r′ |3
Thus we have
ρ(r ′ )(r − r′) ′
Z
∇ × E(r) = ∇× dV
| r − r′ |3
Z 1 1
ρ(r′ ) ∇ × (r − r ′ ) + ′
dV ′
= ∇ × (r − r )
V | r − r′ |3 | r − r′ |3
−3(r − r ′ )
Z
′ ′
dV ′
= ρ(r ) ′ 5
× (r − r ) + 0
|r−r |
= 0
We have shown that the scalar potential φ(r) for an irrotational vector field A(r)
can be constructed via
r Z
φ(r) = −A(r′ ) · dr′
r0
for some suitably chosen r0 and any path which joins r0 and r. Sensible choices
for r0 are often r0 = 0 or r0 = ∞.
We have also shown that the line integral is independent of the path of integration
Chapter 1 20
r′ = λ r where {0 ≤ λ ≤ 1}
Example:
Let A(r) = (a · r) a where a is a constant vector.
It is easy to show that ∇ × ((a · r) a) = 0. Thus
Z r
φ(r) = − A(r′) · dr′
0
Z r
= − (a · r′ ) a · dr′
0
Z 1
= − (a · λ r) a · (dλ r)
0
Z 1
= − (a · r)2 λ dλ
0
1
= − (a · r)2
2
Of course, this is all rather artifical. What we really want to do is to obtain φ
and A from first principles.
Chapter 1 21
We have seen that, for the case of a point-charge at the origin, the electric field
is singular at r = 0. In such cases, it is not possible to obtain the corresponding
scalar potential at r by integration along a path from the origin. All is not
lost - remember that the starting point for our path is arbitrary, and often it is
convenient to take it at infinity.
Example: Electric field due to point charge at r = 0: E(r) = kqr/r3, so that
E(r = 0) is singular, and hence undefined. As in the proof of Gauss’ law, our
region R must exclude an infinitessimal sphere centred at r = 0.
Here we choose a path from r0 = ∞, yielding
Z r Z 1
φ(r) = − E(r′ ) · dr′ = − E(λr) · dλ r
∞ ∞
dλ r2
Z 1
= −kq
∞ λ2 r 3
1
= kq
r
Thus we have the famous 1/r potential due to a point charge.
Because of the linearity of the gradient operation, we can impose the linear su-
perposition principle on the potential, and hence obtain an expression for the
potential due to an extended charge distribution:
Z
ρ(r′ )dV ′
φ(r) = k (1.5)
V | r − r′ |
In this case, ∇× A = 0 does not imply the existence of a scalar potential function.
Example: Work using cylindrical coordinates (ρ, φ, z). A vector field A, with
a
Aρ = Az = 0, Aφ =
ρ
Chapter 1 22
eθ
eρ
ρ
Excluded
region
θ
To see how the name conservative field arises, consider a vector field F (r) cor-
responding to the only force acting on some test particle of mass m. The work
done by the force in going around a closed curve C is
I
W = F (r) · dr
C
For a conservative force, ∇ × F = 0, the earlier theorems tell us:
Chapter 1 23
• The total work done by the force in moving the particle around a closed curve
is zero.
We will now show that for a conservative force, the total energy is constant in
time.
Proof
The particle moves under the influence of Newton’s Second Law:
mr̈ = F (r).
Consider a small displacement dr taking time dt along the path followed by the
particle. Then we have
We have seen that the existence of a scalar potential is associated with the irrota-
tional or conservative nature of a vector field. Where the vector field corresponds
to a force, we have a neat physical motivation for the name: a force is conservative
if the work done in going around a closed path is zero, and if a particle moves
solely under the influence of that force, then the energy is conserved.
F (r) = −∇(qφ(r)).
We have seen that the (conservative) force acting on a particle is minus the gra-
dient of its potential energy: F (r) = −∇U (r).
Chapter 1 25
For the case where φ vanishes at infinity, the potential U (r) is the work done, W ,
in bringing the charge q from infinity to the point r. We will now consider the
work done in assembling a set of point charges qi at ri , i = 1, . . . , N .
We do this by bringing each charge i in turn, one at a time, to position ri , and
then fixing it in position. The work done in bringing charge i is
qi i−1
X qj
Wi =
4πǫ0 j=1 |ri − rj |
and thus the total work done in assembling the charges is
1 X N i−1
X qi qj
W = = U,
4πǫ0 i=2 j=1 |ri − rj |
where U is the potential energy of the system. We can write this in a more
symmetric form as
1 X N X N qi qj
U=
8πǫ0 i=1 j=1 |ri − rj |
where we do not include the self-energy term, i = j.
We can generalise this to a continuous charge distribution in the usual way, viz
1 Z 3 3 ′ ρ(r)ρ(r′ )
U= d rd r ,
8πǫ0 |r − r′ |
and we now use eqn. 1.5 to write
1Z
U= ρ(r)φ(r)dV, (1.8)
2
analogous to eqn. 1.7.
Chapter 1 26
We can also interpret the potential energy in terms of the electric field, by using
ME1
ǫ0 Z
U = dV ∇ · E(r)φ(r)
2 Z
ǫ0
= − dV E(r) · ∇φ(r) (Integration by parts)
2Z
ǫ0
= dV |E|2 . (1.9)
2
We now identify the integrand as the energy density
ǫ0
u(r) = |E(r)|2 .
2
Chapter 1 27
We are now ready to derive a differential equation for the potential. Our starting
point is Maxwell’s First Equation (ME1), derived earlier:
ρ
∇ · E = 4πkρ = .
ǫ0
We now make use of the irrotational nature of E(r) to write E = −∇φ(r). Thus
ME1 becomes
• This equation is Poisson’s Equation. ρ(r) is the source for the electro-
static potential φ(r).
• If we have that the source ρ(r) ≡ 0 everywhere, then this equation becomes
∇2φ = 0.
These are two of the most important equations in physics. They, or close variants,
occur in:
• Electromagnetism, as above
Laplace’s and Poisson’s equations are linear, second order, partial differential
equations; to determine a solution we have also to specify boundary conditions.
Example: One-dimensional problem
d2φ(x)
=λ
dx2
for x ∈ [0, L], where λ is a constant. This has solution
1
φ(x) = λx2 + Ax + B
2
where A, B are constants. To determine these constants, we might specify the
values of φ(x = 0) and φ(x = L), i.e. the values on the boundary.
normal
n. φ
∆
We will proceed to show that the solutions of Laplace’s and Poisson’s are unique,
up to a constant (Neumann), if subject to either of the above boundary conditions.
We begin with a couple of useful vector identities
We begin with a couple of identities that will be useful both in this proof and
later.
Let ψ1 and ψ2 be two continuously differentiable, arbitrary scalar fields defined
in a volume V bounded by a closed surface S. Introduce the vector field A(r) =
ψ1∇ψ2 . From the divergence theorem, we have
Z Z
∇ · A dV = A · n dS
V S
We now proceed to the formal proof. Let φ1 (r) and φ2 (r) be solutions of Poisson’s
equation ∇2φi = −ρ/ǫ0 inside a volume V bounded by surface S, satisfying either:
where f (r) and g(r) are continuous functions defined on the surface S.
Consider the function
ψ(r) = φ1 (r) − φ2 (r).
Then ψ satisfies Laplace’s equation:
∇2ψ(r) = 0 in V
with either
We now apply Green’s first identity for the case ψ1 = ψ2 = ψ, and obtain
Z Z
2
|∇ψ| dV = (ψ∇2ψ + |∇ψ|2 ) dV (since ∇2ψ = 0 in V )
V ZV
= ψ∇ψ · n dS (from eqn. 2.5)
S
= 0 (1.12)
Chapter 1 31
• The uniqueness property means we can use any method we wish to obtain
the solution - if it satisfies the correct boundary conditions, and is a solution
of the equation, then it is the correct solution. A good example: Method of
Images, to be covered in the next chapter.
Chapter 1 32
Within a conductor, the electrostatic field E must be zero. However, the field is
zero because of an induced charge density sufficient to annul the external field.
Now ME1 tells us that ∇ · E = ρ/ǫ0 , where ρ is the charge density. So if E is
zero within the conductor, the charge density must be zero. So where does the
induced charge density reside?
The charge density is confined solely to the surface of the conductor
Chapter 1 34
E · n = σ/ǫ0
Note: the surface charge density discussed here is different to a sheet of charge
of density σ per unit area discussed earlier in the course. The latter may best be
viewed as a charge distribution in an insulator, i.e. a fixed charge distribution.
Unfortunately, the two terms are often confused in the literature, and indeed
probably in these lectures!
~ = P ~q.
φ
Boundary-Value Problems in
Electrostatics
Φ = 0 on surface
Q
a
O
1
Chapter 2 2
Once again, the method was particularly simple in this case because of spherical
symmetry. Similar simplifications occur in the case of cylindrical symmetry.
φ=0
Chapter 2 4
Thus, by our uniqueness theorem, the potential in the upper half plane is the
same as that of a charge q placed above an grounded sheet at z = 0.
Chapter 2 5
a2 n 2
= 2 a − 2ab cos θ + b2
o
b
a2
= 2 |r − b|
b
and hence
1 q qa 1
φ(r)|r=a = − = 0.
4πǫ0 |r − b| b a/b|r − b|
Thus we have
1. The image system satisfies the original Poisson’s equation for r ≥ a since the
only additional charge we have introduced is in the region r < a.
with b′ = a2 /b.
Using
|a − a2 /b2b|r=a = a/b|a − b|r=a
we find
q a n
2 2
o
σ=− b /a − 1 a.
4π |a − b|3
Note that the surface charge density is not uniform, but that
Z
σ dS = q ′
S
as expected.
2.1.3 Point charge near insulated, conducting sphere with total charge
Q
This problem is a slightly more complicated. Our starting point is the point charge
near the grounded conducting sphere, together with the superposition principle.
2. Disconnect the sphere from earth, and add a charge Q−q ′ to the sphere. This
charge will be uniformly distributed, since the charge q ′ is already distributed
to balance the forces due to q.
Chapter 2 8
Appealing to the uniqueness theorem, and noting that, once again, no charges
have been introduced in r ≥ a, we have
Q − q′ q′
1 q
φ(r) = + + .
4πǫ0 r 4πǫ0 |r − b| r − a2 /b2b|
We will now proceed to calculate the Force on the charge q; this is just given by
Coulomb’s law for the forces between q and the two image charges:
1 Q − q′ q′
′
F = q 3 b+ (b − b )
4πǫ0 b |b − b′ |3
qa3 (2b2 − a2 )
1 qb
= Q−
4πǫ0 b3 b(b2 − a2 )2
Note that the force is always attractive at sufficiently small distances irrespec-
tive of Q due to the induced surface charge density on the conductor.
Chapter 2 9
We will now proceed to a formal solution using Green functions. First, however,
a mathematical digression. . .
δ(x − a) = 0 if x 6= a.
2.
Z
1 if a ∈ R
dx δ(x − a) =
R 0 otherwise
The delta function is not strictly a function but rather a distribution; it is de-
fined purely through its effect under an integral. It immediately follows from the
definition that Z
dx f (x)δ(x − a) = f (a) (2.2)
if a lies within the region of integration.
The δ-function δ(x − a) may be thought of as the limit of a Gaussian centred at
a in which the width tends to zero whilst the area under the Gaussian remains
unity.
δ(x − a) = lim δǫ (x − a)
ǫ→0
1 − (x−a)2
δǫ (x − a) = √ e ǫ (2.3)
πǫ
R∞
It is easy to see that limǫ→0 δǫ (x) = 0 if x 6= a and −∞ δǫ (x − a) = 1. Let us check
the property (2.2)
Z ∞
lim δǫ(x − a) f (x)
ǫ→0 −∞
Chapter 2 10
1
Z ∞ (x−a)2 1
= lim √ e− ǫ [f (a) + (x − a)f ′(a) + (x − a)2 f ”(a) + ...]
ǫ→0 −∞ πǫ 2
2 4
= lim [f (a) + ǫ f ”(a) + O(ǫ )] = f (a)
ǫ→0
There are some simple relations that follow from the Eq. (2.2)
1. The δ-function is a derivative of a step function θ(x):
1 x≥0
θ(x) = (2.4)
0 x<0
Indeed, if f (x) vanishes at infinity
Z ∞ Z ∞ Z ∞
′ ′
f (x) θ (x) = − f (x) θ(x) = − f ′ (x) = f (0)
−∞ −∞ 0
Z ∞
f (x) δ(x) = f (0)
−∞
2.
Z Z
′
dx f (x)δ (x − a) = − dx f ′(x)δ(x − a) integ. by parts
= −f ′ (a)
3.
Z XZ 1
dx f (x)δ(g(x)) = dy f (xi(y))δ(y)
i g ′ (xi(y))
X f (xi)
= ′
i |g (xi )|
so that
Z
1 if X ∈ V
d3 x δ(x − X) =
V 0 otherwise
Chapter 2 11
Note that it is this last property that defines the multi-dimensional δ-function,
with this simple representation in a Cartesian basis; you have to be a little
careful when working in curvilinear coordinates.
As a simple illustration of the power of the δ-function, let us return to the expres-
sion, eqn. (1.5), for the potential due to a continuous charge distribution
1 Z 3 ′ ρ(x′ )
φ(x) = dx .
4πǫ0 V |x − x′ |
We now introduce the δ-function to enable us to write a set of N discrete charges
qi at xi as a charge distribution
i
so that
1 Z 3 ′ i qiδ (3) (x′ − xi )
P
φ(x) = dx
4πǫ0 V |x − x′ |
1 X qi
=
4πǫ0 i |x − xi |
which is our familiar expression for the potential due to a set of point charges.
1
∇2 ′
= −4πδ (3) (x − x′ )
|x − x |
Chapter 2 12
where the “primed” denotes differentiation with respect to the primed indices.
Let us apply this for the case ψ1 (x′ ) = |x−1x′ | and ψ2 (x′ ) = φ(x′ ) where
and
∇′2ψ1(x′ ) = −4πδ (3) (x − x′ )
yielding
−ρ(x′ )
Z
3 ′
1 ′ (3) ′
dx + φ(x )4πδ (x − x ) =
|x − x′ | ǫ0
Z 1 ′ ′ 1
dS ′ n · ′ ′
∇ φ(x ) − φ(x )∇ .
|x − x′ | |x − x′ |
Applying our rule for integrating over δ-functions, we obtain
1 Z 3 ′ ρ(x′ )
φ(x) = dx +
4πǫ0 |x − x′ |
′
1 Z 1 ∂φ(x ) ∂ 1
dS ′ ′
− φ(x ) . (2.5)
4π |x − x′ | ∂n′ ∂n′ |x − x′ |
The function 1/|x − x′ | is said to be a Green function for the problem.
The Green function is not unique, and is just a function satisfying
1 Z 3 ′
φ(x) = d x G(x, x′ )ρ(x′ ) +
4πǫ0 V
′ ′
1 Z
∂φ(x ) ∂G(x, x )
dS ′ G(x, x′ ) − φ(x′ )
′
(2.6)
4π S=∂V ∂n ∂n′
The utility of this generalisation is the following. In eqn. 2.5, the surface integral
involved both φ(x′ ), and ∂φ(x′ )/∂n′; in general we cannot specify both simultane-
ously at a point on the surface, since the problem is then overdetermined. Thus in
eqn. 2.5 we have an implicit equation for φ(x), with the unknown also appearing
under the integral on the right-hand side. In eqn. 11.79, we can choose G(x, x′ )
so that the surface integral depends only on the proscribed boundary values of φ
(Dirichlet) or ∂φ/∂n′ (Neumann).
We will now consider the boundary conditions we have to impose on the Green
Functions to accomplish the above aim.
Dirchlet Problem
Here the value of φ(x′ ) is specified on the surface, and therefore it is natural to
impose that the Green function GD (x, x′ ) satisfy
GD (x, x′ ) = 0 for x′ on S,
and thus
′
1 Z 3 ′ ′ ′ 1 Z ′ ′ ∂GD (x, x )
φ(x) = d x GD (x, x )ρ(x ) − dS φ(x ) . (2.7)
4πǫ0 V 4π S ∂n′
Thus the surface integral only involves φ(x′ ), and not the unknown ∂φ(x′ )/∂n′.
Chapter 2 14
Neumann Problem
where the final term is just the average value of φ(x′ ) on the surface S. The
inclusion of this term is perhaps not surprising; recall that the solution to the
Neumann problem is unique only up to an additive constant.
Proof
Apply Green’s theorem for the case ψ1 (x′ ) = GD (x, x′ ), and ψ2 (x′ ) = GD (y, x′ ):
Z
d3x′ (GD (x, x′ )∇′2GD (y, x′ ) − GD (y, x′ )∇′2GD (x, x′ )) =
V Z
dS ′ n · (GD (x, x′ )∇′ GD (y, x′ ) − GD (y, x′ )∇′ GD (x, x′ )).
S
Chapter 2 15
But for the Dirichlet problem GD (x, x′ ) vanishes for all x′ ∈ S, and hence the
right-hand side of the above is zero. Thus we have
Z
d3 x′ GD (x, x′ ){−4πδ (3) (y − x′ )} − GD (y, x′ ){−4πδ (3) (x − x′ )} = 0
n o
and hence
GD (x, y) = GD (y, x)
Chapter 2 16
The secret, then, to the solution of boundary value problems is determining the
correct Green function, or equivalently obtaining the function F (x, x′ ). They are
several techniques
1. Make a guess at the form of F (x, x′ ). Here we recall that F is just the
solution of the homogeneous Laplace’s equation ∇′2F (x, x′ ) = 0 inside V , and
therefore is just the solution of the potential for a system of charges external
to V . In particular, for the Dirichlet problem, since GD (x, x′ ) vanishes at
x′ ∈ S, we have that F (x, x′ ) is just that system of charges external to V
that, when combined with a point charge at x, assures that the potential
vanishes on the surface. And finding that system of charges is precisely what
we were doing in the Method of Images. . .
We saw at the beginning of this chapter how to use the method of images to con-
struct the potential φ(x′ ) for a point charge at x outside an grounded conducting
sphere of radius a. In particular, for a charge q = 4πǫ0 , the potential satisfies
with φ(x′ ) = 0 for x′ on S. Thus now see that φ(x′ ) is precisely the Green func-
tion GD (x, x′ ) that we need. Note that you have to be careful to distinguish the
variable we are integrating over, x′ , and the variable at which we are evaluating
the potential, x. Perhaps counter-intuitively, it is at the point x that we place
our point charge.
Chapter 2 17
z x
x’
θ From eqn. 2.1, we have that the Green function
γ is
11111111111111111111
00000000000000000000 1 a
00000000000000000000
11111111111111111111
00000000000000000000
11111111111111111111 G(x, x′ ) = − ,
00000000000000000000
11111111111111111111 |x′ − x| x|x − a2 /x2x|
′
00000000000000000000
11111111111111111111
and it is easy to check that, indeed, G(x, x′ ) =
G(x′ , x).
ϕ
We can rewrite this as
1 1
G(x, x′ ) = 2
−
(x + x′2 − 2xx′ cos γ)1/2 x2x′2/a2 + a2 − 2xx′ cos γ)1/2
where γ is the angle between x and x′ .
The general solution for the potential is then
1 Z 3 ′ 1 Z ∂G(x, x′ )
φ(x) = d x G(x, x′ )ρ(x′ ) − dS ′ φ(x′) . (2.9)
4πǫ0 V 4π S ∂n′
Thus we need the normal gradient of the Green function to the surface, which
points inward,
∂G ∂G
= −
∂n′ surface ∂x ′ ′
x =a
2x2a/a2 − 2x cos γ
1 −2a − 2x cos γ
= − 2 +
2 (x + a2 − 2ax cos γ)3/2 (x2a2 /a4 + a2 − 2ax cos γ)3/2
x2 − a2
= −
a(x2 + a2 − 2ax cos γ)
Thus we have all the ingredients to solve the Dirichlet problem outside a sphere
of radius a.
Chapter 2 18
Because the source is zero, we only need the surface term from eqn. 2.9
′
1 Z ′ ′ ∂G(x, x )
φ(x) = − dS φ(x ) .
4π S ∂n′
Now dS ′ = a2 dψd(cos θ′), yielding
1 2 Z 2π ′ ′ ∂G ∂G
(
Z 1 Z 0 )
φ(x) = − a dψ V d(cos θ ) ′ + (−V ) d(cos θ′ ) ′
4π 0 0 ∂n −1 ∂n
2 2
V Z 2π 1
Z
a(x − a )
= dψ ′ d(cos θ′ ) 2 −
4π 0 0 (a + x2 − 2ax cos γ)3/2
2 2
Z 0 a(x − a )
d(cos θ′) 2
.
−1 (a + x2 − 2ax cos γ)3/2
We can express cos γ in terms of the spherical polar coordinates of x and x′ by
noting that
cos γ = n · n′ = (sin θ cos ψ, sin θ sin ψ, cos θ) · (sin θ′ cos ψ ′ , sin θ′ sin ψ ′ , cos θ′ )
= sin θ sin θ′ cos(ψ − ψ ′ ) + cos θ cos θ′ ,
where n and n′ are unit vectors in the directions of x and x′ respectively. Finally,
we can combine the two integrals through by making the substitution θ′ → π − θ′
and ψ ′ → ψ ′ + π in the second integral, giving
V 2 2
Z 2π
′
Z 1 1
d(cos θ′ ) 2
φ(x) = a(x − a ) dψ
4π 0 0 (a + x2 − 2ax cos γ)3/2
1
− 2
(a + x2 + 2ax cos γ)3/2
In general, we cannot obtain the solution in closed form; γ is just too complicated
a function of θ′ and ψ ′ . However, we can study the solution in specific cases.
Chapter 2 19
We can also obtain the solution for x >> a, by means of a Taylor expansion. We
begin by writing
where
ax
α= ,
a2 + x2
yielding
V a(x2 − a2 ) Z 2π ′ Z 1
′
1 1
φ(x) = dψ d(cos θ ) − .
4π (a2 + x2)3/2 0 0 (1 − 2α cos γ)3/2 (1 + 2α cos γ)3/2
We now expand the integrand as a power series in α, yielding
The integrals for the first two terms in the expansion are perfectly tractable.
Recalling that cos γ = sin θ sin θ′ cos(ψ − ψ ′ ) + cos θ cos θ′ , we find
Chapter 2 20
1.
Z 2π Z 1 Z 2π Z 1
′ ′ ′
dψ d(cos θ ) cos γ = dψ d(cos θ′) cos θ cos θ′ = π cos θ
0 0 0 0
2.
Z 2π Z 1
′
dψ d(cos θ′ ) cos3 γ = π/4 cos θ(3 − cos2 θ)
0 0
and thus
3V a2 x(x2 − a2 ) 35 a2 x2
2 4 4
φ(x) = cos θ 1 + (3 − cos θ) + O(a /x ) .
2(a2 + x2 )5/2 24 (a2 + x2)2
Note that we can express this power series as a series in a2 /x2, rather than α,
yielding
3V a2 7a2 5
3
!
3 4 4
φ(x, θ, ψ) = cos θ − cos θ − cos θ + O(a /x ) .
x2 12x2 2 2
and we can verify that this gives the correct expression for θ = 0.
As we go to higher order terms in the expansion, the angular integrals become
increasingly intractable, and this approach fails. However, the eagle-eyed amongst
you may recognise the angular terms as the Legendre polynomials P1 (cos θ) and
P3 (cos θ), and this brings us to the next section.
where
2 Z a/2 2πmx
!
Am = dx f (x) cos m = 0, 1, 2, . . .
a −a/2 a !
2 Z a/2 2πmx
Bm = dx f (x) sin m = 1, 2, . . .
a −a/2 a
We can combine the sine and cosine terms by noting
1 h ix
e + e−ix
i
cos x =
2
1 h ix
e − e−ix ,
i
sin x =
2i
and introducing a new set of functions
1
Um (x) = √ ei2πmx/a m = 0, ±1, ±2, . . . ,
a
We get an expansion
∞
X
f (x) = Am Um (x),
m=−∞
where
1 Z a/2 ′
Am = √ dx′f (x′)e−2πimx /a .
a −a/2
Proof of completeness
∞ ′
ein(x−x ) = 2πδ(x − x′ )
X
−∞
for x, x′ ∈ [−π, π]:
Chapter 2 23
⇒ ∞ −∞ e
inx
= 2πδ(x), Q.E.D.
P
Taking the real part of both sides of this equation we reproduce Eq. (2.12).
Suppose we now let a → ∞, so that the discrete sum over m becomes an integral
over a continuous variable k where
2πm
→ k.
a
Then we have
X a Z
→ dk
m 2π
and the discrete coefficients become a continuous function
v
u 2π
u
Am → t
A(k).
a
Thus we may express the Fourier Transforms as
1 Z
f (x) = √ dk A(k)eikx
2π
1 Z
A(k) = √ dx f (x)e−ikx.
2π
Note that the assignment of the coefficients outside the integrals depends on the
convention adopted; in all cases the product is 1/2π.
The orthogonality and completeness relations assume the continuous, and sym-
metric, forms
1 Z∞ ′
dx ei(k−k )x = δ(k − k ′ )
2π −∞
1 Z ′
dk eik(x−x ) = δ(x − x′ )
2π
How does one obtain a complete set of orthonormal functions? We will now show
that, for a certain class of differential equations, the solutions are orthogonal, for
specific boundary conditions.
Chapter 2 25
2.5.4 Theorem
Proof
providing
!#b
dψλ′ dψλ∗
"
p(x) ψλ∗ − ψλ′ = 0.
dx dx a
Corollaries
and hence λ∗ = λ.
2. For λ′ 6= λ, Z b
dxr(x)ψλ∗ ψλ′ = 0,
a
i.e. the functions ψλ are orthogonal.
Chapter 2 27
We will now see how the Sturm-Liouville equation arises in the solution of Laplace’s
equation, and how we can then use the Sturm-Liouville theorem to provide an
orthonormal set of functions. The method we will use will be the separation of
variables. It is best shown by illustration.
Consider the solution of Laplace’s equation in a box 0 ≤ x ≤ a, 0 ≤ y ≤ b,
0 ≤ z ≤ c, with the values of the potential prescribed on the boundary. In
particular, let us consider the case where φ vanishes on the boundary, except on
the plane z = c where φ(x, y, z = c) = V (x, y).
In Cartesian coordinates, the natural coordinate system for the problem, Laplace’s
equation assumes the form
∂2 ∂2 ∂2
φ(x, y, z) + 2 φ(x, y, z) + 2 φ(x, y, z) = 0.
∂x2 ∂y ∂z
We will seek solutions to this equation that are factorisable, i.e.
and build up our final solution from such factorisable solutions. Substituting this
form into Laplace’s equation, we obtain
d2X(x) d2Y (y) d2Z(z)
Y (y)Z(z) + X(x) X(z) + X(x)Y (y) = 0,
dx2 dy 2 dz 2
which we may write as
1 d2 X 1 d2 Y 1 d2 Z
+ + = 0.
X dx2 Y dy 2 Z dz 2
We have separated the equation into three terms, each dependent on a different
variable. Since the equation holds for all x, y, z, we can say that each term must
separately be constant. Thus
1 ′′
X = C1 (2.21)
X
Chapter 2 28
1 ′′
Y = C2 (2.22)
Y
1 ′′
Z = C3 (2.23)
Z
where C1 + C2 + C3 = 0.
Let us consider eqn. 2.21
d2X(x)
− C1X = 0,
dx2
and choose a trial solution
X(x) = eαx .
Then we have that α2 = C1 .
Xn (x) = sin αn x.
hence we immediately know that the functions Xn (x) are orthogonal. We can
treat Y (y) similarly, and obtain
mπ
Ym (y) = sin βm y; βm = , m = 1, 2, . . .
b
Finally, we obtain Z from
Z ′′ 2 2 n2π 2 m2 π 2
= αn + βm = 2 + 2 > 0.
Z a b
In this case, the solution is a real exponential, and imposing the boundary condi-
tion Z(0) = 0 we have
Z(z) = sinh γnmz
where
q
γnm = π n2/a2 + m2 /b2.
Thus the general solution, using the completeness property, is
∞
X
φ(x, y, z) = Anm sin αn x sin βm y sinh γnmz.
m,n=1
We obtain the coefficients Amn by imposing the boundary conditions on the plane
z = c:
∞
X
V (x, y) = Anm sin αn x sin βm y sinh γnmc.
m,n=1
4 Z a Z b
Anm = dx dyV (x, y) sin αn x sin βm y
ab sinh γnmc 0 0
φ(0, y) = φ(a, y) = 0
φ(x, 0) = V
φ(x, y) → 0 as y → ∞
X ′′ + α2 X = 0
Chapter 2 31
Yn′′ − αn2 Yn = 0
Yn (y) = e−αn y .
and thus
4V /nπ n odd
An =
0 n even
with
4V X 1 −nπy/a nπx
φ(x, y) = e sin .
π n odd n a
yielding
nπx
sin = ℑeinπx/a .
a
Thus we may write the general solution as
4V X 1 −nπy/a inπx/a
φ(x, y) = e ℑe
π n odd n
4V X 1 (nπi/a)(x+iy)
= ℑe .
π n odd n
We now introduce the variable
Z = e(iπ/a)(x+iy) ,
and thus
X 1 n 1
Z = − {ln(1 − Z) − ln(1 + Z)}
n odd n 2
1 1+Z
= ln .
2 1−Z
Hence we may write the general solution as
2V 1+Z
φ(x, y) = ℑ ln .
π 1−Z
We will conclude by writing this solution explicitly in terms of x and y. We begin
by noting that Z = |Z| exp iθ where θ is the phase of Z, i.e. tan θ = ℑZ/ℜZ.
Thus
ln Z = ln |Z| + iθ =⇒ ℑ ln Z = θ.
Now,
1+Z (1 + Z)(1 − Z ∗) 1 − |Z|2 + 2iℑZ
= =
1−Z |1 − Z|2 |1 − Z|2
and thus
1+Z 2ℑZ
ℑ ln = tan−1 .
1−Z 1 − |Z|2
But we have
πx
ℑZ = e−πy/a sin
a
|Z|2 = e−2πy/a
and thus
−πy/a
sin πx
2V −1 2e a
φ(x, y) = tan −2πy/a
π 1−e
which, after some simplification, becomes
2V sin πx/a
φ(x, y) = tan−1 .
π sinh πy/a
Chapter 2 34
In practice, such problems can be done in a much simpler way, by observing that
the real and imaginary components, u and v respectively, of an analytic complex
function f (z = x + iy) satisfy the two-dimensional Laplace’s equation
∂ 2u ∂ 2u ∂ 2v ∂ 2v
+ = 0; + = 0.
∂x2 ∂y 2 ∂x2 ∂y 2
This is a direct consequence of the Cauchy-Riemann eqnations.
Each term depends on a different variable, and this must hold for all s and z.
Thus each term is separably constant. For the function T (θ), let us take
1 ∂ 2T
2
= −ν 2.
T ∂θ
Since T must attain the same value at θ = 0 and θ = β, the solution must be
oscillatory rather than exponential, and hence ν 2 must be positive. Thus the
solution is
Aν cos νθ + Bν sin νθ; ν 6= 0
Tν (θ) =
A0 + B0 θ; ν=0
For the radial function, we have
∂ ∂R
!
s s − ν 2R = 0.
∂s ∂s
For ν 6= 0, let us take as trial solution R ∼ sα ,
(α2 − ν 2)sα = 0,
The solution must be valid as s → 0 (note that we are not interested in the
solution for s large), and therefore the terms proportional to ln s and s−ν cannot
contribute. Thus our solution is of the form
A0 + B0 θ; ν=0
φ(s, θ) =
sν (Aν cos νθ + Bν sin νθ); ν > 0
We will now use the boundary conditions on the planes to further constrain the
solution. At θ = 0, β, we have φ = V , independent of s, and therefore we have
Aν = 0
nπ
sin νβ = 0; ν= , n = 1, 2, . . . ,
β
yielding
nπθ
Bn snπ/β sin
X
φ(s, θ) = A0 + B0θ + .
n β
Finally, we impose that the potential be V on the two planes
θ = 0, φ = V =⇒ A0 = V
θ = β, φ = V =⇒ B0 = 0,
nπθ
Bn snπ/β sin
X
φ(s, θ) = V + .
n β
As we get closer into the corner, the first term will dominate,
πθ
φ(s, θ) ∼ V + B1 sπ/β sin .
β
Taking the gradient, we obtain
πB1 π/β−1 πθ πB1 π/β−1 πθ
E = −∇φ = − s sin es − s cos eθ
β β β β
and the induced surface charge density is
πB1 ǫ0 π/β−1
σ = ǫ0 [E · n] = − s
β
Chapter 2 37
Thus we see behaviour familiar from our knowledge of “action at points” - the
fields and surface charge densities become singular near sharp edges.
Chapter 3
In the previous chapter, we saw how we could look for factorizable solutions to
Laplace’s Equation in Cartesian coordinates, and then construct the solution for
more general boundary values using the completeness property of the such fac-
torisable solutions. In this chapter we will employ analogous methods in spherical
polar and cylindrical coordinate systems. In practice, the coordinate system that
is appropriate depends on the symmetry or geometry of the problem.
We will denote our coordinates by (r, θ, ϕ), in terms of which Laplace’s equation
assumes the form
1 ∂ 2 ∂φ 1 ∂ ∂φ 1 ∂ 2φ
! !
2
∇ φ(s, θ, ϕ) = 2 r + 2 sin θ + 2 2 .
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2
We will now seek factorisable solutions of the form
U (r)
φ(r, θ, ϕ) = P (θ)Q(ϕ),
r
1
Chapter 2 2
where the factor of 1/r is conventional. Substituting this into Laplace’s equation,
we have
1 d 1 1 dU (r)
P (θ)Q(ϕ) 2 r2 − 2 U (r) +
r dr r r dr
d2Q(ϕ)
U (r)Q(ϕ) 1 d dP (θ) U (r)P (θ) 1
+ sin θ + = 0,
r r2 sin θ dθ dθ r r2 sin2 θ dϕ2
yielding
P Q d2 U UQ d dP U P d2 Q
!
+ 3 sin θ + 3 2 = 0,
r dr2 r sin θ dθ dθ r sin θ dϕ2
which we may write as
1 d2 Q 2
!
2 2 1 d U 1 1 d dP
+ r sin θ + sin θ = 0. (3.1)
Q dϕ2 U dr2 r2 sin θ P dθ dθ
The first term is a function of ϕ alone, and the remaining term is a function of
(r, θ) alone. Thus they must be separately constant, and we may write
1 d2 Q
2
= −m2 , (3.2)
Q dϕ
where m is a constant. Eqn. 3.2 has solution
Q = e±imϕ .
We now observe that the solution must be periodic, with period 2π, in the az-
imuthal variable ϕ. Thus m must be an integer, and, of course, real. Thus we
may write eqn. 3.1 as
r 2 d2 U 1 1 d dP m2
!
+ sin θ − = 0.
U dr2 sin θ P dθ dθ sin2 θ
We now observe that the first term is purely a function of r, whilst the remaining
terms are purely a function of θ. Thus we may write
r 2 d2 U
= l(l + 1)
U dr2
Chapter 2 3
where l is a constant - we will see the reason for expressing the constant in this
way later. To solve this equation, we will take a trial solution
U (r) = rα ,
yielding
α(α − 1) = l(l + 1)
with solutions α = l + 1, −l. Thus we have
The equation for the polar coordinate θ now assumes the form
m2
1 d dP
!
sin θ + l(l + 1) −
P = 0.
sin θ dθ dθ sin2 θ
It is convenient to introduce the variable x = cos θ, with −1 ≤ x ≤ 1 and
d d
= − sin θ .
dθ dx
After a little algebra, we have
m2
d 2 dP
" #
(1 − x ) + l(l + 1) −
P =0
dx dx 1 − x2
This is the Generalised Legendre Equation, and is, once again, an equation
of Sturm-Liouville type, with p(x) = 1 − x2, q(x) = −m2 /(1 − x2), λ = l(l + 1),
and r(x) = 1.
We will now seek solutions of this equation, first for the case m = 0, where the
equation is known as the Ordinary Legendre Equation
d dP
" #
(1 − x2 ) + l(l + 1)P = 0.
dx dx
Chapter 2 4
We begin by noting that the solutions must be both continuous and single-
valued in the region −1 ≤ x ≤ 1, corresponding to 0 ≤ θ ≤ π. We will obtain
the solutions through series substitution, i.e. by trying a solution of the form
∞
cn xγ+n,
X
P =
n=0
from which
dP ∞
cn (γ + n)xγ+n−1
X
=
dx n=0
dP ∞ ∞
(1 − x2) γ+n−1
cn (γ + n)xγ+n+1,
X X
= cn (γ + n)x −
dx n=0 n=0
d ∞ ∞
2 dP
" #
γ+n−2
cn (γ + n)(γ + n + 1)xγ+n.
X X
(1 − x ) = cn (γ + n)(γ + n − 1)x −
dx dx n=0 n=0
As this equation must be valid ∀x ∈ [−1, 1], we can equate the coefficients of the
powers of x to zero. The leading power of x is xγ−2, and we use this equation, the
indicial equation, to determine γ. Thus
• xγ−2:
c0 γ(γ − 1) = 0 =⇒ γ = 0 or γ = 1
• xγ−1:
c1 undetermined : γ = 0
c1 (γ + 1)(γ + 1 − 1) = 0 =⇒
c1 = 0 :γ=1
• xγ+n, n ≥ 0:
(γ + n)(γ + n + 1) − l(l + 1)
cn+2 = cn .
(γ + n + 2)(γ + n + 1)
Chapter 2 5
Note that the recursion relation relates only even (odd) polynomials for γ = 0
(γ = 1).
We have already noted that the solution must be valid for x ∈ [−1, 1], and in
particular at the end points x = ±1. Thus the series must be finite at x = ±1.
To explore the convergence properties, we note that
cn+2/cn −→ 1 as n −→ ∞,
and thus the series resembles a geometrical expansion x2n. This diverges at
P
x = ±1 unless the series terminates, i.e. unless cn = 0 for some n. Thus our
requirement for convergence is
(γ + n)(γ + n − 1) − l(l + 1) = 0 for some n.
• γ = 0:
n(n + 1) = l(l + 1) =⇒ n = l
• γ = 1:
(n + 1)(n + 2) = l(l + 1) =⇒ n = l − 1.
Note that in both cases the highest power of x is xl ; the two cases are the same.
We call the corresponding solutions Pl (x) the Legendre Polynomials, and con-
ventially we take Pl (1) = 1. The first few are
P0 (x) = 1
P1 (x) = x
1
P2 (x) = (3x2 − 1)
2
1
P3 (x) = (5x3 − 3x).
2
We can write the Legendre polynomials in a more memorable form through Ro-
drigues’ Formula:
1 dl 2
Pl (x) = l l
(x − 1)l .
2 l! dx
Chapter 2 6
Applying our theorem concerning the orthogonality of the solutions of the Sturm-
Liouville equation yields
Z 1 Z 1
′ ′
[l(l − 1) − l (l + 1)] dx Pl (x)Pl′ (x) = 0 =⇒ dx Pl (x)Pl′ (x) = 0, l 6= l′ ,
−1 −1
i.e. the Legendre polynomials are orthogonal. N.B. it is easy to check that our
solutions satisfy the required boundary conditions.
To determine their normalisation, we can use either Rodrigues’ formula, or the
generating function; we use the latter. From eqn. 3.3, we have
Z 1
2
Z 1 1
dx g(t, x) = dx
−1 −1 1 − 2xt + t2
)1
1
(
= − ln(1 − 2xt + t2)
2t −1
2
1 (1 − t)
= − ln
2t (1 + t)2
X∞ 2t2l
= ,
l=0 2l + 1
where we have used the series expansion of ln(1 + t). However, we also have
Z 1 ∞ Z ′
2
dx Pl (x)Pl′ (x)tl+l
X
dx g(t, x) =
−1
l,l′ =0
∞
2l 1
Z
dx Pl (x)2.
X
= t
l=0 −1
Z 1 2
dx Pl (x)Pl (x) =
−1 2l + 1
Chapter 2 7
3.1.4 Completeness
Since the Legendre Polynomials form a complete set, we may write any function
f (x), x ∈ [−1, 1] as
∞
X
f (x) = Al Pl (x).
l=0
We obtain the coefficients using the orthogonality relations
Z ∞
X Z 1
dx f (x)Pl(x) = Al dx Pl (x)Pl′ (x)
−1
l′ =0
2
= Al
2l + 1
whence
2l + 1 Z 1
Al = dx f (x)Pl (x)
2 −1
Example
where we have used the normalisation condition Pl (1) = 1. But we have (from
Rodrigues’s formula, with a little work)
l/2
(−1)l/2 (l−1)!!
: l even
2 (l/2)!
Pl (0) =
0
: l odd
where (l − 1)!! = (l − 1)(l − 3) . . . 3.1. Thus
(−1)(l+1)/2l!! (−1)(l−1)/2(l − 2)!!
Al = − +
2(l+1)/2((l + 1)/2)! 2(l−1)/2((l − 1)/2)!
(−1)(l−1)/2(l − 2)!! l
( )
= (l−1)/2 1+
2 ((l − 1)/2)! l+1
Thus (l−1)/2
(l−2)!!(2l+1)
−1 : l odd
2 2( l+1
2 )!
Al =
0 : l even
and we have
3 7 11
f (x) = P1 (x) − P3 (x) + P5 (x) . . .
2 8 16
Chapter 2 9
We may now write our general solution for the boundary-value problem in spher-
ical coordinates with azimuthal symmetry, i.e. no ϕ dependence, as
∞
Al rl + Bl r−l−1 Pl (cos θ),
X
φ(r, θ) =
l=0
where the coefficients Al and Bl are determined from the boundary conditions.
Example:
Consider the case of a sphere, of radius a, with no charge inside but potential
V (θ) specified on the surface.
Since there are no charges inside the sphere, the potential φ inside must be regular
everywhere. Thus Bl = 0 ∀l, and we may write the solution as
∞
Al rl Pl (cos θ).
X
φ(r, θ) =
l=0
with
∞
Bl a−l−1Pl (cos θ),
X
V (θ) =
l=0
Chapter 2 10
so that
2l + 1 l+1 Z π
Bl = a dθ sin θ V (θ)Pl (cos θ).
2 0
Then we have
2l + 1 l+1
(Z )
π/2 Z π
Bl = a V Pl (cos θ) sin θdθ − Pl (cos θ) sin θdθ
2 0 −π/2
2l + 1 l+1
(Z )
1 Z 0
= a V dxPl (x) − dxPl (x)
2 0 −1
2l + 1 l+1
(Z )
1
= a V dxf (x)Pl(x)
2 −1
where
1 0<x≤1
f (x) =
−1 −1 ≤ x < 0
This is just the expression we evaluated in Section 3.1.4, and thus we have:
l−1 (l−2)!!(2l+1)
V al+1(− 12 ) 2
2( l+1
l odd
2 )!
Bl =
0 l even
so that
3 a2 7 a4 11 a6
φ(r, θ) = V 2 P1 (cos θ) − 4 P3 (cos θ) + P5 (cos θ) + . . . . (3.4)
2r 8r 16 r6
Al rl + Bl r−l−1 Pl (cos θ)
X
φ(r, θ) = (3.5)
l
Let us conclude this section by looking at the expansion of this critical quantity
that occurs in the construction of the Green’s function. We begin by observing
that the result can depend only on r, r′ and γ, the angle between x and x′ . We
may thus simplify the problem by choosing the azimuthal direction (z axis) along
the x′ axis. The problem then displays manifest azimuthal symmetry, and we may
write
1 ∞
′ l ′ −l−1
X
= A l (r )r + B l (r )r Pl (cos γ) (3.6)
|x − x′ | l=0
We now consider the case where x lies parallel to x′ , when cos γ = 1. Then the
l.h.s. of eqn. 3.6 becomes
1 1
= .
|x − x′ | |r − r′ |
There are two cases:
l
r′ ∞ r ′l
1 1 1X ∞
′ X
r>r : = = =
|r − r′ | r − r′ r l=0 r l=0 r
l+1
1 1 1 X∞ r l X
! ∞ rl
′
r<r : = = ′ =
|r − r′ | r′ − r r l=0 r′ l=0 r
′l+1
Let us introduce r> = max(r, r′) and r< = min(r, r′). Then we may write
1 ∞ rl
X <
′
= l+1
|r − r | l=0 r>
and, comparing with eqn. 3.6, we have
1 ∞ rl
X <
′
= l+1 Pl (cos γ)
|x − x | l=0 r>
Chapter 2 13
Let us now consider the case where we no longer assume azimuthal symmetry.
Then we are concerned with solutions of the Generalised Legendre Equation,
m2
d 2 dP (x)
(1 − x ) + l(l + 1) −
P (x) = 0. (3.7)
dx dx 1 − x2
We can obtain a series solution in an analogous way to that of the ordinary
Legendre equation. For solutions to be finite at x = ±1, corresponding to θ = 0, π,
we require that l must be a positive integer or zero, and that m takes the values
m = −l, −l + 1, . . . , l − 1, l.
Recall that we already know that m must be an integer by the requirement that
the azimuthal function Q(ϕ) be single-valued.
For the case where m is positive, we can write the solutions Plm (x) as
dm
Plm (x) m
= (−1) (1 − x ) 2 m/2
Pl (x)
dxm
or for both positive and negative m by adopting Rodrigues’ formula:
(−1)m 2 m/2 d
l+m
Plm (x) = l
(1 − x ) l+m
(x2 − 1)l .
2 l! dx
Note that eqn. 3.7 depends only on m2 . Thus we have that Pl−m (x) must be
proportional to Plm (x), and in fact
(l − m)! m
Pl−m (x) = (−1)m P (x).
(l + m)! l
Eqn. 3.7 is an equation of Sturm-Liouville class, with eigenvalues l(l + 1). We can
apply the orthogonality theorem at fixed m, and we have
Z 1 2 (l + m)!
dxPlm m
′ (x)Pl (x) = δll′ .
−1 2l + 1 (l − m)!
Chapter 2 14
i.e. Z
dΩ Ylm(θ, ϕ)Yl∗′m′ (θ, ϕ) = δmm′ δll′ .
For the case m = 0, the solution clearly reduces to the Legendre polynomial, up
to some normalisation:
v
u 2l +1
u
Yl0(θ, ϕ) = t
Pl (cos θ)
4π
Chapter 2 15
3.4.1 Completeness
We can now write the general solution of the Laplace boundary value problem as
∞ X
l h
Alm rl + Blm r−l−1 Ylm (θ, ϕ)
X i
φ(r, θ, ϕ) =
l= m=−l
Consider two vectors x, x′ , with coordinates (r, θ, ϕ) and (r′, θ′, ϕ′) respectively.
Let γ be the angle between x and x′ , so that
x· x′
cos γ = ′
= cos θ cos θ′ + sin θ sin θ′ cos(ϕ − ϕ′).
|x||x |
Then we have
4π X l
∗
Pl (cos γ) = Ylm (θ′, ϕ′ )Ylm(θ, ϕ)
2l + 1 m=−l
This is proved in Jackson, but is more easily proved using group theory. Note
that we can rewrite this in the form
l (l − m)! m
Pl (cos γ) = Pl (cos θ)Pl (cos θ′ ) + 2 Pl (cos θ)Plm (cos θ′ ) cos m(ϕ − ϕ′ )
X
m=1 (l + m)!
Chapter 2 16
Example
1 ∞ rl
X <
′
= l+1 Pl (cos γ).
|x − x | l=0 r>
Using the addition theorem, we can rewrite this as
∞ X l l
1 X 1 r< ∗ ′ ′
′
= 4π l+1 Ylm (θ , ϕ )Ylm (θ, ϕ).
|x − x | l=0 m=−l 2l + 1 r>
Superficially, this looks like a much more complicated expression, since we have
introduced an additional sum over m. But it is now a sum over terms that factorise
into a function of (θ, ϕ) and a function of (θ′ , ϕ′), and thus much more useful.
θ z
y
ρ
x
Chapter 2 17
Z(z) = e±kz .
d2 R 1 dR ν2
+ + 1− 2 R =0
dx2 x dx x
The lowest power of x is xγ−2, and equating the coefficients of this to zero gives
the indicial equation which determines γ.
Chapter 2 19
• xγ−2 :
c0 γ(γ − 1) + c0 γ + ν 2c0 = 0 ⇒ γ = ±ν, since c0 6= 0.
• xγ−1 :
0 = c1 (γ + 1)γ + c1 (γ + 1) − ν 2c1
= c1 (γ 2 + 2γ + 1 − ν 2)
= c1 (2γ + 1) since γ 2 = ν 2
⇒ c1 = 0 since ν is an integer.
• xn+γ , n ≥ 0 :
As in the case of Legendre’s equation, the recurrence relation relates either odd or
even values of n. However, we have seen that c1 = 0. Thus cn = 0 for all odd n.
Therefore, let us make the substitution n = 2j, and write the recurrence relation
as
1
c2j+2 = − c2j , j = 0, 1, 2, . . .
4(j + 1)(j + 1 + γ)
c2j+1 = 0.
These are just another pair of linearly independent solutions of the Bessel equa-
tion:
These are also known as Hankel Functions. Their utility is that they have a
more straightforward integral representation than Jν (x) and Nν (x).
The sets of solutions of the Bessel equation are collectively known as cylinder
functions, and satisfy recursion relations in the same manner as the Legendre
polynomials, e.g.
2ν
Ων−1(x) + Ων+1(x) = Ων (x)
x
dΩν (x)
Ων−1(x) − Ων+1(x) = 2
dx
From the limiting forms 3.12, we see that each Bessel function has an infinite
number of roots, which we denote xνn , n = 1, 2, 3, . . . where
Jν (xνn) = 0, for x = 1, 2, 3, . . ..
In particular, we have
ν = 0 : x0n = 2.405, 5.520, 8.654, . . .
ν = 1 : x1n = 3.832, 7.016, 10.173, . . .
ν = 2 : x2n = 5.136, 8.417, 11.620, . . .
The roots of the Bessel function Jν (x) are crucial when we consider its orthogonal-
ity properties, which take a rather unexpected form. We introduce the functions
√
sJν (xνns/a), n = 1, 2, 3, . . .
and will now show that, for fixed ν ≥ 0, these functions, identified by n, form an
orthogonal set on 0 ≤ s ≤ a.
Chapter 2 23
Proof
p(x) = s,
q(x) = −ν 2/s,
r(x) = s,
λ = x2νn /a2 .
Thus we have
Z a
(x2νn − x2νn′ ) ds sJν (xνn′ s/a)Jν (xνn s/a) = 0
0
providing
)#a
d d
" (
s Jν xνn′ s/a) Jν (xνn′ s/a) − Jν (xνn s/a) Jν (xνn′ s/a) = 0. (3.13)
ds ds 0
At the upper limit, s = a, this expression vanishes since xνn and xνn′ are roots of
the Bessel function, and at the lower limit, s = 0, the expression vanishes because
of the factor of s. Thus we have
xνn s xνn′ s
Z a ! !
ds sJν Jν = 0, n 6= n′
0 a a
The integral can be evaluated for n′ = n, with the result
xνn s xνn′ s a2
Z a ! !
ds sJν Jν = [Jν+1(xνn)]2 δnn′ .
0 a a 2
Chapter 2 24
3.5.5 Completeness
We now assume that the Bessel functions satisfy the completeness relation, and
therefore we can expand any function on 0 ≤ s ≤ a as
∞
X
f (s) = Aνn Jν (xνns/a)
n=1
where
2 xνn s
Z a !
Aνn = 2 2 ds sf (s)Jν .
a Jν+1(xνn) 0 a
This is a Fourier-Bessel series. This expansion is particularly useful for the
case where f (a) = 0, e.g. the Dirichlet problem, since each term in the expansion
satisfies the boundary conditions. An alternative set of basis functions is provided
by
√ yνn s
!
sJν ,
a
where the yνn are the roots of dJν /dx = 0, because this set still satisfies the
condition of eqn. 3.13. This choice is often more appropriate for the Neumann
problem.
Note that if we had chosed a separation constant such that the solution in the
z-variable was
Z(z) = e±ikz ,
then the equation for R(s) would have been
d2R 1 dR 2 ν 2
+ − k + 2 R = 0,
ds2 s ds s
which, after our usual substitution x = ks, becomes
d2 R 1 d2 R ν2
+ − 1 + 2 R = 0.
dx2 x dx x
Chapter 2 25
with solutions
x≪1
!ν
1 x
Iν (x) →
Γ(ν + 1) 2
x
h i
− ln + γE + . . . ν=0
Kν (x) → 2ν
Γ(ν) 2
ν 6= 0
2 x
x ≫ 1, ν
1 1
" !#
Iν (x) → √ ex 1 + O
2πx x
π −x 1
s " !#
Kν (x) → e 1+O .
2x x
Note again that only Iν (x) is regular as x → 0.
Chapter 2 26
φ(s, ϕ, 0) = 0
φ(a, ϕ, z) = 0; 0 ≤ z ≤ L
φ(s, ϕ, L) = V (s, ϕ)
z φ = V (ρ,θ)
a L
φ=0
φ=0
x
where m is an integer greater than or equal to zero. The z factor is of the form
Z(z) = sinh kz
Chapter 2 27
where k is the separation constant, and we have imposed the boundary condition
Z(0) = 0. Finally, the radial component is of the form
Since there are no charges in the region s ≤ a, the solution must be regular there,
and in particular must be finite at s = 0. Thus we have Dm = 0. Furthermore, R
must vanish at s = a, and thus
Jm (ka) = 0
where xmn is the n-th root of Jm (x) = 0. Thus our general solution may be written
∞
X B0n
φ(s, ϕ, z) = J0(k0ns) sinh(k0nz)
n=1 2
∞ X
X ∞
+ Jm (kmns) sinh(kmn z) [Amn sin mϕ + Bmn cos mϕ].(3.14)
m=1 n=1
and thus
2 Z a Z 2π
Amn = ds s dϕ V (s, ϕ)Jm(kmns) sin mϕ
πa2 sinh(kmnL)[Jm+1(xmn)]2 0 0
2 Z a Z 2π
Bmn = ds s dϕ V (s, ϕ)Jm(kmns) cos mϕ
πa2 sinh(kmnL)[Jm+1(xmn)]2 0 0
Example
Determine φ(s, ϕ, z) in the upper-half plane z ≥ 0, with φ(s, ϕ, 0) = V (s, ϕ), and
φ finite as z → ∞. Then the separable solutions are of the form
but there is now no restriction on the value of k other than it be positive (to
ensure that φ is finite as z → ∞). Thus the sum over discrete values of k becomes
an integral over k, and our general solution is
Z ∞ B0(k)
φ(s, ϕ, z) = dk e−kz J0(ks)
0 2
∞ Z ∞
dk e−kz {Am (k) sin mϕ + Bm (k) cos mϕ)} Jm(ks).
X
+
m=1 0
We still have a Fourier series in ϕ, but the Fourier-Bessel series has evolved to a
Bessel transform.
Imposing the boundary conditions at z = 0, we have
Z ∞ B0(k)
V (s, ϕ) = dk J0(ks)
0 2
∞ Z ∞
X
+ dk {Am (k) sin mϕ + Bm (k) cos mϕ} Jm (ks)
m=1 0
Chapter 2 29
G(x, x′ ) = 0 for x ∈ ∂V .
rl a (a2/r)l
l
∞ X 1
X ∗ ′ ′ <
= 4π Ylm (θ , ϕ )Ylm(θ, ϕ) l+1 − .
l=0 m=−l 2l + 1 r> r r′l+1
where we note that, for the image charge, r> = r′ , r< = a2 /r, since the image
charge is always inside the sphere. Thus we have
l+1
l 2
∞ X
l 1 r<
1 a
G(x, x′ ) = 4π ∗ ′ ′
X
Y (θ , ϕ )Ylm(θ, ϕ) l+1 −
1 lm
l=0 m=−l 2l + r> a rr′
We have thus accomplished our goal of expressing the Green function as an ex-
pansion over orthogonal functions. There are some important observations we can
make by looking at the radial part
l+1 1 l a2l+1
l 2
r′l+1 r − r < r′
r< 1 a
rl+1
{ l+1 − ′ }= 1 a2l+1
.
r> a rr l+1 r ′l
r − r′l+1 r > r′
• It is symmetric under r ↔ r′ .
∇′2G(x, x′ ) = −4πδ(x − x′ ).
m′ =−∞
yielding
1 ∂ ′ ∂Fm m2 ∂ 2 Fm 2
" #
′ ′
s ′
− ′2
F m + ′2
= − ′
δ(s − s′ )δ(z − z ′ )eimϕ .
s ∂s ∂s s ∂z s
Chapter 2 33
′ ′
s ′
− ′2
f m + ′2
= − ′
δ(s − s′ )δ(z − z ′ ). (3.17)
s ∂s ∂m s ∂z s
Thus from our original P.D.E. in three variables we now have a two-variable
P.D.E..
To proceed further, we must say something about the boundary conditions, or at
the very least specify the volume V . We will assume that is covers −∞ ≤ z ≤ ∞,
and then introduce the Fourier transform (F.T.) of fm , defined by
Z ∞
f˜m (s, z; s′, k) =
′
dz ′ eikz fm (s, z; s′ , z ′ )
−∞
1 Z∞
dke−ikz f˜m (s, z; s′ , k).
′
′ ′
fm (s, z; s , z ) =
2π −∞
We apply the F.T. operator to eqn. 3.17,
m2 ∂ 2 fm
1 ∂ ′ ∂fm 2
Z ∞ " # Z ∞
′
′ ikz ′
dz ′ eikz s′ ∂s′
s ′
− ′2
f m + ′2
= − ′
δ(s − s′
) dz e δ(z − z ′ ),
∞ ∂s s ∂z s −∞
yielding
1 ∂ ′ ∂ f˜m m2 ˜
2˜ 2
′ ′
s ′
− ′2
f m − k f m = − ′
δ(s − s′ )eikz
s ∂s ∂s s s
where we have used the well-known properties concerning the F.T. of a derivative.
We have now exhibited the z dependence of the function, and may write
1 ikz
f˜m (s, z; s′, k) = e gm (s, s′; k),
2π
giving
m2
1 ∂ ∂gm 4π
" #
′ ′
s ′ − ′2 + k 2 gm = ′ δ(s − s′ ).
s ∂s ∂s s s
Chapter 2 34
x = ks
x′ = ks′ ,
yielding
∂ ′ ∂gm(x, x′) m2
′
x ′
− x 1 + ′2 gm (x, x′) = −4πδ(x − x′ ).
′
∂x ∂x x
This is just the modified Bessel equation, with inhomogeneous source. As we
noted earlier, the modified Bessel equation (like the Legendre equation) is of
Sturm-Liouville type:
d ′ dg(x, x′)
′
p(x ) ′
+ q(x′)g(x, x′) = −4πδ(x − x′)
dx dx
with
p(x′) = x′
m2
q(x′ ) = −x′ 1 + ′2 .
x
Thus we have finally reduced the problem to the solution of the Green function
for the Sturm-Liouville equation.
′
p(x ) ′
+ q(x′)g(x, x′) = −4πδ(x − x′),
dx dx
defined on the interval x′ ∈ [a, b], with homogeneous boundary conditions at a
and b. Note that we regard x as some arbitrary, fixed parameter.
The Green function must possess the following properties:
Chapter 2 35
1. For x′ 6= x, g(x, x′) satisfies the homogeneous equation, i.e. the Sturm-
Liouville equation with no source on the r.h.s..
leading to x+ǫ
dg(x, x′)
Z x+ǫ
′
p(x )
′
+ dx′ q(x′)g(x, x′) = −4π.
dx x−ǫ
x−ǫ
Both q(x′) and g(x, x′) are continuous at x′ = x, and therefore we have
lim Z x+ǫ
ǫ→0 dx′ q(x′)g(x, x′) = 0,
x−ǫ
p(x)× ǫ → 0 − = −4π
dx′ dx′
which we write as
dg(x, x′)
4π
= − ,
dx′ x =x
′ p(x)
i.e. there is a discontinuity in the slope of the Green function of magnitude 4π/p(x)
at x′ = x.
Chapter 2 36
Discontinuity in
slope
a x b
Thus we will write our Green function as
• a ≤ x′ ≤ x:
g(x, x′) = C1(x)y1(x′),
where y1 (x′) is a solution of the homogeneous equation satisfying the appro-
priate boundary condition at x′ = a.
• x ≤ x′ ≤ b:
g(x, x′) = C2(x)y2(x′)
where y2 (x′) is a solution of the homogeneous equation satisfying the appro-
priate boundary condition at x′ = b.
Note that this method only works if y1 and y2 are linearly independent, since
otherwise the Wronskian vanishes.
Thus our general form for the Green function is
4π y2 (x)y1(x′)
a ≤ x′ ≤ x
−
p(x) W [y1(x), y2(x)]
g(x, x′) =
4π y2 (x′)y1(x)
x ≤ x′ ≤ b
−
p(x) W [y1(x), y2(x)]
So, as we have already observed for spherical polar coordinates, the Green function
in the regions x′ < x and x′ > x comprises two different, linearly independent
solutions of the homogeneous equation.
We will now return to the case of the modified Bessel equation with δ-function
source
d ′ dg(x, x′)
′
p(x ) ′
+ q(x′)g(x, x′) = −4πδ(x − x′).
dx dx
A pair of linearly independent solutions is provided by the modified Bessel func-
tions Im (x) and Km(x). We will now consider the case where we require the
Chapter 2 38
solution over all space, i.e. x ∈ [0, ∞]. The solution must be finite at x = 0, and
thus
y1 (x′) = Im (x′).
If we further require that the solution be finite as x′ → ∞, then we have
y2 (x′) = Km (x′),
which we can see from the limiting behaviour quoted earlier. In this case, the
Wronskian is (see Jackson)
1
W [Im(x), Km(x)] = −
x
and thus our general solution for the Green function is
4π Km(x)Im(x′)
0 ≤ x′ ≤ x
−
x −1/x
gm (x, x′) =
,
4π Km(x′ )Im(x)
x ≤ x′ ≤ ∞
−
x −1/x
1.
2.
1 Z∞
dk e−ikz f˜m (s, z; s′ , z ′ )
′
′ ′
fm (s, z; s , z ) =
2π −∞
1Z∞ ′
= dk eik(z−z ) Im(ks< )Km(ks>)
π −∞
3.
1Z∞ ′
Fm (s, ϕ, z; s′, z ′ ) = dk eik(z−z ) Im (ks<)Km (ks>)eimϕ
π −∞
1 X∞
im(ϕ−ϕ′ ) ∞
Z
′
′
4. G(x, x ) = e dk eik(z−z ) Im (ks<)Km(ks> ).
π m=−∞ −∞
Since we have evaluated the Green function with boundary conditions at infin-
ity, this last expression is just the expansion of |x − x′ |−1 in cylindrical polar
coordinates.
Chapter 2 40
l′ ,m′
Substituting this into the inhomogeneous equation we have
1 ∂ ′2 ∂Fl′ m′
( " #
Yl∗′m′ (θ′, ϕ′)
X
′2 ′
r ′
l′ ,m′ r ∂r ∂r
∗
1 ∂ 2Yl∗′m′
F l ′ m′ 1 ∂ ′ ∂Yl′ m′
!
+ sin θ +
r′2 sin θ′ ∂θ′ ∂θ′ sin2 θ′ ∂ϕ′2
4π
= − ′2 δ(r − r′ )δ(ϕ − ϕ′)δ(cos θ − cos θ′ ).
r
Now the spherical harmonics are solutions of Laplace’s equation on the unit
sphere, and, from eqn. 3.1, satisfy
1 ∂ 2Yl∗′ m′ 1 ∂ ∗
′ ∂Yl′ m′
!
∗
+ l(l + 1)Yl′m′ + sin θ = 0.
sin2 θ′ ∂ϕ′2 sin θ′ ∂θ′ ∂θ′
Thus our Green function equation becomes
1 ∂ ′2 ∂Fl′ m′ F l ′ m′
( " # )
− ′2 l(l + 1) Yl∗′ m′ (θ′ , ϕ′) =
X
′2 ′
r ′
l′ m′ r ∂r ∂r r
4π
− ′2 δ(r − r′ )δ(ϕ − ϕ′)δ(cos θ − cos θ′ ).
r
Chapter 2 41
∗
We now multiply by Ylm (θ′, ϕ′), and use the orthogonality properties of the spher-
ical harmonics:
1 ∂ ′2 ∂Flm Flm
" #
r − l(l + 1)
r′2 ∂r′ ∂r′ r′2
4π Z
= − ′2 dΩ′ Ylm (θ′, ϕ′)δ(r − r′ )δ(ϕ − ϕ′)δ(cos θ − cos θ′ )
r
4π
= − ′2 Ylm(θ, ϕ)δ(r − r′ ).
r
We may then write
Flm (r, θ, ϕ; r′) = gl (r, r′)Ylm(θ, ϕ)
where gl (r, r′) satisfies
d ′2 d
!
′
r ′
gl (r, r′) − l(l + 1)gl (r, r′) = −4πδ(r − r′ ).
dr dr
This is just the radial part of Laplace’s equation. To proceed further, we must
specifiy boundary conditions.
yielding
a2l+1
′ ′l
y1(r ) = A1 r − ′l+1 .
r
Chapter 2 42
Noting that p(r) = r2, we observe that, once again, the Wronskian is independent
of the evaluation point, and we have general solution
a2l+1 1 rl
′l
A1 r − ′l+1 B2 l+1 − 2l+1
r r b
; a ≤ r′ ≤ r
−4π
2l+1
a
−A1 B2(2l + 1) 1 − 2l+1
b
gl (r, r′) =
2l+1
′l
,
l a 1 r
A1 r − l+1 B2 ′l+1 − 2l+1
r r b
; a ≤ r′ ≤ r
−4π
2l+1
a
−A1 B2(2l + 1) 1 − 2l+1
b
Note that it is also possible to recover this result using the method of images, but
in this case an infinite number of image charges are required.
Chapter 2 43
Example:
We can obtain the Green function by taking the a → 0 limit of eqn. 3.20:
l
(r<r> )l ∗ ′ ′
∞ X
l 4π r<
′ X
G(x, x ) = l+1 − Ylm (θ , ϕ )Ylm(θ, ϕ).
l=0 m=−l 2l + 1 r> b2l+1
The potential is then given by
′
1 Z′ 3 ′ ′ ′ 1 Z ′ ′ ∂G(x, x )
φ(x) = d x G(x, x )ρ(x ) − dS φ(x ) .
4πǫ0 V 4π S=∂V ∂n′
In our case the surface integral vanishes, because the potential vanishes there.
The (linear) charge density is given by
Q
ρ(x′ ) = 2
δ(r′ − a)δ(cos θ′).
2πa
Exercise: verify that the total charge is indeed Q. Thus the potential is
1 Z Q
φ(x) = dϕ′ d(cos θ′ )dr′ r′2 2
δ(r′ − a)δ(cos θ′ )
4πǫ0 2πa
rl l
∞ l 1 (r r )
∗ < < >
(θ′, ϕ′)Ylm(θ, ϕ) l+1
X X
× 4π Ylm − 2l+1
l=0 m=−l 2l + 1 r> b
In this case we have azimuthal symmetry, and the only non-vanishing integrals
arise from the terms with m = 0, for which
v
u 2l+1
u
Yl0(θ, ϕ) = t
Pl (cos θ).
4π
Chapter 2 44
Thus we have
rl (r<r>)l
1 Z ′ ′2 Q ′ ∞
X <
φ(x) = r r 2 δ(r − a) Pl (0)Pl (cos θ) l+1 − 2l+1
4πǫ0 a l=0 r> b
∞ (−1)n(2n − 1)!! r 2n 2n
Q X < r>
= r 2n+1
− P2n (cos θ),
4πǫ0 n=0 2nn! > b4n+1
where we have used
P2n+1(0) = 0
(−1)n(2n + 1)!!
P2n (0) = ,
2nn!
and r< = min(r, a), r> = max(r, a).
A closely related method to those discussed above is the expansion of the Green
function in terms of the eigenfunctions of some related problem. Consider the
solution of
∇2 ϕ(x) + [f (x) + λ]ϕ(x) = 0,
in a volume V bounded by a surface S, subject to ϕ satisfying certain homogeneous
boundary conditions for x ∈ S. In general, consistent solutions can be obtained
only for certain (possibly continuous) values of λ, which we will denote λn , the
eigenvalues. The corresponding solutions, the eigenfunctions, we will denote
ϕn(x). The eigenvalue equation is then
Then any function satisfying the same homogeneous boundary conditions may
be expanded as a series in the eigenfunctions. Consider in particular a Green
function, satisfying
n
and, inserting in eqn. 11.11.1, we obtain
We now use that ϕn is an eigenfunction of eqn. 3.21 with eigenvalue λn , and obtain
′ X ϕ∗n (x)ϕn(x′ )
G(x, x ) = 4π
n λn − λ
eik·(x −x) 1
′ !2
Z
′ 3
G(x, x ) = 4π dk
k2 2π
which we observe may be written
1 Z 3 eik·(x −x)
′
1
= dk .
|x − x′ | 2π 2 k2
∇2ϕ + klmn
2
ϕlmn = 0,
The simplest source for an electrostatic field is a point charge; such a source is
sometimes known as a pole. The arrangement of two point charges, of equal but
opposite sign, is known as a dipole. The concept of a dipole plays a crucial rôle
in electrostatics:
• Even in the case of a neutral atom or molecule, the positive and negative
charges can become separated, e.g. by an applied external electric field. In
that case, the atom or molecule gives rise to an electrostatic field that can
be approximated by a dipole
Consider two charges −q and q at x1 and x2 respectively, and let a be the position
vector of q relative to −q.
1
Chapter 2 2
-q a q
x’
x1
x2 P
x
Let x′ be the mid point of the dipole, so that
x1 = x′ − a/2
x2 = x′ + a/2
|x − x′ ± a/2|−1 ′ 2
± a · (x − x′ )
= |x − x | +
4
−1/2
a · (x − x′ ) |a|2
′ −1
= |x − x | 1 ± + .
|x − x′ |2 4|x − x′ |2
Expanding as a series in |a|2 /|x − x′|2 using the binomial expansion, we obtain
a · (x − x′ ) |a|2
!
1
|x − x′ ± a/2|−1 = |x − x′ |−1 1 + −
± +O
2 |x − x′ |2 |x − x′ |2
Thus
1 qa · (x − x′ )
φ(x) = .
4πǫ0 |x − x′ |3
Chapter 2 3
We now take the limit |a| → 0, q → ∞, with aq = p fixed and finite. This defines
a simple or ideal dipole and we have
1 p · (x − x′ )
φD (x) =
4πǫ0 |x − x′|3
In this subsection, we will consider not the field due to a dipole, but rather the
energy and forces on a dipole in an external field E(x) = −∇φ(x).
Recall from Section 3.5 that for a charge q in an electrostatic potential φ(x),
the potential energy is
U (x) = qφ(x)
Let us now apply this to the case of a dipole in an external field; once again, a
is the separation of the charge q from −q.
Chapter 2 4
If the separation between the charges is small, we can expand about x to obtain
φ(x ± a/2) =
1 ∂ 1 ai aj ∂ 2
φ(x) ± ai φ(r) + φ(x) + . . .
2 ∂xi 2! 4 ∂xi ∂xj
1
= φ(x) ± a · ∇φ(x) + O(a2 )
2
Thus we have
1 1
UD (x) = q [φ(x) + a · ∇φ(x) − φ(x) + a · ∇φ(x) + O(a3 )]
2 2
2
= q a · ∇φ(x)[1 + O(a )]
UD (x) = p · ∇φ(x)
Aside: why did I take x to be at the mid point of the dipole? Because for a simple
dipole, all the corrections to the formula above involving even derivatives of φ(x)
vanish. It just makes the expansion neater, but of course I could have performed
the expansion about any point between the charges.
Recalling that E(x) = ∇φ(x), we have
Chapter 2 5
UD (x) = −p · E(x)
Note that the potential energy of a dipole is a minimum when E and p are
parallel
We will now evaluate the torque, or moment of the force, τ on a simple dipole
about its centre. This is just the moment of the forces acting on the two charges
about the centre of the dipole:
1 1
! !
τ = a × (+q) E(x + a/2) + − a × (−q ) E(x − a/2)
2 ! 2
1
1 1 2
= a × q E(x) + (a · ∇) E(x) + E(x) − (a · ∇) E(x) + O(a )
2 2 2
ie τ = p × E(x) in the point dipole limit
• Note that the torque about some point other than the centre of the dipole
will be different.
• τ = p × E(x) is true for dipoles other than point dipoles if E(x) is constant
over the dipole.
Many materials are dipolar; the positive and negative materials are separated.
Here we will consider the force between a dipole p1 at x1 and p2 at x2. The force
F21 on the dipole at x2 due to the electrostatic field E1 produced by the dipole p1
is
· (x2 − x1 ) (x2 − x1 ) − p1 |x2 − x1 |2
3 p
1
F 21(x2) = (p2 ·∇2) E 1 (x2 ) = (p2 ·∇2 ) C |x2 − x1 |5
where ∇2 means that we take derivatives with respect to x2 (the position vector
of dipole p2), and we have used eqn. (4.1). As discussed above, we can express
this as
F21(x2) = ∇2 (p2 · E1 (x2)).
Note that the force F12(x1) is equal and opposite to F21(x2).
Chapter 2 7
In this section, we will see why the concept of dipoles, and more generally multi-
poles, is so important in electrostatics. Consider the case of a charge distribution,
localised to some volume V . For convenience we will take the origin for our vectors
inside V .
x - x’ P
We have that the potential due to the charge
V distribution within V at a point P outside the
x
x’ volume is:
1 Z ρ(x′ ) dV ′
φ(x) =
4πǫ0 V |x − x′ |
For r much larger than the extent of V , i.e. r ≫ r′ for all x′ such that ρ(x′ ) 6= 0,
we can expand the denominator
Thus we have
1 1 x · x′
′
= + 3 + O(r′2/r3).
|x − x | r r
Hence we can write
1 Q P · x 1 X 3 xi xj
φ(x) = + 3 + Qij 5 + O(1/r5)
4πǫ0 r r 2 i,j=1 r
where
Z
Q = ρ(x′ )dV ′ is the total charge within V
V
Chapter 2 8
Z
P = ρ(x′ )x′ dV ′ is the dipole moment of the charge about the origin
ZV
Qij = ρ(x′ )(3x′ix′j − r′2 δij )dV ′ is the quadrupole moment of the charge.
V
• We have defined the moments with respect to a particular point, e.g. the
dipole moment is the integral of the displacement x′ times the charge
density ρ(x′ ). In general, the moments depend on the choice of “origin”.
What about the total dipole moment when the total charge is zero?
• At large distances from the charge distribution, only the first few moments
(Q, P , quadrupole moment, . . .) are important.
Example:
ρ(x, y, z) = f z (a2 − r2 )
where f is a constant. Show that at large distances from the origin the potential
due to the charge distribution is given approximately by
2f a7 z
φ(x) =
105ǫ0 r3
Use the multipole expansion in SI units:
1 Q P ·x 1
!!
φ(x) = + + O
4πǫ0 r r3 r3
In spherical polars (r, θ, ϕ),
We assume that the charge is confined to a sphere of radius a, and take the centre
of the sphere to be the origin for our vectors. Then for the case r > a, we have
r< = r ′
r> = r,
and we have
1 X ∞ X l 1 Ylm(θ, ϕ) Z Z
φ(x) = l+1
dΩ′ dr′ r′2Ylm
∗
(θ′, ϕ′)r′l ρ(x′ ).
ǫ0 l=0 m=−l 2l + 1 r
Chapter 2 10
We now write Z
qlm = dΩ′ dr′r′2 Ylm
∗
(θ′ , ϕ′)r′l ρ(x′ )
so that the expansion may be written
1 X 1 Ylm (θ, ϕ)
φ(x) = qlm .
ǫ0 l,m 2l + 1 rl+1
This is the multipole expansion using spherical harmonics. To make the connec-
tion with our previous expansion, it is useful to consider the few few terms in
Cartesian coordinates
1 Z 3 ′ ′ 1
q00 = √ d x ρ(x ) = Q
4π 4π
v v
u 3 Z u 3
u u
3 ′ ′ ′ ′
q11 = − t
d x ρ(x )(x − iy ) = −t (Px − iPy )
8π 8π
v v
u 3 Z u 3
u u
q10 = t
d3 x′ ρ(x′ )z ′ = t Pz
4π
v
4π v
1 t 15 1u 15
u Z u
3 ′ ′ ′ ′ 2
u
q22 = d x ρ(x )(x − iy ) = t (Q11 − 2iQ12 − Q22)
4 2π 12 2π
v v
u 15 Z 1 u 15
u u
3 ′ ′ ′ ′ ′
q21 = − t
d x ρ(x )z (x − iy ) = − t (Q13 − iQ23)
v
8π 3 8π
v
1ut 5 1u 5
u Z u
q20 = d3x′ ρ(x′ )(3z ′2 − r′2) = t Q33.
2 4π 2 4π
Note that the components for negative m can be trivially obtained using
ql,−m = (−1)mqlm
∗
.
In general, for the l-th multipole moment, there are (l + 1)(l + 2)/2 components
in Cartesian coordinates, while only 2l + 1 components using spherical harmonics.
There is no inconsistency here - the Cartesian tensors are reducible under rota-
tions (i.e. mix with tensors having few indices under rotations) whilst the tensor
moments expressed in spherical harmonics are irreducible (i.e. the qlm for fixed
Chapter 2 11
l mix only amongst themselves under rotations); that is why we remove the trace
in the quadrupole moment Qij , to give us 5 irreducible components.
We can express the electric field components trivially in spherical harmonics. In
particular, the contribution of definite l, m is
1 l+1 1
Er = Ylm(θ, ϕ)qlm l+2
ǫ0 2l + 1 r
1 1 1 ∂
Eθ = − qlm l+2 Ylm(θ, ϕ)
ǫ0 2l + 1 r ∂θ
1 1 1 im
Eϕ = qlm l+2 Ylm(θ, ϕ).
ǫ0 2l + 1 r sin θ
If we now consider the case of a ideal dipole p along the z-axis, then
v
3
u
u
q10 = t
p
4π
q11 = q1,−1 = 0
and we have
v v
1 2u 3 3 p cos θ 2p cos θ
u u
u
Er = t t
3
=
ǫ0 3 4π 4π r 4πǫ0 r3
v v
1 1ut 3 1u 3 p sin θ
u u
Eθ = p 3 t sin θ =
ǫ0 3 4π r 4π 4πǫ0 r3
Eϕ = 0,
which reduces to the expression we derived earlier for an ideal dipole, eqn. 4.1.
There is a danger in using the expression for the electrostatic field due to an ideal,
or point dipole. To see this, consider the electrostatic field E(x) due to a localised
charge distribution ρ(x). In partciular, consider the integral of E over some sphere
of radius R, the centre of which we will take as the origin of our vectors.
Chapter 2 12
-
- -
+
+
+ +
We have Z Z Z
3 3 2
d xE = − d x ∇φ = −R dΩφ(x)n
r<R r<R
where n is a unit normal outward from the surface of the sphere, and we have
used the generalisation of the divergence theorem.
Using Coulomb’s law for an extended charge distribution, we may write
Z
3 R2 Z 3 ′ ′
Z
n
d xE = d x ρ(x ) dΩ .
4πǫ0 |x − x′ |
Now we can evaluate the x integration by writing the vector n = sin θ cos ϕi +
sin θ sin ϕj + cos θk, and then expressing these terms in spherical harmonics as
v
u 8π Y11 (θ, ϕ) + Y1,−1(θ, ϕ)
u
sin θ cos ϕ = −t
3 2
v
u 8π Y11 (θ, ϕ) − Y1,−1(θ, ϕ)
u
sin θ sin ϕ = −t
3 2i
v
u 4π
u
cos θ = t
Y10(θ, ϕ)
3
Chapter 2 13
Thus only the l = 1 terms contribute, and using the orthogonality property of
spherical harmonics we have
Z n 4π r< ′
dΩ = n
|x − x′ | 2
3 r>
where n′ is a unit vector in the direction of x′ . Hence we have
Z
3 R2 Z 3 ′ r< 4π ′ ′
d xE = − dx 2 n ρ(x )
4πǫ0 r> 3
R 2 Z 3 ′ r< ′ ′
= − d x 2 n ρ(x ) (4.3)
3ǫ0 r>
where r< = min(r′, R).
We now consider two cases
1. Sphere completely encloses the charge density. Then we have r< = r′ , and
r> = R, and we have, from eqn. 4.3,
Z
P
d3 xE = − , (4.4)
3ǫ0
where P is the electric dipole moment. Note that this expression is indepen-
dent of the size of the sphere, provided it completely encloses the dipole.
2. Charge density completely outside the sphere. Then we have r< = R, r> = r′ ,
and we have
Z
3 R 3 Z 3 ′ n′
d xE = − d x ′2 ρ(x′)
3ǫ0 r
4π 3
= R E(0).
3
Thus the average value of the electric field over a spherical volume containing
no charge is just the value of the field at the centre of the sphere.
Let us now consider the corresponding expression for the integrated E in the case
of an ideal dipole, eqn. 4.1:
Z
3
Z1 3(p · n)n − p
d xE(x) = d3 x
r<R r<R 4πǫ0 r3
= 0 Exercise: let p = pk, and work in spherical polars.
Chapter 2 14
For this to be consistent with eqn. 4.4, our expression for the electrostatic field
due to a dipole at x0 must be modified
1 3n(p · n) − p 4π
E(x) = − pδ(x − x0) .
4πǫ0 |x − x0|3 3
This expression only changes the electric field at the position of the dipole, and we
can then, with some care, use the expression as if we were using ideal, or point,
dipoles. The δ-function contains information about the finite distribution of the
charge lost in the multipole expansion.
Chapter 2 15
So far we have only considered the case of electrostatics in free space. We will
now consider the case of macroscopic materials in the presence of electric fields.
Such materials are classified according to whether or not electrons, or charges,
can flow over long distances. In the case of conductors, charges can move freely
about the material, and, as we have already seen, generate an induced field that
exactly cancels any applied external field.
Chapter 2 16
In this chapter we consider the case of dielectics. Here the electrons are bound
to atoms, and have only limited freedom to move. The material might have an
inherent dipole moment, or a dipole moment might be generated by the presence
of an external electric field. The crucial property of a dielectric is that
∇ × E = 0.
Thus
In the following, we will assume the applied field induces a dipole moment, but no
higher moments. Now consider the potential at x due to the charge, and dipole
moment, of a volume ∆V at x′ :
1 ρ(x′ )∆V P (x′ ) · (x − x′ )
∆φ(x, x′ ) = + ∆V ,
4πǫ0 |x − x′ | |x − x′ |3
where x is outside the volume ∆V . The dipole moment per unit volume is called
polarization. We now pass to an integral in the usual way, and obtain
1 Z 3 ′ ρ(x′ ) P (x′ ) · (x − x′ )
φ(x) = dx +
4πǫ0 V |x − x′ | |x − x′ |3
1 Z 3 ′ ρ(x′ )
1
= dx ′
+ P (x′ ) · ∇′ ′
(integ. by parts)
4πǫ0 |x − x | |x − x |
′
1 Z 3 ′ 1 ′ ′ ′ 1 Z ′ P (x ) · n
= dx [ρ(x ) − ∇ · P (x )] + dS
4πǫ0 V |x − x′ | 4πǫ0 S=∂V |x − x′ |
This expression can be rewritten as follows
1 Z 3 ′ ρf (x′ ) + ρb (x′) 1 Z ′
′ σb (x )
φ(x) = dx + dS
4πǫ0 V |x − x′ | 4πǫ0 S=∂V |x − x′ |
where σb ≡ P · n is the surface density of the bound charge, ρb ≡ −∇ · P is the
volume density of the bound charge, and the “old” charge density ρ is called the
free charge density ρf to distinguish from the density of the bound charge.
Chapter 2 17
P = ǫo χe E
D = ǫ0 E + ǫ0 χe E = ǫE
Tangential condition
We have that ∇ × E = 0, and thus, applying Stoke’s theorem to the closed curve
C shown above, we have Z
E · dl = 0,
C
yielding
|| ||
E1 = E2
which we can express as
(E 2 − E 1) × n21 = 0
Normal condition
(D2 − D1 ) · n21 = σf
(D2 − D1 ) · n21 = σf
Chapter 2 19
The method we adopt here essentially follows that of the solution of boundary-
value problems in vacua, with the boundaries given by conducting surfaces. The
method is best illustrated by examples.
Example:
In order to determine the potential in the region z > 0, let us try an image charge
q ′ at z = −d. Then the potential at x is
q′
1 q
φ(x)|z>0 = + .
4πǫ1 |x − dez | |x + dez |
We know that the potential in the region z < 0 must satisfy Laplace’s equation in
that region, and therefore, in particular, there cannot be any poles in the region
Chapter 2 20
z < 0. Therefore, let us try the potential due to a charge q ′′ at the postion of our
original charge q:
1 q ′′
φ(x)|z<0 = .
4πǫ2 |x − dez |
We now introduce cylindrical polar coordinates, so that
1 q ′′
z<0
4πǫ2 {ρ2 + (z − d)2}1/2
φ(ρ, θ, z) =
q′
1 q
+ z>0
4πǫ1 {ρ2 + (z − d)2}1/2 {ρ2 + (z + d)2}1/2
Thus we have a solution that satisfies the Laplace’s equation in z < 0, and Pois-
son’s equation in z > 0, and the correct boundary conditions at z = 0. Thus, by
our uniqueness theorem, it is the solution.
To see the form of the field lines we consider two cases, ǫ1 > ǫ2 and ǫ1 < ǫ2 ; in
both cases the field lines for z < 0 are those of a point charge, of magnitude q ′′ ,
at q.
1. ǫ1 > ǫ2 .
Then q ′ is same
sign as q.
2. ǫ2 > ǫ1 .
P i = (ǫi − ǫ0)E i.
Clearly the polarisation charge density vanishes, except at the point charge q, and
at the interface between the two materials. At the interface, there is a disconti-
Chapter 2 22
where n21 is the unit normal from region 1 to region 2, and P 2 and P 1 are the
polarisations at z = 0− and z = 0+ respectively. Thus we have
d q ′′
1 d ′ 1
σb = − (ǫ1 − ǫ0 ) (q − q) − (ǫ2 − ǫ0 )
4πǫ1 (ρ2 + d2)3/2 4πǫ2 (ρ2 + d2)3/2
d ǫ2 − ǫ0 ′′ ǫ1 − ǫ0 ′
( )
= q − (q − q)
4π(ρ2 + d2 )3/2 ǫ2 ǫ1
dq
= {2(ǫ2 − ǫ0 )ǫ1 + (ǫ1 − ǫ0 )(ǫ1 + ǫ2 + ǫ2 − ǫ1 )}
4π(ρ2 + d2 )3/2(ǫ1 + ǫ2 )ǫ1
q ǫ0 (ǫ2 − ǫ1 ) d
= −
2π ǫ1 (ǫ2 + ǫ1 ) (ρ + d2)3/2
2
Note that in the limit ǫ2 /ǫ1 ≫ 1, the electric field in region z < 0 becomes very
small, and the polarisation charge density approaches the value of the induced
surface charge density for a conductor at z = 0, up to the factor of ǫ0 /ǫ1. In that
sense, the material in z < 0 behaves as a conductor.
Example
E0
We will work in spherical polar coordinates, and express our solution as an ex-
pansion in Legendre polynomials:
l
l Al r Pl (cos θ) r<a
P
φ(r, θ, ϕ) = l −l−1
,
l [Bl r + Cl r ]Pl (cos θ) r>a
P
whence
Al rl−1Pl1 (cos θ)
X
− r<a
Eθ = l .
Cl r−l−2Pl1 (cos θ) − B1P11 (cos θ) r > a
X
−
l
The radial component is straightforward,
Al .l.rl−1Pl (cos θ)
X
− r<a
Er = l .
Cl (l + 1)r−l−2Pl (cos θ) − B1P1 (cos θ) r > a
X
l
l l
Al = Cl = 0, l 6= 1.
Chapter 2 25
• Outside the sphere, the field is equivalent to that of the applied field, together
with that due to a point dipole at the origin, of moment
ǫ/ǫ0 − 1 3
p = 4πǫ0 a E0 , (4.12)
2 + ǫ/ǫ0
orientated in the direction of the applied field.
which is just the dipole moment we obtained in eqn. 4.12. Thus the dipole mo-
ment is just the volume integral of the polarisation.
where we have integrated by parts, assuming that the charge is localised so that
the surface term vanishes. Thus the total energy in constructing the system is
Z Z D
3
W = dx E · δD.
0
We can recover our expression eqn. 4.13 either by the substitution E = ∇φ and
using ∇ · D = ρf , or by noting the linear relation between φ and ρ. The crucial
observation is that the expression eqn. 4.13 is valid only if the relation between D
and E is linear.
(ǫ1 − ǫ0 )E = P ,
yielding
1Z 3
W =− d x P · E 0.
2 V1
1
We can interpret w = − P · E 0 as the energy density of the dielectric. The
2
expression can be likened to that for the energy of a dipole distribution derived
at the end of Section 4.3. There we were considering a permanent dipole, whilst
here energy is expended in polarizing the dielectric, and this is reflected in the
factor of 1/2.
Note that the energy tends to decrease if the dielectric moves to a region of
increasing E 0 , providing ǫ1 > ǫ0 . Since the charges are held fixed, the total
energy is conserved, and we can interpret the change in field energy W due to
Chapter 2 30
We will conclude this section by considering the contrasting case where we intro-
duce a dielectric body into a system where the potentials, rather than charges,
are kept fixed. A paradigm is the introduction of a dielectric between the plates
of a capacitor connected to a battery, and hence at a fixed potential difference.
Fixed
Potentials
Dielectric
In this case, charges can flow to or from the conducting plates as the dielectric
is introduced to maintain the potentials, and hence the total energy can change.
Again, we will assume that the media are linear.
It is sufficient to consider small changes to the potential δφ and to the charge
distribution δφ, for which the change in energy δW , from eqn. 4.13, is
1Z 3
δW = d x (ρf δφ + φδρf ).
2
For the case of linear media, these two terms are equal if the dielectric properties
are unaltered. However, in the case where the dielectric properties are altered
during the change, ǫ(x) → ǫ(x) + δǫ(x), this is no longer true, because of a
Chapter 2 31
2. We now reconnect the battery. The potential on the conductors, where the
only macroscopic charges reside, must regain its original value, i.e. δφ2 =
−δφ1 , and there is a corresponding change in charge density δρ2f , yielding
1Z 3
δW2 = d x (ρf δφ2 + φ2 δρ2f ).
2
In this step, the dielectric properties are unaltered and the two terms are
equal, so we have
Z
δW2 = d3 x ρf δφ2
Z
= − d3x ρf δφ1
= −2 δW1
which we write as
δWV = −δWQ ,
i.e. the change in energy at fixed potential is minus the change in energy at fixed
charges. In this case, if a dielectric with ǫ1 > ǫ0 moves into a region at fixed
Chapter 2 32
Magnetostatics
5.1 Introduction
The crucial difference between electric and magnetic phenomena is the absence
of isolated magnetic charges, or magnetic monopoles. Here the basic building
blocks are magnetic dipoles. For a magnetic field, or flux density, B, the torque
τ acting on a dipole of moment µ is
τ = µ × B.
The other concept we need in the study of magnetostatics is the electric current
J, defined as the flow of charge per unit time per unit area, with normal in the
direction of J.
dI dQ ˆ
J= = J
da⊥ da⊥ dt
1
Chapter 2 2
∇ · J = 0.
dl × (x − x′ )
dB = kI
|x − x′ |3
where, in SI units,
For a point charge q moving with velocity v, we can replace Idl by qv, and we
have
µ0 qv × (x − x′ )
B= ,
4π |x − x′ |3
providing v is constant, and small compared to the velocity of light.
Chapter 2 3
We can apply the superposition principle to the magnetic field, and obtain for a
general current density
µ0 Z 3 ′ J(x′ ) × (x − x′ )
B(x) = dx .
4π |x − x′ |3
Example
dF = Idl × B.
Thus the force on a closed loop of current l1 due to magnetic field from closed
loop l2 is
µ0 I I dl × (x − x )
2 1 2
F 12 = I1 I2 dl1 × 3
4π |x1 − x2 |
µ0 I I dl × [dl × (x − x )]
1 2 1 2
= I1 I2 .
4π |x1 − x2 |3
Chapter 2 4
yielding
µ0 I1I2 I I dl1 · dl2 dl1 · x12
F 12 = − x + dl . (5.1)
4π |x12|3 12 2
|x12|3
We will now show that the second term vanishes. Consider the integration around
loop 1, for fixed x2. Then under a change x1 → x1 + dl1, we have
x · dl
= − 12 3 1 .
|x12|
Thus the integrand in the second terms of eqn. 5.1 is an exact differential, and
therefore the integrand around the closed loop vanishes, and we have
µ0 I1I2 Z dl1 · dl2
F 12 = x .
4π |x12|3 12
Now Newton’s third law is satisfied explicity, and we have
F 12 = −F 21 .
In analogy with electrostatics, our starting point is the expression for B due to
general current density
µ0 Z 3 ′ J(x′ ) × (x − x′ )
B(x) = dx .
4π |x − x′ |3
We begin by recalling that ∇ × (ϕa) = ∇ϕ × a, where a is a constant vector.
Thus
′ 1 1 × J(x′ )
∇x × J(x ) ′
= ∇
x ′
|x − x | |x − x |
(x − x′ )
= − ′ 3
× J(x′ )
|x − x |
Thus we can write
µ0 Z 1
B(x) = ∇ × d3 x ′ ′
J(x′ ) (5.2)
4π |x − x |
∇ · B = 0. (5.3)
This is another of Maxwell’s equations, and is just another statement that you
cannot have isolated magnetic charges, and that the total flux of B through any
closed surface vanishes Z
dS · B = 0.
S=∂V
To obtain another differential equation, we evaluate ∇ × B. We begin by recalling
the vector identity
∇ × (∇ × A) = ∇(∇ · A) − ∇2 A,
so that
J(x′ )
µ0 Z 3 ′ µ0 Z 3 ′ ′ 2 1
∇×B = ∇ d x ∇x · − d x J(x )∇x .
4π |x − x′ | 4π |x − x′ |
Chapter 2 6
Now
J(x′ )
1
∇x · ′
= −J(x′ ) · ∇x′
|x − x | |x − x′ |
2 1
∇x ′
= −4πδ(x − x′ ),
|x − x |
and thus
µ0 Z 3 ′ ′ 1 Z
∇ × B = − ∇ d x J(x ) · ∇x′
′
+ µ0 d3x′ J(x′)δ(x − x′ )
4π |x − x |
µ0 Z 3 ′ 1
= ∇ dx ∇ ′ · J + µ0 J(x)
4π |x − x′ | x
For, for magnetostatics, we have ∇ · J = 0, and thus
∇ × B = µ0 J (5.4)
This is the second fundamental differential equation. We can apply Stoke’s theo-
rem to a closed curve C spanned by a surface S to obtain
Z Z Z
∇ × B · dS = B · dl = µ0 J · dS.
Chapter 2 7
∇·B = 0
∇ × B = µ0 J
B = ∇ × A.
In the case where B is the magnetic field, we call A the magnetic vector po-
tential.
∇ × (∇f ) = 0.
Examples
2. We could require
∇ · A = 0 ∀x.
This is the Coulomb Gauge.
Choosing, or fixing, the gauge reduces the number of degrees of freedom, clear in
example (1) above. All the fundamental forces of nature are described by Gauge
Theories, having the property of a gauge, or local, symmetry.
We will specify that we work in the Coulomb gauge, ∇ · A = 0. Then the second
of our governing equation becomes
∇ × B = ∇ × (∇ × A) = ∇(∇ · A) − ∇2 A = µ0 J
and thus
∇2A = −µ0 J.
This is just Poisson’s equation, applied to each of the Cartesian components of
A, and from our investigation of electrostatics has the solution
µ0 Z 3 ′ J(x′ )
A(x) = dx . (5.5)
4π |x − x′ |
Example
y
ϕ
I
x
Chapter 2 9
The current is purely in the azimuthal direction, and in spherical polars, we can
write the current density as
′ ′ δ(r′ − a)
Jϕ = I sin θ δ(cos θ ) .
a
You should convince yourself that this expression is correct. W.l.o.g. we will
consider the case where the observation point is in the x − z plane, so that, in
Cartesian coordinates, the current density is
where we write
′
cos ϕ′ = ℜeiϕ .
Chapter 2 10
Consider a localized current distribution J(x′ ) , and the magnetic vector potential
produced at a point P (x) where |x| ≫ |x′ |. Then we can write
1 1 x · x′
= + + ...,
|x − x′ | |x| |x|3
so that, in the Coulomb gauge
µ0 1 Z 3 ′ ′ x Z 3 ′ ′ ′
Ai (x) = d x J i (x ) + · d x J i (x )x + . . . (5.6)
4π |x| |x|3
Chapter 2 11
where in the second step we have integrated by parts, using the fact that the
surface integral vanishes for a localised current distribution. Thus we have
Z
d3x′ [f J · ∇′ g + gJ · ∇′ f ] = 0 (5.7)
We now consider the first term in eqn. 5.6. Applying eqn. 5.7 for the case f (x′ ) =
1, g(x′ ) = x′i, we have
Z
d3x′ [Jj δij + x′iJj · 0] = 0
Z
⇒ d3x′ Ji = 0.
Thus the first term vanishes. This is just a further restatement that there is no
“monopole” contribution to the multipole expansion for magnetic fields.
We now applying the identity to the case f = x′i , and g = x′j . Then from eqn. 5.7,
we have
Z
3 ′ ′ ∂x′j ∂x′i
dx [xiJk ′ + x′j Jk ] = 0
∂xk ∂x′k
Z
⇒ 3 ′
d x [x′iJj + x′j Ji ] = 0.
Levi-Civita Tensor
To take the discussion further, we recall the definition of the Levi-Civita tensor
0 if any two if i, j, k are equal
ǫijk = 1 if (ijk) is an even permutation of (123)
−1 if (ijk) is an odd permutation of (123)
A × B|i = ǫijk Aj Bk .
Thus we have
1 µ0 1 Z
3 ′ ′
Ai (x) = − x × d x x × J
2 4π |x|3 i
The vector
1Z 3 ′ ′
m= d x x ×J
2
is the magnetic moment, whilst
1
µ = x′ × J
2
is the magnetic moment density. Thus we can write
Chapter 2 13
µ0 1
A(x) = m×x
4π |x|3
This is the lowest, non-vanishing term in the multipole expansion of the magnetic
vector potential for a localised current density. Applying B = ∇ × A, we have
µ0 3(x · m)x − r2m
B= ,
4π r5
exactly analogous to the electrostatic field due to a point dipole.2
Example
Example
We conclude this section by considering the case where the current distribution
arises from the motion of a number of charged point-like particles:
X
J= qi vi δ(x − xi ),
i
2
It is possible to introduce a vector potential to describe electric dipole fields
Chapter 2 14
where vi is the velocity of the ith particle, which we assume is much less than the
velocity of light.
Then we have
1X
m= qi x i × v i .
2 i
Now the orbital angular momentum of a particle is given by
Li = Mi xi × v i ,
First, consider a magnetic dipole in the uniform magnetic field B. Let us visualize
magnetic dipole m as a wire loop with area a carrying current I such as m = Ia.
The total force acting on the loop is zero:
I I
F = I dl × B = − IB × dl = 0
I I
′ II ′
′
N = − B × (x × dx ) = ( x × dx′ ) × B = m × B
2 2
so the torque in a uniform external field is a cross product of the magnetic moment
and the field.
Let us now consider a small dipole in the non-uniform external field (the size of
the dipole ≪ characteristic size of the field). The formula for the torque remains
the same: N = m × B where the magnetic field should be taken at the position
of the dipole. However, the total force is no longer zero.
I
F = I dl × B 6= 0
Chapter 2 16
Since our dipole is small we can expand B(x′ ) in powers of x′ . For simplicity,
suppose that the dipole is located at the origin. We get
because ∇ · B = 0.
Since F = −∇U we see that the potential energy of a (small) magnetic dipole in
the external magnetic field is
U = −m·B
M (x′ )
Z
3 ′ ′ ′ 1 Z
3 ′ ′
d x (∇ × M (x ) + dx ∇ × ,
| x − x′ | | x − x′ |
Chapter 2 18
Using the divergence theorem for vector fields (again, see the cover)
Z Z
3 ′ ′ ′
d x (∇ × A(x ) = n × A dS (5.9)
V S=∂V
∇ × B = µ0 J f ,
we have
∇ × B = µ0 {J f + ∇ × M }.
It is now conventional to introduce the magnetic field H, where
1
H = B − M.
µ0
In the context of media, the field B is known as the magnetic induction or
magnetic flux density. In terms of H and B, the fundamental equations of
magnetostatics in matter are
∇·B = 0
∇ × H = Jf
Chapter 2 19
M = χm H,
B = F(H)
Chapter 2 20
We will now obtain boundary conditions for the normal and tangential compo-
nents of the field at the boundary between two materials. Note that the followign
discussion is independent of whether or not there is a linear relation between the
H and B.
δΑ
n
Normal Condition
where n is a unit normal from medium 1 to medium 2, and δA is the surface area
of the pillbox. Thus we have
B⊥ ⊥
1 = B2
Chapter 2 21
Tangential Condition
We now apply Stoke’s theorem to get the boundary conditions on the tangential
components: I Z Z
H · dl = (∇ × H) · da = J f · da,
c S S
H 2 −H 1 = K ×n ⇒ n×(H 2 −H 1) = K
We will now look at various methods of solving boundary value problems be-
tween different media. The method depends on nature of the constitutive relation
between B and H, and on whether there is non-zero current density.
The magnetic field is always solenoidal, and therefore we can essentially always
introduce a vector potential A such that B = ∇ × A.
The dynamical information for the magnetostatics of media is provided by the
equation
∇ × H = Jf.
We will now specialise to the case where we have a linear constitutive relation,
B = µH, enabling us to write
∇×A
" #
∇× = Jf.
µ
Chapter 2 22
∇2 A = −µJ f .
This is analogous to the case discussed in Section 5.4.2, and the solution is that
of eqn. 5.5, with µ0 replaced by µ.
∇2φM = 0,
where we assume that µ is piecewise constant, i.e has a constant value in each of
the different media we are considering.
M constant
Since J f ≡ 0, we can solve this problem using either a scalar or a vector potential.
∇·B = 0 (5.10)
∇×H = 0 (5.11)
1
H = B−M (5.12)
µ0
We will introduce a scalar potential for the magnetic field,
H = −∇φM .
∇2φM = −ρM ,
where
ρM = −∇ · M.
In the case where there are no boundaries, this equation has the solution
1 Z 3 ′ ρM
φM = dx
4π |x − x′ |
Chapter 2 24
1 Z 3 ′ ′ 1
= − dx∇ ′
M (x′ ) (integration by parts)
4π |x − x |
1 Z
1
= − ∇ · d3 x ′ ′
M (x′ ).
4π |x − x |
Note that if we are far away from a non-zero M , i.e. f ≫ r′ , then we have
1 1 Z 3 ′ 1
!
φM ≃− ∇ · d x M (x′ ) = m · x.
4π r 4πr3
where Z
m= d3x′ M (x′ ).
Suppose now that we had a hard ferrormagnet confined to a volume V , with
surface S. Then there is a contribution arising from the discontinuity in M at the
surface, which we can express as a surface magnetisation density,
σM = n · M ,
ez
M
∞ X l l
1 X 1 r< ∗ ′ ′
′
= 4π l+1 Ylm (θ , ϕ )Ylm (θ, ϕ).
|x − x | l=0 m=−l 2l + 1 r>
√
Noting that cos θ′ = P1 (cos θ′ ) = 4πY10(θ′, ϕ′ ), and using orthogonality, we can
write
1 r<
φM (r, θ) = M0 a2 2 cos θ,
3 r>
where r<{>} = min{max}(r, a).
Inside the sphere, we have r< = r and r> = a. Thus
1 1
φM = M0 r cos θ = M0 z.
3 3
Thus we have
Hin = −∇φM = − 13 M
,
Bin = µ0 (H + M ) = 32 µ0 M
B0 = µ0 H0
a
b
Chapter 2 28
Since the current density is zero, we can once again write H = −∇φM . Further-
more, B = µH, and thus ∇ · H = 0 so that the scalar potential satisfies
∇2φM = 0,
where we have imposed that there be a uniform field at infinity for the case r > b,
and that the solution is regular as r → 0.
We now impose the boundary conditions at the interfaces r = a and r = b
B ⊥ is continuous
H k is continuous
which become:
∂φM ∂φM
(b+) = (b−)
∂θ ∂θ
∂φM ∂φM
µ0 (b+) = µ (b−)
∂r ∂r
∂φM ∂φM
(a+) = (a− )
∂θ ∂θ
∂φM ∂φM
µ0 (a−) = µ (a+)
∂r ∂r
We now use these equations to determine the coefficients αl , βl , γl , noting that
∂
Pl (cos θ) = Pl1 (cos θ).
∂θ
Chapter 2 29
All the coefficients vanish for l > 1 (exercise), and we have (see Jackson)
(2µ′ + 1)(µ′ − 1)
α1 = (b3 − a3 )H
0
2a 3
(2µ′ + 1)(µ′ + 2) − b3 (µ′ − 1)2
′
9µ
δ1 = − ′ ′ 2a3 ′ 2
H ,
0 (5.16)
(2µ + 1)(µ + 2) − b3 (µ − 1)
where µ′ = µ/µ0 .
For r > b, we have the uniform field together with a dipole of moment α1 , parallel
to H1:
α1
φM = −H0 r cos θ + 2 cos θ.
r
For r < a, there is a uniform magnetic field parallel to H0, of magnitude −δ1 :
φM = −(−δ1)r cos θ.
From eqn. (5.16), we see that δ1 ≃ 1/µ′ as µ′ → ∞: the effect of a shell of high
permeability is to shield the interior from the magnetic field.
Chapter 6
∇·D = ρ
∇×E = 0 (6.1)
and
∇×H = J
∇ · B = 0. (6.2)
Electric and magnetic phenomena are completely separate, except for the fact
that current density is associated with the motion of charges.
Faraday (1831) observed that a current could be induced in a closed loop of wire
by varying the flux of magnetic field through a surface spanning the loop.
1
Chapter 2 2
B C
dφ
E = −k ,
dt
where, in SI units, k = 1. Note that the sign here is a consequence of Lenz’s law:
the induced current is in such a direction as to oppose the change of flux producing
it. You could argue that the whole application of electricity in the modern world
dates rests on Faraday’s law; the observation that a changing magnetic field can
produce an electric current.
Chapter 2 3
We can can generalise this integral equation as applying to any closed curve in
space, spanned by a surface,
I
d Z
E · dl = − B · dS.
C dt S
We now apply Stokes’ theorem to the l.h.s.,
I Z
E · dl = (∇ × E) · dS.
S
Specialising to the case where both C and S are fixed in time, we have
d Z Z
∂B
B · dS = · dS,
dt S S ∂t
and thus Z Z ∂B
(∇ × E) · dS = − · dS,
S S ∂t
yielding
Z
∂B
(∇ × E + ) · dS = 0.
S ∂t
Since both C, S are arbitrary, we obtain the differential form of Faradays law,
∂B
∇×E+ =0
∂t
∇ · D = ρ,
With this final modification of Ampere’s law, and Faraday’s law, we have the
completed the construction of Maxwell’s equations
∇·B =0
B = ∇ × A.
B = ∇×A (6.3)
∂A
E = −∇φ − (6.4)
∂t
The two remaining equations (ME1 and ME3) determine the dynamical be-
haviour, i.e. the dependence of A and φ on t and x. To solve them, we need
some constitutive relation between (D, H) and (E, B). We will initially restrict
ourselves to the case of the vacuum, where we have
D = ǫ0 E
1
H = B.
µ0
Coulomb’s law, ME1, is thus
∇ · E = ρ/ǫ0
whilst Ampère’s law, ME3, is
1 ∂E
∇ × B = J + ǫ0 .
µ0 ∂t
Thus, in terms of the potential (φ, A), ME1 becomes
∂
∇2 φ + (∇ · A) = −ρ/ǫ0 (6.5)
∂t
1 ∂ 2A 1 ∂φ
" #
2
∇ A− 2 2 −∇ ∇·A+ 2 = −µ0 J (6.6)
c ∂t c ∂t
Thus we have derived two, coupled second-order PDE’s that are, with the def-
initions of the potentials in eqns (6.3) and (6.4), equivalent to the original four
Maxwell equations.
Chapter 2 7
A −→ A′ = A + ∇Λ (6.7)
∂Λ
φ −→ φ′ = φ − (6.8)
∂t
1 ∂φ
∇·A+ = 0. (6.9)
c2 ∂t
This is known as the Lorentz condition, and the dynamical equations assume
the form
2 1 ∂ 2φ
∇ φ − 2 2 = −ρ/ǫ0 (6.10)
c ∂t
2 1 ∂ 2A
∇ A − 2 2 = −µ0 J. (6.11)
c ∂t
The A and φ fields have become decoupled, and the simplified equations are just
the wave equations, with a inhomogeneous source. But is it actually possible
to find a gauge transformation that satisfies eqn. (6.9)?
Let (A, φ) be potentials satisfying eqns. (6.6) and (6.5), and let Λ be a gauge
transformation such that the transformed fields satisfy eqn. (6.9). Then we have
1 ∂φ′
′
∇·A + 2 = 0
c ∂t
1 ∂φ ∂ 2Λ
2
=⇒ ∇ · A + ∇ Λ + 2 − 2 = 0
c ∂t ∂t
Thus we need to find Λ satisfying
1 ∂ 2Λ 1 ∂φ
" #
2
∇ Λ− 2 2 =− ∇·A+ 2 . (6.12)
c ∂t c ∂t
Chapter 2 9
Note that the Lorentz condition does not specify a gauge uniquely. Let (A, φ)
satisfy the Lorentz condition. Now consider the transformation
A −→ A′ = A + ∇Λ
∂Λ
φ −→ φ′ = φ − .
∂t
Then the Lorentz condition transforms as
1 ∂φ ′ 1 ∂φ′ 2 1 ∂ 2Λ
∇·A+ 2 −→ ∇ · A + 2 =∇ Λ− 2 2
c ∂t c ∂t c ∂t
Thus the new gauge also satisfies the Lorentz condition, providing
1 ∂ 2Λ
∇2 Λ − = 0.
c2 ∂t2
The Lorentz gauge is important because:
• (A, φ) are treated on an equal footing and, when we discuss Special Relativity,
we will see that the Lorentz condition is Lorentz covariant, i.e. independent
of the choice of coordinate system.
2 1 ∂ 2A 1 ∂φ
∇ A − 2 2 = −µ0 J + 2 ∇ . (6.14)
c ∂t c ∂t
Note that the scalar potential φ(x, t) is the instantaneous Coulomb potential
due to a charge density ρ(x, t), i.e. we do not take account of “causality” through
the use of a retarded potential.
The equation for the vector potential contains a gradient operator, ∇∂φ/∂t arising
from the solution of Poisson’s equation for the scalar potential, and this term is
irrotational,
∂φ
" #
∇× ∇ = 0.
∂t
It would be useful to completely decouple the equations governing the vector and
scalar potentials, as in the case of the Lorentz gauge. To accomplish this, we will
separate the current into an irrotational, or longitudinal, piece and a solenoidal,
or transverse, piece,
J = Jl + Jt (6.15)
with
∇ × Jl = 0
∇ · Jt = 0.
d3k ik·(x−y ) ki kj ∂ ∂ Z d3 k
eik·(x−y )
Z Z Z
3 2 3
= d yJ(y) 3
e (δij − 2 ) = ( − δij ∇ ) d yJ(y) 3
8π k ∂xi ∂xj 8π
∂ ∂ Z J(y)
=( − δij ∇2 ) d3 y
∂xi ∂xj 4π|x − y|
and similarly
∂ ∂ Z J(y)
Jil (x) =− d3 y
∂xi ∂xj 4π|x − y|
Using the formula
1
∇2 = −4πδ(x − y)
|x − y|
it is easy to check the self-consistency Jit (x) + Jil (x) = Ji (x).
Thus we have performed the decomposition of eqn. (6.15) with
1 Z 3 ′ ∇′ · J(x′ )
Jl = − ∇ d x (6.17)
4π |x − x′ |
′
1 Z J(x )
Jt = ∇ × ∇ × d3 x′ (6.18)
4π |x − x′ |
Now, from the continuity equation, we have
∂ρ
∇ · Jl + = 0.
∂t
and substituting in eqn. (6.17) we obtain
1 Z 3 ′ 1 ∂ρ
Jl = ∇ dx .
4π |x − x′ | ∂t
We now identify the r.h.s. of this equation with our expression for the scalar
potential of eqn. (6.13) and observe that
∂φ
Jl = ǫ0 ∇
∂t
1 ∂φ
=⇒ µ0 Jl = 2 ∇ ,
c ∂t
Chapter 2 12
where we have used µ0 ǫ0 = 1/c2. Thus, returning to the equation for the vector
potential, eqn. (6.14), we find
1 ∂ 2A
∇2 A − = −µ0 Jt . (6.19)
c2 ∂t2
Only the transverse part of the current is a source for A. Thus this gauge is also
known as the transverse or radiative gauge, and once again we have decoupled
the scalar and vector potentials.
Chapter 2 13
In both the Lorentz and Coulomb gauges, we have reduced the problem of finding
the potentials to the solution of the wave equation
1 ∂ 2ψ
∇2 ψ − = −4πf (x, t), (6.20)
c2 ∂t2
where f is some known source, and c, as we have intimated earlier, is the velocity
of wave propagation.
Such a hyperbolic equation, like the elliptic equations encountered in electrostat-
ics, can be solved by means of Green functions. In particular, we will find the
Green function G(x, t; x′ , t′) satsifying
1 ∂2
∇2 − G(x, t; x′ , t′ ) = −4πδ(x − x′ )δ(t − t′ ). (6.21)
2
c ∂t2
The solution to the inhomogeneous wave equation, eqn. (6.20), for a general source
is then Z
ψ(x, t) = ψ0(x, t) + d3x′ dt′ G(x, t; x′ , t′)f (x′ , t′ )
where ψ0 is a solution of the homogeneous equation. Note that this is essen-
tially an initial-value problem, rather than the boundary-value problem encoun-
tered with elliptic equations.
To obtain the Green function, we take the Fourier transform with respect to t:
1 Z
G(x, t; x′ , t′) = dωe−iωtg(x, ω; x′ , t′ )
Z2π
′ ′
g(x, ω; x , t ) = dt eiωt G(x, t; x′ , t′ )
d3 xe−iq·xg(x, ω; x′ , t′ ),
Z
′ ′
g̃(q, ω; x , t ) =
Chapter 2 14
yielding
e−iq ·x eiωt
′
′ ′
=⇒ g̃(q, ω; x , t ) = 4π 2
q − k2
where k ≡ ω/c is the wave number. We can invert this expression to obtain
eiq|x−x | cos θ
′
′ ′ 4π iωt′ Z ∞ 2 2π
Z Z 1
g(x, ω; x , t ) = e dq q dψ d(cos θ) 2
(2π)3 0 0 −1 q − k2
q 2 eiq|x−x | e−iq|x−x |
′ ′
4π iωt′ Z ∞
= e dq 2 −
(2π)2 0 q − k 2 iq|x − x′| iq|x − x′ |
′
4π eiωt Z ∞ dq q iq|x−x′ |
= e
(2π)2 i|x − x′ | −∞ q 2 − k 2
The integrand has poles at q = ±k, and therefore we have to specify how to treat
the poles in order to evaluate the integrals. We will do this by displacing the poles
off the real axis as follows:
(±) ′ ′ 4π eiωt Z ∞ dq q iq|x−x′ |
g (x, ω; x , t ) = e ,
(2π)2 i|x − x′ | −∞ q 2 − k 2 ∓ iη
where η is small. We now write
q 2 − k 2 ∓ iη = (q − k ∓ iǫ)(q + k ± iǫ),
k+i ε
-k-i ε
We can complete the contour in the upper-half plane, where the contribution from
the semi-circle at infinity vanishes, and obtain
1
eiωt +ik|x−x | .
′ ′
g (+) (x, ω; x′ , t′) = ′
|x − x |
Similarly, in the case of g (−) , we have a pole in the upper half plane at q = −k +iǫ,
and performing the contour integration we obtain,
1 iωt′ ±ik|x−x′ |
g (±) (x, ω; x′ , t′) = e .
|x − x′ |
We now invert the temporal Fourier Transform
(±) ′ 1 Z
′ 1 iωt′ ±k|x−x′ |
G (x, t; x , t ) = dω e−iωt e .
2π |x − x′ |
The ω integration is straightforward, and we find
1 1
" #
(±) ′ ′ ′
G (x, t; x , t ) = ′
δ (t − t) ± |x − x′ | (6.22)
|x − x | c
The Green function G(+) is known as the retarded Green function, because a
change at time t arises from an effect at an earlier time
1
t′ = t − |x − x′ |.
c
Chapter 2 16
The use of the retarded Green function ensures that the observer only feels
the effect of the source after it is turned on.
The use of G(−) means that, once the source ceases, the effects from the
source are no longer felt, or more precisely they are contained within ψout .
Case 1 above is the more commonly encountered, for example in the case ψin ≡ 0
so that there is no wave in the distant past, and a source f (x, t) switches on
at some time. Then inserting our explicit expression for the Green function, we
obtain
f (x′ , t′ret)
Z
ψ(x, t) = d3 x ′
|x − x′ |
where the subscript ret denotes that the function f is evaluate at time
1
t′ret = t − |x − x′|.
c
In this section, we will derive laws expressing conservation of energy and momen-
tum for electric and magnetic fields.
Chapter 2 17
The force acting on a particle carrying charge q, and moving with velocity v is
F = q(E + v × B).
∂u
Z Z " #
3 3
− d x J ·E = dx + ∇ · (E × H) (6.25)
V V ∂t
Since this applies for any volume V , we have a differential energy continuity
equation
∂u
+ ∇ · (E × H) = J · E (6.26)
∂t
The vector
S = E×H
is the Poynting Vector. It only enters through a divergence in the above expres-
sions but, when we come to consider its properties under Lorentz transformations
later in the course, we will discover that it is essentially unique.
We can reduce the integral over the Poynting vector in eqn. (6.25) to a surface
integral using the divergence theory. Thus we can interpret the Poynting vector
as the energy flux across a surface, and the Poynting theorem in essence says:
“The rate of change of electromagnetic energy in a volume together with energy
flux across the boundary is equal to minus the total work done by sources within
the volume”.
Chapter 2 19
The field energy density of eqn. (6.24) contains not only the fundamental fields,
but also the “derived” fields H and D. Thus they include contributions associ-
ated with the polarization and magnetisation of the medium which are in essence
mechanical, and should be associated with the J · E term.
Let Emech be the mechanical energy in some fixed volume V . We have seen
that the work done per unit time per unit volume J · E is the rate of increase of
mechanical energy,
dEmech Z 3
= d x J · E.
dt V
where now we have expressed the field energy solely in terms of the fundamental
fields. It is this expression that is more naturally associated with the field energy,
and Poynting’s theorem reads
d I
(Emech + Efield ) = − dA · S (6.27)
dt
Chapter 2 20
Again we work with the microscopic fields. The force on a particle of charge q
is
F = q(E + v × B).
Thus Newton’s second law may be expressed as
d Z
P mech = d3x [ρE + J × B]
dt
where P mech is the total momentum of the particles in a volume V . To evaluate
this expression, we once again use Coulomb’s law (ME1) and Ampère’s law (ME3),
yielding for the integrand
1 ∂E
" #
ρE + J × B = ǫ0 E(∇ · E) − B × ∇ × B − ǫ0 .
µ0 ∂t
We now use
∂ ∂E ∂B
E×B = ×B+E×
∂t ∂t ∂t
∇·B = 0
to write
ρE + J × B =
∂B ∂
ǫ0 [E(∇ · E) + c2 B(∇ · B) − c2 B × (∇ × B) + E × − (E × B)].
∂t ∂t
We now use Faraday’s law (ME2) to write
d d Z
P mech + ǫ0 d3x E × B =
dt Z dt V
ǫ0 d3x [E∇ · E + c2 B∇ · B − E × (∇ × E) − c2 B × (∇ × B)], (6.28)
where we assume that the volume V is fixed. The second term on the l.h.s. we
associate with the momentum carried by the field
Z
P field = ǫ0 d3x E × B, (6.29)
Chapter 2 21
∇·B = 0
∂B
∇×E+ = 0
∂t
∇·D = 0
∂D
∇×H − = 0.
∂t
In the case of plane waves, it is sufficient to consider those propagating with
a definite frequency ω, and hence time dependence exp −iωt; essentially this is
equivalent to taking the Fourier Transform. We have a set of linear, homogeneous
equations and hence all fields have the same harmonic behaviour. Thus we may
write Maxwell’s equations as
∇·B = 0
∇·D = 0
∇ × E − iωB = 0
∇ × H + iωD = 0.
1
Chapter 2 2
We will now specialise to the case of a linear constitutive relation between the
fields: D = ǫE and B = µH. We will also assume ǫ, µ are real. Note that later
we will consider the complex case; taking them to be real corresponds to there
being no energy losses. Then the last two equations of eqn. (7.1) become
∇ × E − iωB = 0
∇ × B + iωǫµE = 0,
∇2 E + ω 2 ǫµE = 0
∇2B + ω 2 ǫµB = 0 (7.1)
These are known as the Helmholtz wave equations. As is well known, they
support the plane-wave solutions
E E0 ik ·x−iωt
=
e , (7.2)
B B0
√
where k = ω µǫ, and
√
v = ω/k = 1/ µǫ
is the phase velocity.
We now recall the velocity of light in a vacuum is given by
√
c = 1/ µ0 ǫ0 .
E(x, t) = E 0 eikn·x−iωt
B(x, t) = B eikn·x−iωt
0 (7.4)
with
k 2 = µǫω 2 .
Thisis actually shorthand for
E(x, t) = ℜ E0eikn·x−iωt .
n o
n · B0 = 0
n · E 0 = 0. (7.7)
∇ × E − iωB = 0
∇ × B + iωµǫE = 0,
to yield
√
B0 = µǫn × E 0 . (7.8)
√
Setting c = 1/ µǫ to be the volocity of light in the medium, we see that both cB
and E have the same magnitude.
N.B. Had we chosen to work with H, rather than B, then we would have
n × E0
H0 =
Z
q
where Z = µ/ǫ is the impedence
We will now specialise to the case where n is indeed real. Then B 0 is perpendicular
to E 0 , and has the same phase.
Chapter 2 5
ε2
ε1
The vectors E, B and n form an orthogonal triad, and it is usual to introduce three
mutually-orthogonal basis vectors ǫ1 , ǫ2 and n and to write the electromagnetic
field as
Note that E1 and E2 can be complex to account for a phase shift between the two
plane waves.
The general solution for the wave equation is
Linear Polarization
ε2
θ
ε1
If E1 and E2 have the same phase we talk about a linearly polarized wave;
the direction of the E field is constant, with the angle given by
θ = tan−1 E2/E1.
θ=ω t
k
The different signs correspond to rotating to the left or rotating to the right; these
are more commonly known as positive and negative helicities.
Since it is possible to use any two mutually orthogonal vectors as polarization
vectors, it is usually for circularly polarized waves to introduce
1
ǫ± = √ (ǫ1 ± iǫ2 ) (7.10)
2
with the properties
ǫ± ∗ · ǫ± = 1, ǫ± ∗ · ǫ∓ = 0, ǫ± ∗ · n = 0,
An important question is, given an electric field E(x, t), how can one determine
its polarization properties; one way of specifying the relative importance of the
different components is through the Stokes Parameters. This is described in
Jackson 7.2.
Chapter 2 8
The laws describing the behaviour of a wave at the interface between two media
are well known:
2. sin θi/ sin θt = n′/n (Snells’s law) where n′, n are the refractive indices of the
final and initial media respectively.
These are simple kinematic laws; we would like to determine dynamic properties
- intensities and phase changes.
θT kT
kr
ki
θi θr
We begin by writing
We first observe that the boundary conditions must be satisfied ∀x, y at all times
t. Thus all fields must have the same phase factor at z = 0. N.B.: We have
implicitly assumed this in saying that the frequency in z > 0 must be the same
as that in z < 0.
Thus ki · x = kr · x = kT · x at z = 0. The k’s lie in a plane - plane of incidence.
From the figure, we see that
θi = θr
Similarly,
and thus
v
sin θi u ′ ′
uµ ǫ n′
=t =
sin θT µǫ n
Chapter 2 10
D⊥ is continuous
B⊥ is continuous
Applying to the fields at the interface, we have
(E0i + E0R − E0T ) × n = 0 (7.11)
1 1
" #
i r T
(ki × E0 + kr × E0 ) − ′ kT × E0 × n = 0 (7.12)
µ µ
ǫ(E0i + E0r ) − ǫ′ E0T · n = 0
h i
(7.13)
ki × E0i + kT × E0r − kT × E0T · n = 0.
h i
(7.14)
We we now consider two cases; where the electric polarization vector is normal to
plane of incidence and where it is parallel to plane of incidence.
ex
kr
ki
Bi θi θr Br
Ei Er
The z axis is normal to the interface, and we choose the x axis to be in the plane of
incidence as shown. Thus the electric field is along the y axis. The first boundary
Chapter 2 11
µ µ
v
1 i uǫ
u
= E0|ki | cos θi = ω t cos θi E0i .
µ µ
We treat the other two terms similarly, and we find
v
v u ′
uǫ uǫ
u
(E0i − r
E0 )ω t cos θi − ωt ′ cos θT E0T = 0,
µ µ
yielding v
v u ′
uǫ uǫ
u
cos θi t E0i − E0r − cos θT E0T = 0
t (7.16)
µ µ′
The remaining boundary conditions yield no new information, so combining eqns. (7.15)
and (7.16) we find
r
1 − ǫǫµµ′ cos θT
′
E0r cos θi 1 − µµ′ tan
tan θi
θT
i = = µ tan θi (7.17)
r
E0 ǫ ′ µ cos θ
1 + ǫµ′ cos θTi 1 + µ′ tan θT
E0T 2 2
= = µ tan θi (7.18)
E0i
r
1 + ǫǫµµ′ cos
′ θT 1 + µ′ tan θT
cos θi
= = (7.20)
E0i 1 + ǫǫ′ tan
tan θi
r
ǫµ′ cos θT
1 + ǫ′ µ cos θi θT
r
1 − ǫǫµµ′
′
E0r n − n′
= −→ if µ = µ′
E0i
r
1 + ǫǫµµ′
′
n + n′
E0T 2 2
= −→ if µ = µ′
E0i
r
1 + ǫǫµµ′
′
1 + n′/n
In this formula we assume that the directions of E0r and E0i are the same. (contrary
to Eq. (7.42) from Jackson where the directions of E0r and E0i are assumed to be
opposite).
Thus we see that, if both refractive indices are equal
E0r = 0
E0T = E0r
Chapter 2 13
E0r
tan θi
1 − ǫǫ′ tan θT
i = ǫ tan θi .
E0 1 + ǫ′ tan θT
θT kT
kr
ki
θi θr
i0
If light passes from a medium of higher optical density to one of lower optical
density, the angle of refraction is greater than the angle of incidence.
Hence there is a θi for which θT = π/2, given by
We see that the refracted wave propagates parallel to the surface, and is ex-
ponentially attenuated with increasing z. The attenuation occurs over only a
few wavelengths unless θi ≈ i0 .
Chapter 2 15
2
Now HT = (kT × ET )/µ′ ω, and thus
= |ET |2 n · kT /µ′ ω,
whence
1 h
ℜ |ET |2 n · kT /µ′ ω
i
hS · ni =
2
1 h
= ℜ |ET |2 kT cos θT /µ′ ω
i
2
= 0,
since cos θT is purely imaginary; there is no time-averaged energy flux across the
interface.
The principle of total internal reflection has many applications, most notably
in fibre-optic cables. The analysis presented here assumes, of course, that the
material is wide compared to the wave length of light.
Chapter 2 16
7.5 Dispersion
k 2 = µǫω 2 .
E (t)
Nucleus
We now apply an external electromagnetic field (E, B). Then the force on the
electron is
F (t) = −e(E(t) + v × B(t)).
Chapter 2 17
Providing the velocity is small compared to that of light, the magnetic force will
be negligible; recall that c|B| ≈ |E|. Thus the equation of motion of the electron
is
d2 d
m( 2 x + γ x + ω02 x) = −eE(t).
dt dt
The dipole moment of the system is just p = −ex. We now assume that the
external field has frequency ω, so that the time dependence is
E = E0e−iωt.
Thus the displacement will have the same frequency dependence, and we have an
equation of motion
m(−ω 2 − iωγ + ω02 )x = E0 ,
yielding a dipole moment
e2 2
p = (ω0 − ω 2 − iωγ)−1.
m
We now consider the case of N atoms/unit volume, each having Z electrons of
which fj electrons have resonant frequency ωj . We will take this as a model for a
linear medium, in which the polarization P arises solely from the applied external
field. Thus, recalling that
P = ǫ0 χe E
we find
ǫ(ω) N e2 X
= 1 + χe = 1 + fj (ωj2 − ω 2 − iωγj )−1.
ǫ0 ǫ0 m j
with fj = Z. We will rewrite this expression as
P
j
ǫ N e2 X (ωj2 − ω 2) + iωγj
=1+ fj . (7.24)
ǫ0 ǫ0 m j (ωj2 − ω 2 )2 + ω 2γj2
We have thus seen how even a simple model gives a frequency-dependent permit-
tivity.
Chapter 2 18
In general, we can assume that the damping factor γ is small. From the form of
eqn. (7.24), it is clear that at very log frequencies, the susceptibility is positive and
the relative permittivity greater than one. As successive resonant frequencies are
passed, more negative terms contribute and eventually the relative permittivity
is less than one.
Particularly interesting is the behaviour in the neighbourhood of a resonance.
Re ε
Im ε
ω
ωj
Here the real part of ǫ(ω) is peaked around ωj , and furthermore displays anoma-
loous dispersion in which light of higher frequency is less refracted than light of
lower frequency.
The presence of an appreciable imaginary part of ǫ(ω) near ω = ωj represents
absorption; energy dissipated in the medium. To see how this arises, consider a
wave propagating in the z-direction. We will write the wave number as
which gives
2 2 ω2
β − α /4 = c2 ℜ ǫ/ǫ0
ω2
.
αβ = c2 ℑ ǫ/ǫ0
In a conductor, some of the electrons can move freely. Thus there are some
electrons with resonant frequency ω0 = 0, whose contribution to the permittivity
is
N e2 f0
ǫ(ω) = ǫ̃(ω) + i ,
mω(γ0 − iω)
where ǫ̃ represents the background permittivity coming from all the other modes.
We see from this that ǫ(ω) is singular as ω −→ 0, and we will now relate this
property to electrical conductivity.
Our starting point is the Maxwell-Ampère law (ME3):
∂D
∇×H =J + . (7.25)
∂t
We will now impose that J and E are related through Ohm’s law
J = σE
where σ is the conductivity. If we assume the usual frequency behaviour exp −iωt,
and assume the background permittivity is a constant ǫb = ǫ̃(ω), eqn. (7.25)
Chapter 2 20
becomes
σ
" #
∇ × H = −iω ǫb + i E. (7.26)
ω
An alternative procedure is to ascribe all properties, including current flow, to
the dielectric properties of the medium. In that case we have
N e2 f0
Suppost that ω is much larger than the highest resonance frequency. Then we
have
N e2 X (ωj2 − ω 2 ) + iωγj
ǫ/ǫ0 = 1+ fj
ǫ0 m j (ωj2 − ω 2 )2 + ω 2 γj2
ω/ωj ≫1 N e2 X
−→ 1− fj ω 2 /ω 4
ǫ0 m j
= 1 − ωP2 /ω 2 (7.29)
where
N Ze2
ωP2 = (7.30)
ǫ0 m
is the plasma frequency , so called because all the electrons essentially behave
as if free.
Recalling that v
√ 1u
uǫ
k= µǫω = t ω,
c ǫ0
where c is the velocity of light in vacua, we have
q
ck = ω 2 − ωP2
whence
ω(k)2 = ωP2 + c2 k 2 . (7.31)
Such an expression, describing the relationship between wave number and fre-
quency, is known as a dispersion relation . Similar expressions occur in many
places in physics, including special relativity and sound propagation.
In a typical dielectric, when ω 2 ≫ ωP2 , the dielectric constant is slightly less than,
but close to, unity.
In a true plasma, such as the ionosphere, all the electrons are essentially free, and
the expression eqn. (7.29) is valid for a range of frequencies, including ω < ωP .
The wave number k is purely imaginary for frequencies less than the plasma
Chapter 2 22
The above plasma model for the ionosphere is modified considerably through the
presence of the earth’s magnetic field. In the model we now construct, we assume
propagation parallel to the earth’s field B0 . We assume that there is a force acting
on the charges due to a propagating electric field, but that the only magnetic force
is that arising from the earth’s field; recall once again that c|B| ≃ |E|.
Thus the equation of motion for an electron of charge −e and mass m is
d2 x
m 2 = −ev × B0 − eE.
dt
Once again, we consider a monochromatic plane wave with time dependence
e−iωt.
Thus we have
x = x+ǫ+ + x−ǫ− + x3ǫ3 ,
so that the equation of motion becomes
d2 x + d2 x − d2 x 3 dx+ dx− dx3
m[ 2
ǫ+ + 2
ǫ− + 2
ǫ3 ]−eB 0 ǫ3 ×[ ǫ+ + ǫ− + ǫ3 ] = −e[E+ǫ+ +E− ǫ− ]e−iωt.
dt dt dt dt dt dt
First, it is easy to see that since ǫ3 ×ǫ3 = 0, the motion along the Z direction is free:
x3 = x30 + v3t. Since the forces acting in XY plane are periodic, the motion of the
charges in the XY plane will be periodic, too: x+ (t) = x+ e−iωt, x− (t) = x− e−iωt.
Chapter 2 23
Now
1
ǫ3 × ǫ+ = √ (ǫ3 × ǫ1 + iǫ3 × ǫ2 )
2
i
= − √ (ǫ1 + iǫ2),
2
yielding
ǫ3 × ǫ+ = −iǫ+
ǫ3 × ǫ− = iǫ− (7.32)
and thus
ωP2
ǫ± /ǫ0 = 1 − (7.33)
ω(ω ± ωB )
ε-
ω/ωB
ε+
ε-
Thus, in this highly simplified model, we see that the permittivity depends on
the polarization of the incident wave. Indeed, for certain ranges of ω we find that
the permittivity can be negative, and hence one or both polarizations no longer
propagate.
So far we have considered monochromatic waves, but have seen that, if the medium
is dispersive, different frequencies will travel with different velocities. In the sec-
tion, we will describe how, for a general plane wave, the rate of energy flow is in
general different from the phase velocity, or velocity of propagation of a particular
Chapter 2 25
ω ≡ ω(k)
u(x, 0) = eik0 x
yielding
A(k) = 2πδ(k − k0).
In practice we virtually never deal with pure monochromatic plane waves of fixed
frequency k0, but rather with pulses, centred about a frequency k0. In particular,
we will consider the propagation of a Gaussian wave packet, of width ∆x, centred
at x = 0. Then !1/4
1 −x2 /4∆x2 ik0 x
u(x, 0) = e e .
2π∆x2
Chapter 2 26
∆x
This satisfies
Z
2
hx i = dx |u(x, 0)|2x2
1/2 Z
1
!
−x2 /2∆x2 2
= dxe x
2π∆x2
!1/2
1 d Z
2 2
= 2
(−2) 2
dx e−x /2∆x
2π∆x d(1/∆x )
= ∆x2,
By analogy with the width of the wave packet, we see that the amplitude A(k) is
centred at k = k0, with width
1
∆k = .
2∆x
Chapter 2 27
Thus we have the important observation that a short pulse, even of “fixed” fre-
quency k0 , contains a spread of monochromatic components. This expression, of
course, is more familiar from Heisenberg’s Uncertainty Principle.
To see how this spread of frequencies effects the propagation of a wave, we con-
sider the simple case of two monochromatic waves, of the same amplitude and
of neighbouring frequencies (k1, ω1) and (k2, ω2), where k1 , k2 ∼ k0 . Then the
resulting “wave packet” propagates as
= Aei[(k1 +k2 )x/2−(ω1 +ω2)t/2] ei[(k1 −k2 )x/2−(ω1−ω2 )t/2] + ei[(k2 +k1 )x/2−(ω2+ω1 )t/2]
n o
known as the group velocity , and a rapidly moving “phase” with velocity
ω1 + ω2 ω
vp −→ = as k2 → k1 . (7.36)
k1 + k2 k
Since the energy density is associated with the amplitude of the wave, we see
that, in this approximation, energy is transmitted with the group velocity, given
by eqn. (7.35) with k0 the central value of the wave number.
We now recall the relationship between ω and k
ck
ω= , (7.37)
n(k)
Chapter 2 28
where n(k) is the index of refraction, and c is the velocity of light in a vacuum.
The phase velocity can then be written
ω(k) c
vp = = . (7.38)
k nk
This can be either less than or greater than the speed of light; for most media at
optical frequencies, n(k) > 1. We can rewrite the group velocity using eqn. (7.37),
regarding k = k(ω), and find
dn dk
n(ω) + ω = c
dω dω
dω c
=⇒ vg = = dn .
dk k0 n(ω) + ω dω
Providing dn/dω > 0, we have vg < c. However, if dn/dω < 0 (anomalous
refraction), vg can be greater than c.
First, let us recall the propagation of a Gaussian pulse in a linear medium without
dispersion
1 1/4 (x − vt)2
( )
u0(x, t) = ( 2 ) exp − + ik0(x − vt) (7.39)
πL 2L2
√
where L = ∆x 2 is the width of the Gaussian wave packet.
Suppose at t = 0 we switch on the dispersion so that ω = ω(k) (some non-linear
function). What will happen with the pulse?
Starting from t = 0, the solution of the wave eqn is
dkZ
u(x, t) = ℜ A(k) e−iω(k)t+ikx
2π
A solution of the second-order duifferential eqn is specified if we know both u(x, 0)
and u̇(x, 0). It is easy to prove that
Z i
A(k) = dx e−ikx[u(x, 0) + u̇(x, 0)] (7.40)
ωk
Chapter 2 29
The initial condition should be taken from the form of a non-dispersive Gaussian
pulse (7.39) at t = 0:
1 1/4 x2
! ( )
u(x, 0) = u0(x, t) = exp − 2 + ik0 x
πL2 2L
!1/4
1 ivx x2
! ( )
iu̇(x, 0) = iu̇0(x, t) = k0 v + 2 exp − 2 + ik0 x
πL2 L 2L
From Eq. (7.40) we obtain
kv (k − k0 )2L2
" # ( )
2 1/4
A(k) = 4πL 1+ exp − (7.41)
ωk 2
A typical behavior of ω(k) is given by eq. (7.31). For simplicity, we will consider
an approximate model of the behavior of frequency in the vicinity of ω0 in the
form
a2 k 2
ω(k) = ω0 1 + (7.42)
2
where ω0 = vk0 is the center of our Gaussian wave packet.
We obtain
dk kω0 (k−k0 )2 L2
!
2 1/4
Z
u(x, t) = ℜ(4πL ) 1+ e− 2 e−iωk t+ikx
2π k 0 ωk
kω0
The term k0 ωk is approximately 1 in the vicinity of k0 so
and thus
1 Z∞
( )
′
D(x, t) = ǫ0 E(x, t) + dωdt′ eiω(t −t) χe (ω)E(x, t′) .
2π −∞
By a change of variable τ = t − t′ , we can rewrite this as
Z ∞
D(x, t) = ǫ0 E(x, t) + dτ G(τ )E(x, t − τ ) (7.44)
∞
where
1 Z
G(τ ) = dω χe (ω)e−iωτ . (7.45)
2π
We have essentially just used the convolution theorem of Fourier transforms, and
have exhibited the non-local connection between D and E.
Chapter 2 31
-i γ/2 − ν -i γ/2 + ν
To evaluate G(τ ) we use contour integration, noting that there are two cases
1. τ < 0: circle at ∞ vanishes in lower half plane.
sin ν0 τ
= ωP2 e−γτ /2 ,
ν0
and thus
sin ν0τ
G(τ ) = ωP2 e−γτ /2 θ(τ ). (7.46)
ν0
We can make two observations
7.8.1 Causality
Because G(τ ) vanishes for τ < 0, D only depends on the values of E at earlier
times, i.e. Z ∞
D(x, t) = ǫ0 E(x, t) + dτ G(τ )E(x, t − τ ) .
0
We can thus write the dielectric constant as
Z ∞
ǫ(ω)/ǫ0 = 1 + dτ G(τ )eiωτ .
0
We now consider a pole just above the ω-axis, by writing z = ω + iδ. Then
1 1
!
′
= P ′
+ iπδ(ω ′ − ω)
ω − ω − iδ ω −ω
whence
1 Z∞ ǫ(ω ′ )/ǫ0 − 1
ǫ(ω)/ǫ0 = 1 + P dω ′ .
πi −∞ ω′ − ω
Thus taking the real and imaginary parts, we find
1 Z∞ ℑǫ(ω ′ )/ǫ0
ℜǫ/ǫ0 = 1 + P dω ′
π −∞ ω′ − ω
′
1 Z∞ ′ ℜǫ(ω )/ǫ0 − 1
ℑǫ/ǫ0 = − P dω (7.47)
π −∞ ω′ − ω
These are the Kronig-Kramer relations; they relate absorption (imaginary
part of ǫ) to dispersion (real part of ǫ) through analyticity.
Chapter 8
In this chapter we will consider propagation of waves in hollow, metal wave guides
and cavities.
(H2 − H1 ) × n = K
(B2 − B1) · n = 0
(D2 − D1 ) · n = σ
(E2 − E1 ) × n = 0.
Inside a conductor, the electrons are completely free, with infinitely fast response,
such that B = E = 0.
Thus our boundary conditions just below the conducting surface reduce to
H ×n = K
B·n = 0
1
Chapter 2 2
D·n = σ
E × n = 0.
∇ × E = iωB
∇·B = 0
∇ × B = −iµǫωE
∇·E = 0
Note: this does not mean that the propagation vector is in the z direction as such.
We now write
∇2 = ∇2T + ∇2z
where
∂2 ∂2
∇2T = +
∂x2 ∂y 2
2 ∂2
∇z = .
∂z 2
Then our wave equation eqn. (8.1) reduces to
∇ × E = iωB
∇ × B = −iµǫωE. (8.3)
∇T × E T = iωB z (8.4)
∇T × E z + ∇z × E T = iωB T . (8.5)
iω∇z × B T = ∇z × [∇T × E z + ∇z × E T ]
= ∇T [∇z · E z ] − ∇2z E T
Chapter 2 4
∇T × B T = −iµǫωE z
∇T × B z + ∇z × B T = −iµǫωE T .
iω[−iµǫωE T − ∇T × B z ] = k 2E T + ∇T [∇z · E z ],
yielding
We have now shown that the propagation of the waves can be solved solely by
solving the two-dimensional wave equation
Ez (x, y)
(∇2T + µǫω 2 − k 2)
= 0, (8.8)
Bz (x, y)
n × E|S = 0
n · B|S = 0.
Chapter 2 5
Ez = 0 (8.9)
∂Bz
= 0. (8.10)
∂n
Thus we are in principle simultaneously solving two boundary-value equations
subject to each of the above conditions. However, in general the eigenvalue equa-
tion (8.2) will have different eigenvalues for the two different sets of boundary
conditions. Hence we cannot satisfy both simultaneously unless one is trivial.
Thus we classify the solutions as
∂Bz
Ez = 0 everywhere, and ∂n = 0 on boundary. Hre we must solve the Eq. (8.8)b
with Neumann boundary condition. The transverse fields are
iµω ~ T Hz , HT = ik ∇ ~ T Hz
ET = − 2 ê3 × ∇ (8.12)
γ γ2
Finally, we must consider
∇T × E TEM = 0
∇z × E TEM = iωB TEM.
Chapter 2 6
In addition, we have
∇T · E TEM = 0.
Combining the first and third of these equations, we find
∇2T E TEM = 0,
k 2 = µǫω 2 .
Ez = φ(x, y)e±ikz−iωt.
Then ψ satisfies
(∇2T + µǫω 2 − k 2 )ψ = 0,
subject to φ = 0 on the boundary.
We now introduce
γ 2 = µǫω 2 − k 2 ,
Chapter 2 7
(∇2T + γ 2)φ = 0.
If kλ2 > 0, kλ is real, and the propagation is oscillatory. If it is negative, the wave
number is imaginary and the wave will no propagate.
We define the cut-off frequency ωλ by
γλ
ωλ = √ (8.14)
µǫ
where
vp = ω/k
1 1
= √ q
µǫ 1 − ωλ2 /ω 2
c
= q
1 − ωλ2 /ω 2
which is always larger than the velocity of light, and diverges as ω → ωλ .
In contrast, the group velocity
!−1
dk q
vg = = c 1 − ωλ2 /ω 2 ,
dω
Chapter 2 8
which is always smaller than the infinite-space velocity of light, and vanishes as
ω → ωλ . In this limit the wave no longer propagates. Note that
vpvg = c2 .
a
For the sake of illustration, we will consider the case of TE modes. In Cartesian
coordinates, we have to solve the eigenvalue equation
∂2 ∂2
2
+ 2 + γ 2 ψ = 0
∂x ∂y
subject to
∂ψ(0, y) ∂ψ(a, y)
= = 0,
∂x ∂x
∂ψ(x, 0) ∂ψ(x, b)
= = 0.
∂y ∂y
This clearly has eigenfunctions for Hz
mπx mπy
! !
ψmn (x, y) = H0 cos cos
a b
Chapter 2 9
with eigenvalues
2
n2
2 2 m
γmn =π + 2 .
a2 b
We denote the modes TEm,n . The lowest non-trivial mode is TE1,0 if a > b, with
cut-off frequency given by
2
γ10 = π 2 /a2 .
For this mode, for wave propagating in the positive direction, we have
πx ik1,0 z−iωt
!
Hz = H0 cos e .
a
We can obtain the transverse components of the field from eqn. (8.12)
ika πx ikz−iωt
!
HT = − H0 sin e ex
π a!
iωaµ πx ikz−iωt
ET = H0 sin e ey ,
π a
with k = k1,0.
The analysis of TM modes proceeds likewise. However, here the lowest propa-
gating mode is TM1,1, with a higher cut-off frequency. Wave guides are often
constructed such that TE1,0 is the only propagating mode. Recalling that
q
kλ = µǫ (ω 2 − ωλ2 )
√
we can show kλ / µǫω as follows:
Chapter 2 10
ω
TE1,0 TE0,1 TE1,1
TM1,1
ω
8.5.1 Energy Flux along Waveguide
The time-averaged energy flux is given by the real part of the Poynting Vector
1
S= E × H ∗.
2
Let us evaluate this for TE modes
1 1
S = E × H ∗ = (E T × H ∗T − Hz∗ ê3 × E T ).
2 2
Since Hz = ψ(x, y)e−iwt+ikz we get from Eq. (8.12)
ωkµ iωµ ωkµ ωµ
S= 4
∇T Hz ×(ê3×∇T Hz∗ )− 2 Hz∗ ê3 ×(ê3×∇T Hz ) = 4
ê3 |∇t ψ|2 −i 2 ψ ∗ ∇T ψ
2γ γ 2γ γ
Taking the real part, we get
ωkµ
ℜS = 4
|∇T ψ|2 ê3 .
2γ
Chapter 2 11
This is in the z-direction, and we see that energy propagation is along the waveg-
uide.
Similarly, for the TM wave Ez = φ(x, y)e−iωt+ikz one obtains
ωkǫ
ℜS = 4 |∇T φ|2 ê3 .
2γ
The total power transmitted by the TE wave is
Z ωkµ Z
P = ℜ S · ez dA = 4
dA (∇T ψ)∗ · (∇T ψ).
A 2γ
where A is a cross-section through the wave guide. Recalling Green’s identity, we
have Z I
∂ψ
∗ 2
(ψ ∇T ψ + ∇T ψ · ∇T ψ) dA = ψ ∗ dl.
∗
C ∂n
∂ψ
Because of the boundary conditions, either ∂n or φ (for the TM mode) vanish on
the surface. Thus
ωkµ Z ∗ 2
P = − 4 ψ ∇T ψ dA
2γ A
ωkµ 2 Z
= 4
γ |ψ|2 dA,
2γ A
√
r
ω2
where we represented k as ω µǫ 1 − ωλ2 and γ 2 as µǫωλ2 ).
Similarly, for the TM modes we get
1/2
ωλ2
!2
µ ω Z
P = √ 1 − 2 φ∗ φ dA, (8.16)
2 µǫ ωλ ω A
From Chapter 7, we have that the field energy per unit length is given by
1Z 1Z
hU i = [ǫE · E ∗ + µH · H ∗ ] dA = [ǫE T · E ∗T + µH T · H ∗T + µHz · Hz∗ ] dA
4 4
µ 2 2
Z
2 2 µ Z
= 4
(µǫω + k ) [|∇T ψ| + µ|ψ| ] dA = 2
(µǫω + k + γ ) |ψ|2 dA
2 2 2
4γ 4γ
Chapter 2 12
where we have used the fact that |∇T ψ|2 = γ 2 |ψ|2 since ∇2T ψ = −γ 2ψ. Finally,
R R
we obtain
µ2 ǫω 2 Z 2 µ ω2 Z
hU i = 2
|ψ| dA = 2 |ψ|2 dA.
2γ 2 ωλ
Using eqns. (8.15) and (8.5.1), we find
1/2
ω2
1
P/U = √ 1 − λ2 ≡ vg (8.17)
µǫ ω
Thus we see that the energy propagates with the group velocity. N.B. you
should convince yourself that this expression has the correct dimension.
For the TM wave, we get
µ ω2 Z
hU i = 2 |φ|2 dA.
2 ωλ
yielding the same result (8.17) for group velocity.
n·B = 0
n×E = 0
n·D = Σ
n×H = K (8.18)
conductor
H, E
ξ
Hc , E c
n
∇ × Hc = σEc − iωǫEc .
∂
∇ = −n
∂ξ
and our equations become
i ∂Ec
Hc = n×
µc ω ∂ξ
∂ Hc
Ec = − σ1 n × ∂ξ .
Hc = Hk e(i−1)ξ/δ . (8.19)
Chapter 2 15
Thus the magnetic field is tangential and falls off exponentially as we go into
the conductor. We can differentiate this, to obtain
µω
s
Ec = (1 − i)(n × Hk )e−ξ/δ eiξ/δ . (8.20)
2σ
Thus Ec is also tangential to the surface, but of much smaller magnitude.
We now go back to our boundary condition
n × (E − Ec) = 0.
Since E c has a small tangential component, so does E just outside the conductor.
µω
s
Ek = (1 − i)(n × Hk ).
2σ
Thus there is a non-zero component of the Poynting vector into the conductor,
and hence a net flow of energy, given by
dP 1
h i = ℜ [E × H ∗ ] · (−n)
da 2
µc ωδ
= |H k |2
4
1
= |H k |2.
2σδ
It can be demonstrated that this power is dissipated into heat as ohmic losses in
the skin of the conductor.
Applying this to our wave guide, we see that we have an energy loss/unit length
given by
dP 1 I 2 1 I
= − dl |H k | = − dl |n × H|2
dz 2σδ C 2σδ C
∂φ 2
1
1 ω
!2 I
2 2 (TM)
= dl µ 1ωλ ∂n ω2 2
2
ωλ 2
2σδ ωλ C
µǫω 2 1 − ω 2 |n × ∇T ψ| + ω 2 |ψ|
λ
(TE)
λ
A resonant cavity differs from a wave guide in being closed. Thus, rather than
having wave propagation, we have standing waves.
Chapter 2 16
As before, we can have both TM and TE fields. However, now the z-dependence
is of the form, for the case of TM modes,
so ET = − iωµ
γ2
(A sin kz + B cos kz)ez × ∇T ψ. From the boundary conditions
E T |z=0,d = 0 we get
pπz
Hz = ψ(x, y) sin
d
iωµ pπz
ET = − 2 sin ez × ∇T ψ
γ d
pπ pπz
HT = cos ∇ ψ. (8.24)
dγ 2 d T
The function ψ(x, y) now satisfies the wave equation
!2
pπ
∇2T ψ 2
+ [µǫω − ψ =0
d
where
p2 π 2
γ 2 = µǫω 2 − .
d2
We can solve this eigenvalue problem as for propatation along a wave guide, but
now the eigenvalues γλ determine not the cut-off frequencies but the allowed
frequencies:
2 2
2 1 γ 2 +
p π
ωλp = (8.25)
µǫ λ
d2
Chapter 2 18
R y
ψ(s, ϕ) = ψ(s)e±imϕ
ψ(s, ϕ) = Jm(γmns)e±imϕ .
γmn R = xmn,
Chapter 2 19
where xmn is the nth root of Jm(x) = 0. Thus the resonant requencies are given
by
2 2
2 1 x
mn +
pπ
ωmnp = 2 2
(TM mode). (8.26)
µǫ R d
The solution for TE modes is similar and the resonant frequencies are given by
′
2
p2 π 2
2 1 xmn
ωmnp = + 2 (TE mode), (8.27)
µǫ R2 d
Radiating Systems
with
µ0 Z 3 ′ 1 ik|x−x′ |
A(x) = d x J(x′ ) e .
4π |x − x′ |
1
Chapter 2 2
We will now consider the form of the field a distance r away from a localised,
time-varying source of extent d. We begin by introducing the wavelength
2π 2πc
λ= ≡ ,
k ω
where λ ≫ d.
We now consider the form of the potential in three different regions:
−iωρ + ∇ · J = 0.
pck 2 eikr 1
H = − sin θ (1 − )ϕ̂ (9.6)
4π r ikr
ikr
p 2e 1 ik
E = [−k sin θθ̂ + (2 cos θr̂ + sin θθ̂)( 3 − 2 )eikr ]
4πǫ0 r r r
It is interesting to examine their limiting forms
• Radiation Zone: r ≫ λ ≫ d:
• Near Zone: λ ≫ r ≫ d:
Here the leading behaviour of the fields is given by
1 1
E = [3n(n · p) − p] 3
4πǫ0 r
1 i k
H = (n × p) 2 .
4πǫ0 Z0 r
Thus at very short distances, there is essentially an electric dipole field with
time dependence exp −iωt, and a magnetic field suppressed by kr/Z0 that
vanishes as k → 0
In order to show that this solution does indeed correspond to radiation, we will
look at the time-averaged power flux in the radiation zone. This, of course, is
Chapter 2 6
θ=0
The total power transmitted is just obtained by integrating eqn. (9.7) over the
unit sphere, and is independent of the phases of p:
c2 Z0 k 4 2
P = |p| .
12π
Once again we assume that the dimensions of the antenna are much less that the
wavelength. The antenna consists of two conductors of length d/2, along the z
Chapter 2 7
ez
θ
This current flow gives rise to a line charge density Λ through the continuity
equation
∂I
iωΛ(z) = .
∂z
yielding
2iI0
Λ(z) = sgn(z).
ωd
This charge density has a non-zero dipole moment
Z d/2
2iI0
p = dz z ez
−d/2 ωd
iI0d
= ez .
2ω
N.B. if we had current flowing in opposite directions in the two arms of the
antenna, there would have been no dipole radiation term.
Thus, from eqn. (9.7), we see that this apparatus gives dipole radiation, with
power distribution
dP Z0 I02
= 2
(kd)2 sin2 θ
dΩ 128π
Chapter 2 8
Z0I02 (kd)2
P = . (9.8)
48π
If we identify the power radiated with energy dissipation through an effective
resistance, the coefficient of I02 /2 is eqn. (9.8) is the radiation resistance - the
factor of 2 arises from time-averaging, in the usual way.
In this section we’ll derive the formulas for the dipole radiation again - this time
without Fourier transformation dωe−iωt implied.
R
The general formulas for vector and scalar potentials due to an arbitrary source
are:
1 Z 3 ′ ρ(x′ , tr )
φ(x, t) = dx
4πǫ0 |x − x′ |
µ0 Z 3 ′ J(x′ , tr )
A(x, t) = dx (9.9)
4π |x − x′ |
|x−x′ |
where tr = t − c is the retarded time.
To study the behavior of these expressions in the radiation zone |x| ≫ |x′ |, we
choose the origin somewhere inside the radiating body and expand the denomi-
nators in a usual way:
1 1 n̂ · x′
= (1 − + ...) (9.10)
|x − x′ | r r
where r ≡ |x| and n̂ ≡ r̂ is the propagation vector for our would-be shperical
′
wave. We need also to expand the retarded time in powers of rr :
|x − x′ | r n̂ · x′
tr = t − ≃t− +
c c c
so that
′ ′ n̂ · x′
ρ(x , tr ) = ρ(x , t0 ) + ρ̇(x′ , t0 ) + ... (9.11)
c
Chapter 2 9
where t0 ≡ t− rc is the retarded time for our origin. The parameter of the expansion
(9.11) is λd ≪ 1 (see previous Section). Indeed, ρ̇ ∼ ωcharρ where ωchar are the
characteristic frequencies of the emitted radiation, hence dcρρ̇ ∼ dω d
c = λ ≪ 1. )
Substituting the expansions (9.10) and (9.11) in the expression (9.9), one obtains:
1 Z 3 ′ ′ n̂ · x′ ′ n̂ · x′
φ(x, t) = d x [ρ(x , t0 ) + ρ̇(x , t0 )](1 − + ...)
4πǫ0 r c r
Q n̂ · p(t0) n̂ · ṗ(t0)
= + + + ...
4πǫ0 r 4πǫ0 r2 4πǫ0 rc
For the vector potential in Eq. (9.9), the first term in the expansions (9.10) and
(9.11) is sufficient:
µ0 Z 3 ′ J(x′ , tr ) µ0 Z 3 ′
A(x, t) = dx ′
≃ d x J(x′ , t0)
4π |x − x | 4πr
In the previous Section, we demonstrated that
Z
d3x′ J(x′ , t) = ṗ(t)
Let us calculate the magnetic vector potential due to this setup. W.l.o.g. we
can assume that the point x lies in the XZ plane. The general formula for the
magnetic vector potential has the form
′
µ0 I ′ e−iωtr
A(x, t) = dl ′
êφ′ Ie−iωt (9.19)
4π |x − x |
x·x′ r̂·x′
Expanding tr′ ≃ t0 − c (where t0 = t − rc ) and |x−1x′ | ≃ 1r (1 + r ) we get
µ0 bI eikr Z 2π ′ b ′
A(x) = dφ (−ê1 sin φ′ + ê2 cos φ′ )(1 + sin θ cos φ′ )e−ikb sin θ cos φ
4π r 0 r
Since kb = 2π λb ≪ 1 we can expand the exponential in the r.h.s. of this equation
and get
µ0 bI eikr Z 2π ′ b
A(x) = dφ (−ê1 sin φ′ + ê2 cos φ′ )(1 + sin θ cos φ′ − ikb sin θ cos φ′ )
4π r 0 r
Performing integration over φ′ we obtain
ikµ0 Ib2 1 ikr
A(x) = ê2 (1 − )e sin θ (9.20)
4r ikr
Chapter 2 12
For our setup ê2 = êφ so the final result for the vector potential takes the form
ikµ0 m̂ 1 ikr
A(x) = êφ (1 − )e sin θ (9.21)
4πr ikr
which coincides with Eq. (9.18).
Let us find now electric and magnetic fields of the magnetic dipole radiation.
Taking the curl of Eq. (9.18), we find
eikr
1 2 1 ik ikr
!
H= k (n × m) × n + [3n(n · m) − m] − e . (9.22)
4π r r3 r2
The field H due to the magnetic dipole is of the same form as the field E due to
the electric dipole (see Eq. (9.5)). Similarly we have
Z0 2 eikr 1
!
E = − k (n × m) 1− , (9.23)
4π r ikr
so that the electric field due to a magnetic dipole is of the same form as the
magnetic field due to an electric dipole:
ǫ0 m µ0 m
Hmag.dipole = Eel.dipole, Emag.dipole = Hel.dipole ,
p p
Since the radiated power is proportional to n · (E × H),
mag.dipole m2 el.dipole µ0 m2 ω 4
Prad = P = (9.24)
p2 c2 rad 12πc3
In order to get an estimate of the relative strength of the electric and magnetic
dipole radiation, consider a physical dipole made from two charges q and −q
separated by distance d which rotate with angular velocity ω around the center
of the dipole. The magnetic moment of this system can be approximated by an
d2 ω
oscillating current I = Tq = 2πq
ω
so we get an oscillating magnetic moment m = 8
.
The ratio of powers for this example is
Pmag ω 2 d2 v2
∼ = (9.25)
Prmel 4c2 c2
Chapter 2 13
where v is the linear velocity of the rotating charges. We see that for charges
moving with non-relativistic velocities the electric dipole radiation is the most
important part while the magnetic dipole radiation is of the size of the relativistic
corrections.
The interesting part is the quadrupole moment, obtained from the symmetric
part of eqn. (9.17). We use
1Z 3 ′ ′ ′ iω Z 3 ′ ′
d x {(n · x )J + (n · J)x } = − d x ρx (n · x′ ),
2 2
using the same tricks we encountered earlier, and write
µ0 ck eikr 1
!Z
A(x) = − 1− d3 x′ρ(x′ )x′ (n · x′ ). (9.26)
8π r ikr
In the limit r ≫ λ, we find
H = ikn × A/µ0
E = ikZ0(n × A) × n/µ0 . (9.27)
If we now recall our expression for the quadrupole moment
Z
Qαβ = d3xρ(x)(3xαxβ − r2 δαβ )
then we find that H can be written
ick 3 eikr
H=− n × Q(n)
24π r
where Q(n) is defined by
X
Qα = Qαβ nβ .
β
The power dissipation is
dP c2 Z0 6
= 2
k |[n × Q(n)] × n|2 .
dΩ 1152π
We encountered a simple model of a quadrupole moment in the multipole expan-
sion last term:
Q33 = Q0
1
Q11 = Q22 = − Q0, (9.28)
2
Chapter 2 14
The complete description requires the full multipole expansion which is beyond
what I am going to do in this course.
q Z 3 ′Z ′ δ(r′ − w(t)) ~ ′ |r − r′ |
φ(r, t) = d x dt δ(t − t + )
4πǫ0 |r − r′ | c
|r −r′ |
q Z ′ δ(t′ − t + c ) Z ′ 1 δ(t′ − tr )
= dt = dt
4πǫ0 ~ ′ )|
|r − w(t ∂
(t′ − t + |r −w(t
~ ′ )|
) ~ ′ )|
|r − w(t
∂t ′ c t′ =tr
q Z ′ δ(t′ − tr ) c
= dt 1 =
4πǫ0 ~ ′ )| − c v(t′ ) · (r − w(t
|r − w(t ~ ′ ) c|r − w(t
~ r )| − v(tr ) · (r − w(t
~ r ))
∂
where v(t) ≡ ∂t w(t)
~ is the velocity of the particle and tr is the solution of the
equation c(t − tr ) = |r − w(t
~ r )| = 0.
Similarly,
µ0 q c
A(r, t) = v(tr )
4π c|r − w(t
~ r )| − v(tr ) · (r − w(t
~ r ))
The potentials
q c
φ(r, t) =
4πǫ0 c|r − w(t
~ r )| − v(tr ) · (r − w(t
~ r ))
v
A(r, t) = 2 V (r, t) (9.32)
c
are called the Lenard-Wiechert potentials for a point charge. The corresponding
electric and magnetic fields are (see Jackson or Griffiths)
Chapter 2 16
q ς
E(r, t) = 3
[~u(c2 − v 2 ) + ~ς × (~u × a)]
4πǫ0 (~ς · ~u)
ςˆ
B(r, t) = × E(r, t) (9.33)
c
~ς
where v = v(tr ), a = a(tr ), ~ς ≡ r − w(t
~ r ), and ~u ≡ cˆ
ς − v(tr ) (as usually, ςˆ ≡ |~ς| ).
The electric and magnetic fields due to a point charge moving along an arbitrary
trajectory w(t)
~ are given by Eq. (9.33)
q ~u(c2 − v 2) q ςˆ × (~u × ~a)
E(r, t) = +
4πǫ0 ς 2 (ˆ ς · ~u)3 4πǫ0 ς ς · ~u)3
(ˆ
~ς
B(r, t) = × E(r, t). (9.34)
c
where ~ς = ~r − w(t
~ r ), ~u = cˆς − ~v , and tr is defined as a solution to the equation
c(t − tr ) = ς. As usuallly, velocity and acceleration in Eq. (9.34) are taken at
t = tr . The first term (∼ ~u) is called the velocity field and the second (∼ ~a) is
called the acceleration or the radiation field.
The Poynting vector is
~= 1E
S ~ ×B~ = 1 E ~ × (ˆ ~ = 1 [E 2ςˆ − (ˆ
ς × E) ~ E]
ς · E) ~ (9.35)
µ0 µ0 c µ0 c
Some of the energy is radiation; another part is just a field energy carried along
by the particle as it moves. To calculate the power radiated by the particle at
time t∗ , we draw a large sphere with radius ς = R, wait for t − t∗ = Rc , and
integrate Poynting vector over the surface. Since the velocity field is ∼ 1/R2
the corresponding Prad is ∼ R2 R14 = R12 so it does not contribute to the radiated
power at large R. The power due to the acceleration field (∼ 1/R) is finite:
Prad ∼ R2 R12 = 1. We get
→ ~ς × E ~rad = ςˆ E 2 .
~ rad(r, t) = 0 ⇒ S (9.36)
µ0 c rad
For simplicity, consider the charge which is instantaneously at rest at t = t∗ . Since
~v(t∗ ) = 0, ~u(t∗) = cˆ
ς so the Eq. (9.36) reduces to
~ ςˆ µ0 q 2 2 2 µ0 q 2 a2 sin2 θ
Srad = ( ) [a − (ˆ
ς · ~a) ] = ςˆ (9.37)
µ0 c 4πR 16π 2 c R2
The total power is given by the following Larmor formula
I 2 2 Z
~ = µ0 q a
~rad · da sin2 θ 2 µ0 q 2 a2
Prad = S R sin θdθdφ = (9.38)
S 16π 2 cR2 R2 6πc
which we have already obtained using the electric dipole radiation, see the Eq.
(9.16).
We have derived the Larmor formula under the assumption that v = 0 but one can
demonstrate that it holds true as long as v ≪ c. In the general case of arbitrary
velocity, the radiation is given by the Lienard formula
µ0 q 2 γ 6 2 (~v · ~a)2
Prad = (a − ) (9.39)
6πc c2
r
v2
where γ ≡ 1/ 1 − c2 .
and therefore
v
v2 2
u
q u
(c2t − v · r)2 − (c2 − v 2)(c2t2 − r2 ) = Rc 1 −
t
sin θ (9.43)
c2
Fields
For w
~ = tv (and a = 0) the electric field in Eq. (9.33) reduces to (recall ~u ≡
cˆ
ς − v(tr ) = cˆ
ς −v )
q ς~u 2 2 q(c2 − v 2) c~ς − ςv
E(r, t) = (c − v ) = (9.44)
4πǫ0 (~ς · ~u)3 4πǫ0 (cς − v · ~ς)3
Similarly, we get
and
t′ = t
x′ = x − vt (11.1)
1
Chapter 2 2
Suppose now that we look at the equation of motion in K ′ . Then we have v′i =
vi − v, and
dv
mi i = − ∇x′i Vij (|x′i − x′j |).
X
dt j
Now under eqn. (11.1),
∂ ∂
′ =
∂xi ∂xi
and we have
|x′i − x′j | = |xi − xj |,
and we see that the eqn. of motion in K ′ is of exactly the same form as that
in K - we say that classical Newtonian mechanics transforms covariantly under
Galilean Transformations.
We have seen that electric and magnetic propagation in a vacuum satisfies the
wave equation
2
∇2 −
1 ∂ ψ(x, y, z; t) = 0. (11.2)
c2 ∂t2
Let us now consider the transformation of this equation under eqn. (11.1). We
have
∂ ∂x′j ∂ ∂t′ ∂
= +
∂xi ∂xi ∂x′j ∂xi ∂t′
∂ ∂
= δij ′ + 0 = ′
∂xj ∂xi
∂ ∂x′j ∂ ∂t′ ∂
= +
∂t ∂t ∂x′j ∂t ∂t′
∂ ∂
= −vi ′ + ′ .
∂xi ∂t
Thus the wave equation (11.2) becomes
1 ∂ ∂
" ! !#
′2
∇ − 2 ′
− v · ∇′ ′
− v · ∇′ ψ = 0
c ∂t ∂t
Chapter 2 3
1 ∂2
′2 2 ′ ∂ 1
i.e. ∇ −
2 ′2
+ 2
v · ∇ ′
− 2
(v · ∇′ )(v · ∇′ ) ψ = 0 (11.3)
c ∂t c ∂t c
This equation is clearly of a different from to equation (11.2). The wave equation
does not transform covariantly under Galilean Transformations. For sound waves
there is no problem; they propagate in a medium, and it is natural to formulate
the wave equation in a frame in which the medium is at rest. Thus the natural
question arose - Is there a frame in which the “ether” is at rest”?. Of course, we
all know the answer (Michelson-Morley) that the velocity of light is the same in
all frames, and the resolution of this nasty transformation property is the Special
Theory of Relativity.
1. The same laws of nature hold in all systems moving uniformly with respect
to one another.
2. The velocity of light has the same value in all systems moving uniformly
with respect to each other, independent of velocity of observer relative to the
source.
In K ′: c2 t′2 − (x′2 + y ′2 + z ′2 ) = 0
where
β = v/c
γ = (1 − β 2)−1/2
x·v
xk = . (11.8)
|v|
In vector form, this is
ct′ = γ(ct − β · x)
γ−1
x′ = x + (β · x)β − γβct. (11.9)
β2
Chapter 2 6
Alternative Parametrisation
Introduce β = tanh ζ, so that γ = cosh ζ. Then, for frames moving parallel to the
x axis, we have
Given two events (ct1 , x1 ) and (ct2 , x2), Lorentz transformations leave the interval
invariant. Thus we can classify the interval by the sign of δa2 , as follows
• ∆s2 < 0. This is timelike separation. We have c|t2 − t1 | > |x2 − x1 |, so that
the two points can communicate by a signal travelling at less than the speed
of light, and indeed a frame can be chosen such that |x2 − x1| = 0.
• ∆s2 < 0. This is spacelike separation, with c|t2 − t1 | < |x1 − x2 |. The two
space-time points cannot communicate, and indeed a frame exists in which
t1 = t2 .
Points that can be connected with the space-time origin by a light signal are said
to lie on the light cone.
Chapter 2 8
ct
Points within the light cone can be causally connected with the origin, whilst
those outside cannot. The forward (ct > 0) and backward (t < 0) cones define
absolute future and absolute past, and the ordering is preserved under Lorentz
transformations.
Consider a rocket moving with constant velocity v along the x direction relative
to the lab frame K. Let us denote the rest frame of the rocket by K ′. We assume
that the axes of the frames are parallel, and the origins coincide at t = 0.
On the side of a rocket is a meter rule. We also have, in the lab. frame, a high
density of observers, each with a very accurate clock synchronised in the frame
K.
Chapter 2 9
y y’
v
111111111
000000000
11111111
00000000 x
00000000
11111111
00000000
11111111 x’
z z’
Simultaneity
At time t, an observer in the lab frame, co-incident with one end of the meter
rod, records his position (ct, x1 ), and an observer coincident with the other end
does likewise (ct, x2 ). Thus (ct, x1) and (ct, x2) denote two events, which are
simultaneous in the lab. frame.
In the rocket rest frame K ′ we have
We immediately see that t′1 = t′2 iff x1 = x2; in general the points as not simula-
taneous in the rocket rest frame.
Chapter 2 10
Length Contraction
In the rocket frame, our meter rule has length x′2 −x′1 . However, from eqn. (11.11),
we see that in the laboratory frame the length is given by
i.e.
x′1 − x′2
x1 − x2 =
γ
Time Dilation
0 = γ(x − βct2 )
=⇒ x = βct2
t2
t′2 = .
γ
We now generalize the discussion to the case where the rocket is moving with a
velocity v(t) along some path relative to the lab frame K. We will now introduce
K ′ as the instantaneous rest frame of the rocket.
Consider two closely separated points on the trajectory, with coordinates in the
two frames {(ct, x), (c[t+dt], x+dx)} and {(ct′ , x′ ), (c[t′+dt′ ], x′ +dx′ )} respectively.
The interval between the points is the invariant, and we have
c2 dt′2 − dx′2 = c2 dt2 − dx2 .
But dx′ = 0 in k ′ , and furthermore dx2 = v2 dt2 , and thus
q
cdt′ = cdt 1 − β(t)2,
where
v(t)
. β(t) =
c
Then the elapsed time in the rocket between two events is
Z t
2
q
t′2 − t′1 = dt 1 − β(t)2 < t2 − t1 .
t1
The proper time τ is the elapsed time in the frame in with the object is at rest.
Thus
cdτ = ds
where ds is the interval introduced earlier. In this case we have
q
dτ = dt 1 − β(t)2. (11.12)
Note that proper time can only be defined for time-like quantities.
Suppose now that a projectile is fired with velocity u′ from the rocket, relative to
the rocket. Then the co-ordinates of the projectile in K ′ satisfies
dx′
u′ = ′ .
dt
Chapter 2 12
while in K we have
dx
u= .
dt
Using the inverse Lorentz transform we have
xk = γv [x′k + βct′ ]
dx′k dt′ ′
dxk dt
=⇒ uk ≡ = γv ′ + βc
dt dt dt dt
′ ′
dxk
dt
= γv ′ + βc ,
dt dt
where we use k to denote the component along v. We also have
ct = γ[ct′ + βx′k]
dt′ ′
′ dt
=⇒ c = γc c + βuk
dt dt
dt′ c
=⇒ =
dt γv [c + βu′k]
Combinding these two results, we find
u′k + v
uk = (11.13)
1 + βu′k/c
Similarly
dx⊥ dx′ dt′
u⊥ = = ⊥′ · ,
dt dt dt
yielding
u′⊥
u⊥ = . (11.14)
γ(1 + βu′k/c)
In vector notation, this becomes
u′k + v
uk =
1 + v · u′ /c2
u′⊥
u⊥ = (11.15)
γ(1 + v · u′ /c2)
As expected, this reduces to the Galilean result u = u′ + v for the case u′, v ≪ c.
Chapter 2 13
We can formulate this picture in a much more convenient fashion through the
introduction of four vectors. To see how these work, let us return briefly to
Galilean transformations, and rotations in Euclidean space.
Consider two co-ordinate systems P , P ′ whose origins coincide, but which are
related by rotation through an angle θ.
z, z’
1
0
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
00000000000
11111111111
0
1
00000000000
11111111111
y’
0
1
00000000000
11111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
y
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1 00000
11111
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1 00000
11111
00000000000
11111111111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
0
1 00000
11111
00000000000000000
11111111111111111
0000000000000000
1111111111111111
111111111111111111
000000000000000000
00000
11111
00000
11111
00000
11111
x
00000
11111
00000
11111
00000
11111
00000
11111
θ
00000
11111
00000
11111
where R is a rotation matrix. You will note I have put the indices upstairs on
the vectors - I will return to this later. For the specific case of a rotation through
Chapter 2 14
′i ∂x′i j
A = Rji Aj = A (11.17)
∂xj
are called vectors.
A simple example of a vector is dx, which transforms as
∂x′i j
′i
dx = dx ,
∂xj
Scalars
Co-vectors or Forms
X · Y = gij X i Y j .
The tensor gij must be isotropic. There is only one isotropic rank-two tensor:
gij = δij .
Xi = gij X j
X · Y = X i Yi = XiY i .
We only have the luxury of indentifying vectors with covectors in Cartesian co-
ordinates in Euclidean space, where the components of the two are numerically
equal.
Example
x · y = gµν xµ y ν , (11.21)
The covariant four vector or form can be obtained as before by using the
raising and lowering properties of the metric tensor
xµ = gµν xν .
In our example we have that xµ = (ct, −x, −y, −z) - the components of a covector
are numerically different to those of the vector.
Chapter 2 17
Let us return to our two frames K and K ′. The relation between vectors in the
two frames is given by
′µ ∂x′µ ν
x = ν
x = Lµν xν (11.23)
∂x
Let us assumed a similar transformation law for the covectors
x′µ = Lµν xν .
where
1 if ν = σ
δν σ =
(11.25)
0 if ν 6= σ
Note that
∂xσ
Lµσ = , (11.26)
∂x′µ
the characteristic transformation property of a form.
Thus the various quantities we will encouter in the remainder of this course are
• Contravariant Vectors:
A′µ = Lµν Aν
• Covariant Vectors:
Bµ′ = Lµν Bν
• Tensors:
′ ′
C ′µ νρ′... ρ σ
′ ′
µ ν µν...
σ ′ ... = L µ L ν . . . Lρ′ Lσ . . . C ρσ...
Chapter 2 18
• Scalars:
A · B = Aµ B µ = gµν AµB ν
11.6.5 Derivatives
c2 dτ 2 = ds2 ,
Chapter 2 19
α dxα
v = (11.30)
dτ
we have
1 d
vα = √ (ct, x) = γ(c, v),
1 − β 2 dt
yielding
v α = (γc, γv), (11.31)
whose spatial components clearly reduce to our familiar definition of velocity in
the non-relativistic (NR) limit.
= m2 γ 2c2 γ −2 = m2 c2 .
Thus we have
pµ pµ = p2 = m2 c2 (11.33)
For a particle at rest, we have perhaps the most famous equation in physics.
The use of four-vectors is essential to solve problems in special (and general. . . )
relativity. Whilst simple kinematical problems can be solved using three vectors,
it is very clumsy indeed.
J µ = (ρc, J) (11.36)
∂µ J µ = 0.
However, it remains to be shown that the J µ thus constructed does indeed trans-
form as a four vector.
Consider J µ defined through eqn. (11.36) under a transformation to a frame K ′
moving with velocity v along the x axis. Then, if J µ we indeed a four vector we
would have
v
" #
′
ρ c = γ ρc − Jx
c
′
Jx = γ [Jx − vρ]
Jy′ = Jy
Jz′ = Jz .
In the NR limit
′
J = J − ρv
,
ρ′ = ρ
as expected.
Consider now the case Jx = 0. Then we have
Jx′= −γvρ
′
.
ρ = γρ
The second equation would appear to violate charge conservation. However, let
us consider what happens to a volume element under this transformation. In the
frame K, we have
dV = dx dy dz.
Chapter 2 22
However
dx = γ(dx′ + v dt′ )
v
dt = γ(dt′ + 2 dx′ )
c
′
dy = dy
dz = dz ′ .
dV = dx dy dz = γdx′dy ′ dz ′ = γdV ′ ,
Thus both the charge densities and volumes are not separately conserved under
this Lorentz transformation, but the chage itself is.
There is much experimental evidence that ρ′ = γρ, and we will postulate that
J µ in eqn. (11.36) is indeed a four vector, and that
∂µ J µ = 0 (11.37)
11.8.2 Units
At this point, Jackson changes from SI to Gaussian units - the aim being to avoid
carrying superfluous factors of c. In my youth I did everything in SI units, and
then in units in which c ≡ 1 (a huge simplification!). But to avoid confusion (!),
I will also make the switch so as to be in keeping with Jackson.
Gaussian Units
∇ · D = 4πρ (11.38)
Chapter 2 23
4π 1 ∂D
∇×H = J+ (11.39)
c c ∂t
1 ∂B
∇×E + = 0 (11.40)
c ∂t
∇·B = 0 (11.41)
D = ǫE = E + 4πP (11.42)
H = B/µ = B − 4πM (11.43)
You will notice that in these units ∂/∂t has an associated factor of 1/c, correspond-
ing to our definition of a four vector. Also, ǫ and µ are the relative permittivity
and permeability respectively.
B = ∇×A
1 ∂A
E = −∇φ − (11.44)
c ∂t
In a vacuum (ǫ = µ = 1), the inhomogeneous equations become:
1 ∂∇ · A
∇2 φ +
= −4πρ
c ∂t
1 ∂ 2A 1 ∂φ 4π
" #
2
∇ A− 2 2 −∇ ∇·A+ = − J.
c ∂t c ∂t c
In the Lorentz gauge, we have
1 ∂φ
∇·A+ = 0,
c ∂t
and the dynamical equations become
2 1 ∂ 2φ
∇ φ − 2 2 = −4πρ
c ∂t
1 ∂ 2A 4π
∇2 A − 2 2 = − J. (11.45)
c ∂t c
Chapter 2 24
We now recognise the operator on the l.h.s. of these equations as the four-dimensional
Laplacian introduced in eqn. (11.29), and the r.h.s. as the temporal and spatial
components of the current J µ of eqn. (11.36). We will therefore introduce a four-
vector potential
Aµ = (φ, A), (11.46)
so that both equations in (11.45) can be unified in the manifestly covariant form
4π µ
2Aµ = J . (11.47)
c
Furthermore, the Lorentz gauge condition is also manifestly covariant:
∂ µAµ = 0. (11.48)
Chapter 2 25
Ei = Ei
Ei = −E i .
We can see that (E, B) are related to a second-rank tensor, and there are six
independent components of the two fields.
For a general second-rank tensor T µν , we can write
µν
T µν = Tsym
µν
+ Tanti−sym .
The symmetric part has ten components, but the anti-symmetric part has the six
independent components that we could associate with fields E and B. Thus we
introduce the anti-symmetric Maxwell Field-Strength Tensor
F µν = ∂ µAν − ∂ ν Aµ (11.49)
Chapter 2 26
We see that E and B are not components of four vectors, but rather of an anti-
symmetric, second-rank tensor. Note that we can lower the indices in the usual
way
Fµν = gµα gνβ F αβ ,
so that the components corresponding to E change sign, whilst those correspond-
ing to B are unaltered.
Finally, we will introduce the dual field-strength tensor. But as a precursor we
will return to the Levi-Civita tensor.
Levi-Civita Tensor
This is the four-dimensional version of the ǫijk encountered in 3-D Euclidean space.
It is defined by
1 if µ, ν, ρ, σ is an even perm of 0, 1, 2, 3
ǫµνρσ =
−1 if µ, ν, ρ, σ is an odd perm of 0, 1, 2, 3 (11.51)
0 if any two indices are equal
ǫµνρσ = −ǫµνρσ .
E −→ B
B −→ −E,
so that
0 −Bx −By −Bz
Bx 0 Ez −Ey
F̃ µν = .
By −Ez 0 Ex
Bz Ey −Ex 0
Thus F̃ µν reverses the roles of the electric and magnetic fields.
Finally, using eqn. (11.52), we have
µν 1
F̃˜ = ǫµνρσ F̃ρσ
2
1
= ǫµνρσ ǫµνλτ F λτ
4
= −F µν (11.54)
∇ · E = 4πρ (11.55)
1 ∂B
∇×E + = 0 (11.56)
c ∂t
4π 1 ∂E
∇×B = J+ (11.57)
c c ∂t
∇ · B = 0. (11.58)
Chapter 2 28
These are all first-order differential equations expressed in terms of E and B. Thus
we might suspect that the covariant form of Maxwell’s equations will contain terms
of the form
∂µ Fνρ .
Looking at eqn. (11.55), we see that it may be written
∂ i J0
E = 4π .
∂xi c
i i0 00
Recalling that E = F , and noting that F vanishes, we can rewrite (11.55) as
4π 0
∂µ F µ0 = J . (11.59)
c
Turning now to the second inhomogeneous equation, eqn. (11.57), we see that it
may be written
∂ 4π i 1 ∂ i
ǫ0ijk j F̃ k0 = J + E,
∂x c c ∂t
where we use B k = F̃ k0 . To put this equation in a form analogous to eqn. (11.59),
we perform a clever piece of manipulation:
ǫ0ijk F̃ k0 = ǫk0ij F̃k0
1
= ǫµνij F̃µν
2
ij
= F̃˜
= −F ij
where in the second line we have used that one of µ or ν must be the temporal
component, and the other a spatial component. Thus eqn. (11.57) can be written
∂ 4π i 1 ∂ i0
− j F ij = J + F
∂x c c ∂t
∂ ji ∂ 0i 4π i
=⇒ F + F = J (11.60)
∂xj ∂x0 c
4π i
=⇒ ∂µ F µi = J. (11.61)
c
Thus we see that both the inhomogeneous Maxwell equations can be written in
the unified form
Chapter 2 29
4π ν
∂µ F µν = J (11.62)
c
Turning now to the homogeneous equations, we see that eqn (11.58) can be written
∂ i0
F̃ = 0
∂xj
=⇒ ∂µ F̃ µν = 0.
Thus the two homogeneous Maxwell equations can be written in the unified form
∂µ F̃ µν = 0 (11.63)
Eqns. (11.62) and (11.63) constitute the covariant formulation of Maxwell’s equa-
tions.
Note that we can rewrite eqn. (11.63) as
1 µνρσ
∂µ ǫ Fρσ = 0
2
=⇒ ǫµνρσ ∂µ Fρσ = 0,
dτ = γ −1dt,
and write
dpi dpi dτ 1 dpi
= = .
dt dτ dt γ dτ
Thus the force law may be expressed as
dpi 1
( )
= γq E i + ǫ0ijk v j B k .
dτ c
We now introduce the four-velocity V µ = (γc, γv), yielding
dpi q
= {V0F i0 + ǫ0ijk V j B k }
dτ c
q
= {V0F i0 + ǫ0ijk Vj F̃k0 }
c
q
= {V0F i0 + F ij Vj } (using eqn. (11.54)).
c
Thus the Lorentz force law becomes
dpi q
= Vµ F iµ . (11.65)
dτ c
The analogous equation for the energy is
d mech dt d mech
E = E = qE · v.
dt dτ dτ
Thus we have
dE mech
= γqF i0v i
dτ
= qF 0i Vi ,
Chapter 2 31
yielding
E mech q
d = Vµ F 0µ.
dτ c c
Identifying E mech/c with the component p0 , we see that both this equation and
the Lorentz law, eqn. (11.65), can be expressed as
dpµ q
= Vν F µν (11.66)
dτ c
Lorentz Invariants
There are two invariants we can construct from the field-strength tensor
1.
Thus
1
E 2 − B 2 = Fµν F µν (11.67)
2
is a Lorentz Scalar.
2.
Thus
−4E · B = Fµν F̃ µν (11.68)
is a Lorentz Scalar
′ γ2
E = γ[E + β × B] − β(β · E)
γ+1
γ2
B ′ = γ(B − β × E) − β(β · B), (11.71)
γ+1
where β = v/c. Thus the E and B fields mix under a Lorentz transformation.
Chapter 2 34
Consider a charge q moving along a line at velocity (in K) v = ve1 . The charge
is at rest in the frame K ′.
y y’
1
0 1
0
0
1 0
1
0000000
1111111
0
1 1111111
0000000
1
0 0000000
1111111
P 0
1
0
1
1
0
0000000
1111111
1111111
0000000
0000000
1111111
0000000
1111111
v
111111111
000000000
0
1 1111111
0000000
0
1
1 0000000
1111111
0
0
1 0000000
1111111
1
0 1111111
0000000
0
1 0000000
1111111
b 0
1
1
0
0
1
1
0
1111111
0000000
0000000
1111111
0000000
1111111
1111111
0000000 θ
0
1
0
1
1
0
r
0000000
1111111
1111111
0000000
0000000
1111111
0
1 0000000
1111111
1111111
1
0 0000000
0
1
0
1
0000000
1111111
1111111
0000000
1111111
0000000
1111111
0000000
x
x’
z z’
At t = t′ = 0, the origins of the two frames co-incide. We have an observer P at
impact parameter b (i.e. distance of closest approach) as shown above.
We will begin by looking at electric and magnetic fields at point P in frame K ′
at time t′ .
P has coordinates
x′ = −vt′
y′ = b
z ′ = 0.
E3′ = 0.
We now use our transformation laws eqn. (11.70) to write
qγvt
E1 = E1′ = − 2
(b + v 2 γ 2t2 )3/2
γqb
E2 = γE2′ = 2
(b + v 2γ 2t2 )3/2
E3 = γE3′ = 0
B1 = 0; B2 = γB2′ = 0
B3 = γβE2′ = βE2
Thus in the laboratory frame we see a magnetic induction.
Note that in the limit v −→ c, we have β −→ 1 and the magnetic induction equals
the transverse electric field. In the Galilean limit v −→ 0,
v γqb vqb
B3 = −→
c (b2 + v 2 γ 2t2 )3/2 c(b2 + v 2 t2 )3/2
qv ×r
=⇒ B ∼
c r3
where we have used vb = vr sin θ, which we observe is just the Biot-Savart Law.
Finally, let us look at the field lines. We have that
E2 b
=− ,
E1 vt
Chapter 2 36
so that the electric field is still a central field in the frame K. If we now look at
the magnitude of the field, however, we find
γq
|E| = 2 2 2 2 3/2
(b2 + v 2t2 )1/2.
(b + v γ t )
Setting b = r sin θ, vt = −r cos θ, we have
γqr γq
|E| = 3 2 2 3/2
∼ 2
(1 + γ cos2 θ)3/2.
r (1 + γ cos θ) r
So the lines of force, whilst central, are no longer isotropic - they are predomi-
nantly transverse in strength.
Let us look at the propagation of a plane wave in vacuum. Our starting point is
the Jacobi identity eqn. (11.64). Applying ∂ α we find
2Fµν = 0 (11.73)
k 2 = ω 2 /c2
Introducing θ as the angle between k and v, we can use the second eqn. of (11.76)
to compute the Doppler shift:
ω′ ω v ω v ω
" # " #
=γ − cos θ|k| = γ − cos θ .
c c c c c c
Thus we have the Doppler Shift formula
where β = v/c. This is modified from the usual Galilean formula through the
factor of γ.
Chapter 2 38
11.11.1 Aberation
This is the change in direction of a wave vector between the two frames.
k k’
θ θ’
v
v
We can calculate this from
′ |k ′⊥ | |k ⊥|
tan θ = ′ = ′ .
kk kk
sin θ
tan θ′ = (11.78)
γ(cos θ − β)
Summary of First Three Chapters
Introduction
• Coulomb’s Law:
1 r̂21
F21 = q1 q2
4πǫ0 |r2 − r1 |2
• Principle of Linear Superposition
The resultant force on a test particle due to several charges is the
vector sum of the forces due to the charges individually.
In the case of a continuous charge distribution, these two principles yield
1 Z ρ(r′ )(r − r′ )dV ′
E(r) = .
4πǫ0 V | r − r′ |3
Gauss’ Law
If charge density ρ(r) is the sole source of the electrostatic field E(r), the flux of
E out of a closed surface S bounding a volume V is given by
Z
Q
E · dS = 4πCQ = in SI units
S ǫ0
where Q = total charge within S. This can be expressed in differential form as
Maxwell’s First Equation (ME1)
∇ · E = ρ/ǫ0 .
39
Chapter 2 40
Scalar Potential
E(x) = −∇φ(x),
∇2 φ(x) = −ρ(x)/ǫ0
Uniqueness Theorem
This is perhaps the most important theorem in the course, which we will use
implicitly in much of the following.
φ(x) = constant
E · n = σ/ǫ0.
Chapter 2 41
• Method of Images
Here we introduce an image charge outside the region we are seeking a solu-
tion such that image system satisfies boundary conditions of original problem.
Two crucial things to check
– Only additional charges are introduced outside the region so that Pois-
son’s equation is unaltered.
– Image system satisfies correct boundary conditions.
subject to
4π S=∂V ∂n ′ ∂n′
Green functions can be obtained using, e.g., method of images, or in terms
of orthonormal eigenfunctions.
Chapter 2 42
Spherical Harmonics
+ + 1− 2 R =0
dx2 x dx x
• Solutions are the Bessel Functions Jν (x) and Nν (x), which are linearly
independent. A further set is provided by Hankel Functions.
′
p(x ) ′
+ q(x′)g(x, x′) = −4πδ(x − x′ ),
dx dx
defined on the interval x′ ∈ [a, b], with homogeneous boundary conditions at
a and b.
Chapter 2 45
• Green function
4π y2 (x)y1(x′)
a ≤ x′ ≤ x
−
p(x) W [y1(x), y2(x)]
g(x, x′) =
4π y2 (x′)y1(x)
x ≤ x′ ≤ b
−
p(x) W [y1(x), y2(x)]
where
W [y1(x), y2(x)] = y1(x)y2′ (x) − y2 (x)y1′ (x).
is the Wronskian, and y1 and y2 are solutions to the homogeneous equation
′ X ψn∗ (x)ψn(x′ )
G(x, x ) = 4π
n λn − λ