Electrodynamics
Electrodynamics
J.E. Avron1
May 5, 2013
1
Comments and typos welcome. Send to [email protected].
2
I thank my TAs Dr. Dana Levanony and Mr. Yaroslav Pollak for all their
help and Prof. Amos Ori and Dr. Oded Kenneth for all they taught me. The
students the class in 2012, especially Itai Schlesinger, who pruned many typos
and Daniel Klein who showed me how to query Wolfarm α.
Contents
1 Tensor calculus 9
1.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.1.1 Euclidean geometry . . . . . . . . . . . . . . . . . . . . . 9
1.1.2 The metric tensor . . . . . . . . . . . . . . . . . . . . . . 10
1.1.3 Einstein summation convention . . . . . . . . . . . . . . . 11
1.1.4 Coordinate transformations . . . . . . . . . . . . . . . . . 11
1.2 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.1 Covariants components . . . . . . . . . . . . . . . . . . . 14
1.2.2 Orthogonal coordinates . . . . . . . . . . . . . . . . . . . 15
1.2.3 Contraction makes scalars . . . . . . . . . . . . . . . . . . 16
1.3 Scalars, vectors, tensors, densities . . . . . . . . . . . . . . . . . . 16
1.3.1 Symmetric and anti-symmetric tensors . . . . . . . . . . . 16
1.3.2 Weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.3 Levi-Civita symbol . . . . . . . . . . . . . . . . . . . . . . 17
1.3.4 Volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.3.5 Areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4 Mirror . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.5 Isometries of Euclidean space . . . . . . . . . . . . . . . . . . . . 19
1.6 Tensorial equations are coordinate free . . . . . . . . . . . . . . . 20
1.7 Differential operators . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.7.1 Grad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.7.2 Div . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.7.3 Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.7.4 Laplacian . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.8 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2 Minkowski space-time 25
2.1 The principle of relativity . . . . . . . . . . . . . . . . . . . . . . 25
2.1.1 Causality . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2 Space-time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.1 Cartesian coordinates and Inertial frames . . . . . . . . . 28
2.2.2 Events, world lines, light cones, etc. . . . . . . . . . . . . 29
2.3 Everything is relative . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3.1 Time dilation . . . . . . . . . . . . . . . . . . . . . . . . . 30
3
4 CONTENTS
4 Variational principle 57
4.1 The action begins . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.1.1 Action for a free particle . . . . . . . . . . . . . . . . . . . 58
4.1.2 Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.1.3 Gauge invariance . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 Variation of the action . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2.1 Variation of Sp . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2.2 Variation of Sint . . . . . . . . . . . . . . . . . . . . . . . 61
4.2.3 Euler-Lagrange equation . . . . . . . . . . . . . . . . . . . 61
4.2.4 The non-relativistic limit . . . . . . . . . . . . . . . . . . 61
4.2.5 Existence and uniqueness . . . . . . . . . . . . . . . . . . 62
4.2.6 Consistency check . . . . . . . . . . . . . . . . . . . . . . 63
CONTENTS 5
4.3 Supplement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.3.1 Fermat principle . . . . . . . . . . . . . . . . . . . . . . . 63
4.3.2 Rainbow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3.3 Geodesics in Curved space-time . . . . . . . . . . . . . . 65
5 Maxwell Equations 69
5.1 Technical preliminaries . . . . . . . . . . . . . . . . . . . . . . . . 69
5.1.1 Space-time volume element . . . . . . . . . . . . . . . . . 69
5.1.2 Densities . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.1.3 Distributions . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.2 Charge densities and currents . . . . . . . . . . . . . . . . . . . . 72
5.2.1 4-current-density . . . . . . . . . . . . . . . . . . . . . . . 72
5.2.2 Charge conservation . . . . . . . . . . . . . . . . . . . . . 73
5.2.3 Current conservation and gauge invariance . . . . . . . . 74
5.2.4 Gauge invariance and the continuity equation . . . . . . . 75
5.3 Lagrangian field theory . . . . . . . . . . . . . . . . . . . . . . . 75
5.3.1 The Lagrangian of the electromagnetic field . . . . . . . . 76
5.4 Variation of the field: Rules of the game . . . . . . . . . . . . . . 77
5.4.1 Variation of the field: Calculations . . . . . . . . . . . . . 77
5.4.2 Variation of the interaction . . . . . . . . . . . . . . . . . 78
5.4.3 The inhomogeneous Maxwell equations . . . . . . . . . . 78
5.4.4 Current conservation . . . . . . . . . . . . . . . . . . . . . 79
5.4.5 3-D form . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.4.6 Time reversal . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.4.7 Maxwell equations: Structure . . . . . . . . . . . . . . . . 80
5.5 Dielectric and magnetic media . . . . . . . . . . . . . . . . . . . 81
5.6 New Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.6.1 The quantum Hall effect and Chern-Simon action . . . . . 83
5.6.2 Axion electrodynamics . . . . . . . . . . . . . . . . . . . 85
5.6.3 Quantum interface . . . . . . . . . . . . . . . . . . . . . . 86
5.6.4 Magnetic response to an electric field . . . . . . . . . . . . 88
5.6.5 Phantom monopoles . . . . . . . . . . . . . . . . . . . . . 88
5.6.6 Electrodynamics in 1+1 dimensions . . . . . . . . . . . . 90
6.2.1 Blah . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6.2.2 Variation of the metric in mechanics . . . . . . . . . . . . 100
6.2.3 Variation of the metric . . . . . . . . . . . . . . . . . . . . 101
6.2.4 Matrix calculus . . . . . . . . . . . . . . . . . . . . . . . . 101
6.2.5 The stress tensor . . . . . . . . . . . . . . . . . . . . . . . 102
6.3 Energy momentum conservation . . . . . . . . . . . . . . . . . . 102
6.3.1 The source term j · F and conservation of energy . . . . . 103
6.3.2 The interpretation of T jk : Stress . . . . . . . . . . . . . . 104
6.3.3 Field lines as rubber bands . . . . . . . . . . . . . . . . . 105
6.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.4.1 Radiation pressure . . . . . . . . . . . . . . . . . . . . . . 106
6.4.2 Solar sails . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.4.3 Halbach array . . . . . . . . . . . . . . . . . . . . . . . . . 108
9 Radiation 147
9.1 Wave equation with a source term . . . . . . . . . . . . . . . . . 147
9.1.1 Retarded waves for an arbitrary source . . . . . . . . . . . 147
9.1.2 Application: Lorenz gauge . . . . . . . . . . . . . . . . . . 147
9.1.3 Scalar wave generated by a moving point source . . . . . 148
9.2 Maxwell equation in the Lorenz gauge . . . . . . . . . . . . . . . 149
9.3 Lienard-Wiechert: Retarded potentials . . . . . . . . . . . . . . 150
9.3.1 The Lorenz Gauge condition . . . . . . . . . . . . . . . . 150
9.4 Retarded Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
9.4.1 Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . 153
9.5 Particle instantaneously at rest . . . . . . . . . . . . . . . . . . . 154
9.5.1 The Magnetic field: . . . . . . . . . . . . . . . . . . . . . 154
9.5.2 The electric field . . . . . . . . . . . . . . . . . . . . . . . 155
9.5.3 Slow particles . . . . . . . . . . . . . . . . . . . . . . . . . 156
9.6 Retardation from a distant source . . . . . . . . . . . . . . . . . 156
9.6.1 The dipole approximation: k` 1 . . . . . . . . . . . . . 157
9.6.2 Beyond dipole: Slow motion . . . . . . . . . . . . . . . . . 159
9.7 Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
9.8 Classical instability of atoms . . . . . . . . . . . . . . . . . . . . 160
9.9 Orienting an antenna . . . . . . . . . . . . . . . . . . . . . . . . . 161
9.10 Reciprocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Tensor calculus
1.1 Geometry
Minkowski space-time is the stage on which electrodynamics takes place. Before
describing the geometry of space-time lets us collect the tools we need from
(Riemanian) geometry.
Remark 1.1 (Euclid, Gauss and Einstein). Euclid took it for granted that
physical space is Euclidean. The first to seriously entertain the possibility that
the physical space need not be Euclidean was Gauss. In a Euclidean world the
angles of all triangles sum up to π. So, when one says that the world is to a good
approximation Euclidean one means that the deviations from π are small. Gauss
who had experience in lad surveying made an experiment which was did not show
deviation from Euclidean geometry. Later Einstein taught us that space-time is
actually curved and there are many physical tests of this. However, this is a
another story.
9
10 CHAPTER 1. TENSOR CALCULUS
Figure 1.1: Cartesian, rotated Cartesian and polar coordinates for the Euclidean
plane
g is called the metric tensor, also known as the Riemann metric tensor who in
his Thesis founded Riemanian geometry. It is a second rank tensor which means
it has two indices. It is also symmetric gij = gji . The components are written
downstairs. Downstairs components are called covariant components.
If the two Cartesian coordinate systems use the same yardstick then we may
choose the length scale of xj so that
gij = δij
With this choice g is dimensionless and dxj have dimension of length.
In polar coordinates, x1 = r cos θ, x2 = r sin θ
(d`)2 = (dr)2 + r2 (dθ)2
Exercise 1.2. Verify.
The moral of this is that there are many distinct metrics for a given space: As
many as coordinate transformations. The surface of the sphere is geometrically
distinct from the plane. It has the metric
(d`)2 = (dθ)2 + sin2 θ (dφ)2
The geometric distinction means that it can not accommodate Cartesian coor-
dinates. This is why maps drawn on a sheet of paper never accurately represent
regions earth.
Exercise 1.3. Compute det g in Cartesian and polar coordinates.
1.1. GEOMETRY 11
0.7
0.6
0.5
0.4
0.3
0.2
0.1
Figure 1.2: The plot on the left shows a Cartesian y-mesh. The plot in the
middle shows a curvilinear y-mesh in the plane. The deformation is due to a
deformation of the radial polar coordinate r → r(tanh r) shown on the right.
gja v a = gjb v b
Remark 1.5 (Warning). If you get an equation where the indices are not nicely
paired, such as
va ua , va ua wa
it is a good idea to search for a typo.
This says
∂xi
(g 0 )ab = Λi a Λj b gij , = Λj b gji Λi a , Λi a = (1.1)
∂(x0 )a
1
If one thinks of Λ and g as matrices the relation above can be written as
g 0 = Λt gΛ (1.2)
1 First index is row second is column.
12 CHAPTER 1. TENSOR CALCULUS
1.2 Vectors
Consider a vector δx (say, in the Euclidean plane) associated with a small change
of the coordinates δxj . For the sake of masochism, we allow non-orthogonal
coordinate system.
e2
v
v2
v1
e1
the vector is a scalar, and this aloows us to relate the basis vectors ej with the
metric X
δx · δx = gij δxi δxj , gij = ei · ej (1.4)
We recover that fact that g is symmetric and positive
You also see form this formula that g is diagonal in orthogonal coordinate sys-
tems and that the covariant basis vectors ej are, in general, not unit vectors.
Exercise 1.9. Show that in a polar coordinate system the covariant basis vectors
and the normalized unit vectors are related by
er = r̂, eθ = rθ̂
Remark 1.10. Normalized unit vectors are defined only for orthogonal coordi-
nate systems.
so the transformation law for the contravariant components is, by the chain rule
∂xj
j
δx = δ(x0 )a ⇐⇒ δxj = Λj a δ(x0 )a ⇐⇒ δx = Λδ(x0 ) (1.6)
∂(x0 )a | {z }
Contravariant
v = (rθ̇) θ̂ = θ̇ eθ (1.7)
v1
e1
From the definitions of the metric tensor, and the notion of duality we get
vk = va ea · ek = v a ea · ek = gka v a (1.11)
We learn from this that the metric tensor allows us to push indexes down.
Exercise 1.13. Show that
ej = (ej · ea ) ea
Everything one can do with contravariant components has an analog in the
covariant components. In particular, the length of a vector is evidently given
by
v · v = va vb g ab , g jk = ej · ek (1.12)
Taking the scalar product of exercise 1.13 with ek we conclude that the two
metric tensors are inversely related
δkj |{z}
= ej · ek |{z}
= (ej · ea )(ea · ek ) = g ja gak =
|{z} gj k
duality ex.10 index gym
1.2. VECTORS 15
The metric tensor with indexes up is the inverse matrix of the metric tensor
with indexes down. It raises indexes since
From this one can figure the transformation rules for the covariant components
(v 0 )a = (g 0 )ab (v 0 )b
= (g 0 )ab (Λ−1 v)b
= Λi a Λj b gij (Λ−1 )b k v k
= Λi a Λj b (Λ−1 )b k gij v k
= Λi a (ΛΛ−1 )j k gij v k
= Λi a gij v j
= Λi a vi
Remark 1.14. If you think of the covariant components as a row vector then
the transfromation rule is
0
|v ={zvΛ}
covarinat
Note the similarities and differences with the rule of transfromation of the
contravariant components.
er · er = 1, eθ · eθ = r2 , er · eθ = 0
r r θ θ −2
e · e = 1, e ·e =r , er · eθ = 0
r̂ · r̂ = 1, θ̂ · θ̂ = 1, r̂ · θ̂ = 0
16 CHAPTER 1. TENSOR CALCULUS
u · v = u0 · v0
v j = Λj k (v 0 )k , (v 0 )j = vk Λk j (1.15)
Tensors are multi-index objects and the metric tensor is an example. The num-
ber of indices is called the rank of the tensor. Each index transforms according
to whether it is up or down. For example, the second rank tensor T j k transform
like v j uk .
1.3.2 Weights
There are interesting physical quantities that are neither scalars nor tensors.
det g is and example. From Eq. 1.2
2
det g 0 = det Λ det g
Objects with such a rule of transformation are called weights. det g has weight
−2.
1.3. SCALARS, VECTORS, TENSORS, DENSITIES 17
Remark 1.19. The conventions for spherical coordinates is such that you get
a right handed frame provided or order the coordinates r, φ, θ, i.e.
εrφθ = 1
Remark 1.20. In even dimensions cyclic permutations are odd, while in odd
dimensions cyclic permutations are even.
εij... εij... = n!
1.3.4 Volume
In a 3-dimensional Euclidean space, with Cartesian coordinates xj the volume
element is
dV = dx1 dx2 dx3
and, of course, dx1 dx1 dx2 has no volume. If we also allow volume to take a
sign according to the handedness of the frame, we naturally associate volume
element with the components of the completely antisymmetric tensor:
and ∧ means that order matters: dxi ∧ dxj = −dxj ∧ dxi . In Cartesian coordi-
√
nates, where g = 1, this can be re-written as
√
√ 1 2 3 g
dV = g dx ∧ dx ∧ dx = εijk dxi ∧ dxj ∧ dxk (1.16)
3!
√
We have seen in Ex.1.18 gεijk are the covariant components of a bona-fide
(completely anti-symmetric third rank) tensor. Similarly, dxi ∧ dxj ∧ dxk are
the conravariant components of a completely anti-symmetric third rank tensor.
Contracting the two we get that dV behaves like a scalar under coordinate
transformation. Hence, Eq. 1.16 holds in any curvilinear coordinate system,
(provided 1, 2, 3 is right handed frame). This works in any dimension.
Remark 1.23. It is useful to assign signs to volume (and areas): Positive for
right handed frames and negative for left handed frames.
Exercise 1.24 (Spherical coordinates). Let (x, y, z) be Cartesian coordinates
off Eucliden space with metric d`2 = (dx)2 + (dy)2 + (dz)2 . Show that the usual
spherical coordinates
1.3.5 Areas
The area element dS3 associated with the parallelogram (dx1 , dx2 ) is naturally
defined as
dV = dS3 dx3 = (dS3 e3 ) ·(dx3 e3 )
| {z }
area vector
Exercise 1.25. Verify the formula for the area for spherical area elements dSr .
1.4. MIRROR 19
Figure 1.5: The area element, black arrow, directed along e3 , is defined so that
its scalar product with the red arrow, e3 , gives the volume. The black arrow is
perpendicular to the blue arrows e1 and e2 .
1.4 Mirror
A mirror flips some things and does not flip others. It does not flip up and down
but does flip right and left, it flips a right handed coordinate system to a left
handed one etc.
The length of an object is a scalar. It does not flip in a mirror. However,
when one integrates it makes a sign difference if you go right or left. So, for
example, you want the volume element to flip a sign when reflected in a mirror.
For the sake of concreteness consider the 3-dimensional Euclidean space.
Under coordinate inversion (x0 )j = −xj the contravariant components of say
a position or velocity vector flip sign. (The vector still points in the same
direction.) Now consider the angular velocity. since you use your right hand
to determine the direction of the vector of angular velocity, in a mirror you’ll
use your left hand. The vector of angular velocity now points in the opposite
direction. It is a pseudo vector. Under inversion its components do not flip sign.
The cross product of two vectors is therefore a pseudo-vector. Since
(a × b)i = εijk aj bk
we conclude that the Levi-Civita symbol is also pseudo.
This is what is meant by saying that vectors in Cartesian do not have a location.
Rotation keeps the origin fixed. This is also the case for general linear
transformation,
(x0 )j = Λj a xa (1.17)
A linear transformation is a symmetry provided it leave the metric invariant.
In Cartesian coordinates g = 1 and by Eq. (1.2)
1 = (g0 ) = Λt gΛ = Λt 1Λ = Λt Λ
This says that Λt is the inverse of Λ:
Λt Λ = 1
det Λ2 = 1 =⇒ det Λ = ±1
Example 1.27. In a two dimensional Euclidean space there are two second
rank tensors that are invariant under rotations
1 0 0 1
, (1.19)
0 1 −1 0
T jk... = 0
holds in one (fixed) coordinate system, it hold in any other coordinate system.
For example, Newton’s equation
f j = maj
Fjk = ∂j Ak − ∂k Aj
1.7.1 Grad
The chain rule
∂ ∂xk ∂ ∂
0
= = Λk j k
∂(x ) j ∂(x0 )j ∂xk ∂x
says that partial derivatives behave like covariant components of a vector. In
particular, if φ(x) is a scalar valued function then ∇φ give the components of a
covariant vector field.
Exercise 1.30. Show that ∇φ in spherical coordinates is
Here you see why covariant components often lead to simpler formulas than
normalized coordinates.
1.7.2 Div
We shall show that
1 √
∇ · E = √ ∂j ( gE j ),
g
22 CHAPTER 1. TENSOR CALCULUS
Consider a smal cube in the coordinates dxj . The putative expression for div
indeed satisfies Gauss law
√
dV (∇ · E) = gdx1 dx2 dx3 (∇ · E)
√ f
= dx2 dx3 g E 1 + . . .
i
3√ 1
f
2
= dx dx ge · E 1 e1 + . . .
i
√ f
= g dx2 dx3 e1 ·E + . . .
| {z } i
dS
= dS · E
1
∂r (r2 sin θE r ) + ∂θ (r2 sin θE θ ) + ∂φ (r2 sin θE φ )
∇·E =
r2
sin θ
1 1
= 2 ∂r (r2 E r ) + ∂θ (sin θE θ ) + ∂φ (E φ )
r sin θ
1 1
= 2 ∂r (r2 Er̂ ) + ∂θ (sin θEθ̂ ) + ∂φ (Eφ̂ )
r r sin θ
1.7.3 Curl
The last differential operator we shall need to discuss is the curl:
εijk
(∇ × E)i = √ ∂j Ek
g
The putative formula for curl indeed gives Stokes for dS a small square dx1 ×dx2 .
dS · (∇ × E) = dS3 (∇ × E)3
ε3ij
√
= gdx1 dx2 √ ∂i Ej
g
= dx1 dx2 (∂1 E2 − ∂2 E1 )
f
= dx2 E2 − . . .
i
f
= (dx e2 ) · (E2 e2 ) − . . .
2
i
f
= (dx2 e2 ) · E − . . .
i
= d` · E
2
Example 1.32. The φ components of curl in spherical coordinates (recall
Remark1.19 ) is :
1
(∇ × E)ϕ = (∂θ Er − ∂r Eθ )
r2 sin θ
and in normalized components
1
(∇ × E)ϕ̂ = (∂θ Er − ∂r Eθ )
r
1
= ∂θ Er̂ − ∂r (rEθ̂ )
r
Exercise 1.33. Compute the (∇ × E)r and (∇ × E)r̂ .
∇ × (∇φ) = 0
∇ · (∇ × E) = 0
1.7.4 Laplacian
The Laplacian of a scalar function is defined by
1 √ 1 √
∆φ = ∇ · ∇φ = √ ∂j ( gg jk ∂k φ) = √ ∂j ( g ∂ j φ)
g g
∇ × (∇ × E) = −∆E + ∇(∇ · E)
2 Note that Wolfram Mathematica notation compares with mine by interchanging ϕ ↔ θ
24 CHAPTER 1. TENSOR CALCULUS
1.8 Bibliography
• S. Weinberg, Gravitation and Cosmology, gives all a physicist needs to
know about tensors.
• B. Schutz
• Flanders
Chapter 2
Minkowski space-time
Remark 2.1 (c is large). Human length scale, say the length of the foot, is
` = 1[meter].
p When you walk the foot behaves like a pendulum, and its period
is 2π `/g ≈ 2 [s] with g ≈ 9.8 [m/s2 ]. A human speed is then ≈ 1 [m/s].
On this scale c ≈ 3 × 108 [m/s] is essentially infinite. I do not know who first
entertained the thought that c may be finite but large, but the first to estimate c
from astronomical data was the Danish astronomer Rømer (1644-1710) .
25
26 CHAPTER 2. MINKOWSKI SPACE-TIME
2.1.1 Causality
Causality is a way to distinguish past from future: We remember the past but
can’t change it and we can affect the future but do not know it. When one takes
into account the fact that c is a constant of nature, one finds that one needs
to reconsider what one means by the notion of “future” , “now” and “past”.
Einstein said that if you could break the speed of light, you could use this to
send a message to your dead grandmother. An alternative characterization of c
is therefore the ultimate speed at which information propagates.
2.2 Space-time
Space-time is the stage on which events happen. An event, like my typing this
text, is something that happens in space and time and is labeled by 4-coordinates
xµ , µ ∈ 0, 1, 2, 3 where x0 = ct with t time. It is natural to give space and time
the same dimension. This is what we do when we say that the sun is about 8
light-minutes away from earth.
Remark 2.2. We shall use the conventions that Greek indices µ, ν run from 0
to 3, while Roman indices j, k from 1 to 3.
t t'
Future x'
Past
Figure 2.1: Space-time: The red disk is an event. The t axis is the world line
of a black clock that sits at the origin x1 = x2 = x3 = 0. The x axis represents
a Euclidean 3-space. The blue line is the world line of a blue clock moving at
constant speed c/2. In the frame of the blue clock t0 is the time axis. The red
lines represent the future and past light cones relative associated with a signal
that is emitted (or absorbed) at the origin. The Euclidean distance on the paper
reflects badly (d`)2 in Minkowski space.
2.2. SPACE-TIME 27
with g a real, symmetric tensor. The notation is taken from geometry where
(ds)2 ≥ 0 but is potentially misleading because we allow (ds)2 to have either
sign: When the two events are simultaneous in an inertial frame the interval is
indeed the distance squared and (ds)2 > 0. When two events are separated in
time, but occur at the same point the interval is the time elapsed on a clock
squared, but (ds)2 < 0. Let us motivate this choice of signs.
We can always choose coordinates so that g is diagonal at a given point.
(Take Λ an orthogonal transformation.) By scaling we can then make the diag-
onal entries ±1. If all the entries are +1 we have a something that looks locally
like 4-dimensional Euclidean space, not space-time. Space-time has three spa-
tial coordinates, naturally associated with the three entries +1. The remaining
time coordinate is different. It comes with the −1. This gives a 3+1 space-time
manifold1 .
2GM
(ds)2 = − (1 − Φ) (cdt)2 + (1 + Φ) dx · dx, Φ(x) =
c2 r
Far from the star, the clock rate coincides with the coordinate time rate.
2V
Exercise 2.4. Compute the correction terms to Minkowski, c2 on earth due
the sun. (Answer: Φ = 2 × 10−8 )
28 CHAPTER 2. MINKOWSKI SPACE-TIME
Figure 2.2: Two Cartesian coordinates in Minkowski space. In both the t axis
and the x axis are Minkowsi orthogonal: The light-cone (red) bisects the angles
between the t and x axes. If you have installed Mathematica cdf player you can
view a simulation courtesy of Slava Pollak.
1
t0 = γ(1, v/c, 0, 0), γ=p ≥1
1 − (v/c)2
What is the direction of the x0 of the blue inertial frame? Since the blue frame
is inertial, x0 and t0 must be Minkowski orthogonal and the contravariant com-
ponents of x0 must be
1 This would also be the case with three −1 and one +1.
2.3. EVERYTHING IS RELATIVE 29
Now is relative
Fig. 2.4 shows that notion of now is relative. In the black coordinate system
events separated by x occur simultaneously, in the blue ones events separated
by x0 are simultaneous.
30 CHAPTER 2. MINKOWSKI SPACE-TIME
b
S
O
a
Figure 2.3: The interval between the two space like events O and S is related to
the clock readings of an inertial clock. The red lines in the figure are light-like
vectors and a and b are the time intervals measured by the clock at O.
t t'
x'
Figure 2.4: The black coordinate system is inertial. So is the blue coordinate
system. In both the t and x axes are Minkowski orthogonal. The two red
dots are two events which are simultaneous in the black inertial frame but not
simultaneous in the blue frame. The light cone is a coordinate independent
entity.
nized. (Hint: What happens to time delay if you half the speed and double the
travel time?)
(v/c, 1)`
(since the length of the rod, measured by x1 , in the frame where it is stationary
is `). The interval, being a scalar, relates the proper length ` with the apparent
32 CHAPTER 2. MINKOWSKI SPACE-TIME
ct ct¢
x¢
1
Figure 2.5: The blue frame moves to the right at u and the black frame moves
to the left at −u. Because of the symmetry, the length on the paper along t
and t0 correctly reflect the interval. The thick cyan line is one second measured
by a black clock. Its end point is on the line t = 1. The line of events t0 = 1 in
the blue frame intersects the cyan line: The moving clock is slow– Time dilates.
ct ct¢
x¢
1
Figure 2.6: The thick blue interval is a meter stick at rest in the black frame.
The black frame moves left and the blue frames righht at equal speeds. This
makes the Euclidean lengths along the x and x0 axes proportional to the interval.
The blue thick interval is longer than the green interval: The meter stick has
shorter length in the blue frame. Moving objects contract.
2.3. EVERYTHING IS RELATIVE 33
length dx0
`2
(dx0 )2 =, γ≥1
γ2
You are biggest in your own rest frame.
Exercise 2.8. Earth actually rotates pretty fast. To get and idea compute Ωρ/c
at the equator . (Answer: 1.5 × 10−6 )
In the case of earth the centrifugal correction can normally be neglected, but
the Sangac term is important and to a good approximation
Exercise 2.9 (Coordinate times and clock times). 1. Compare the change in
coordinate time dt in the rotating earth frame with the proper-time dτ
measured by a clock at a fixed location in the rotating frame
2. Compare the change in coordinate time dt0 in the inertial frame with the
change in coordinate time dt in the rotating coordinates
3. A clock is taken for a one year trip around earth equator. Show that the
time lag relative to a clock that stayed is
2AΩ
∆τ = ± , A = πRe2
c2
34 CHAPTER 2. MINKOWSKI SPACE-TIME
t t'
Φ'
Φ
Figure 2.7: The coordinates in the inertial frame are drawn blue and the
coordinates in the rotating frame black. The light cone in red. The gray line is
the synch line in a local inertial frame.
where Re is the earth radius and the ± depends on whether the trip was
towards the east or towards the west.
5. Explain why the result implies that one can not synchronize clocks on earth.
Light-cone coordinates
Light cone coordinates in Minkowski space are
√ √
2u = x − ct, 2v = x + ct, y = y, z=z
Exercise 2.10 (Metric in light cone coordinates). Show that the Minkowski
metric tensor in light-cone ordinates is
0 1 0 0
1 0 0 0
ηµν = η µν =
0 0 1 0 ,
0 0 0 1
2.4. LORENTZ TRANSFORMATIONS 35
u v
Figure 2.8: Light-cone coordinates. This is not a Cartesian frame since u·v = 1.
Rindler coordinates
Rindler coordinates are the analog of polar coordinate system in a two dimen-
sional Minkowski space time
x = ρ cosh τ, t = ρ sinh τ
η 0 = Λt ηΛ = η (2.5)
det Λ = ±1
36 CHAPTER 2. MINKOWSKI SPACE-TIME
(x0 )µ = xµ + aµ
This gives
∂(x0 )µ
Λµ ν = = δµν ⇐⇒ Λ = 1
∂xν
which is the trivial Lorentz transformations. This expresses the homogeneity of
Minkowski space time.
ηL + Lt η = 0 (2.6)
The Λ(t) genertaed in this way is a Lorentz transformation for any t. This
follows from the following exercises
2.4. LORENTZ TRANSFORMATIONS 37
Exercise 2.12. Use Eq. 2.6 and the differential equation to show that det Λ(t) =
1.
1. if λ is an eigenvalue of Λ so is λ∗ .
2.4.3 Rotations
Rotation by θ about the x-axis and its generator are given by
1 0 0 0 0 0 0 0
0 1 0 0 0 0 0 0
Λyz (θ) =
Lyz =
0 0 cos θ sin θ 0 0 0 1
0 0 − sin θ cos θ 0 0 −1 0
Similarly for rotations about the x, y space axes. The isometry expressses the
isotropy of space in of Minkowski space-time.
Remark 2.15. An airplane has three rotation controls: Stick, for pitch, rudder
for yaw, and ailerons for roll. The three are lineraly independnet, but non-
linearly dependent: You can always generate one from the other two.
2.4.4 Boosts
Lorentz transformations relating different inertial frames are called as boosts.
A boost in the x direction and its generator are
cosh φ sinh φ 0 0 0 1 0 0
sinh φ cosh φ 0 0 1 0 0 0
Λtx (φ) = , Ltx = (2.7)
0 0 1 0 0 0 1 0
0 0 0 1 0 0 0 1
If you remember your quantum mechanics, then an easy way to see that Eq. (2.6)
is to recall that for Pauli matrices σx σz + σz σx = 0. and so are isometries of
Minkowski space. The isometry expresses the fact that Minkowski space looks
the same in all inertial frames.
Exercise 2.18. Show that the commutator of two boosts is a rotation:
2.4.5 Rapidity
φ of Eq. (2.7) is called the rapidity. To related the rapidity to the relative
velocity between the frames consider the origin in the primed frame x0 = 0.
0 = (x0 )1 = Λ1 µ xµ = −(sinh φ) ct + (cosh φ) x
From this we conclude that
v = c tanhφ (2.8)
vc
Rapidity
Exercise 2.19 (Rapidity add). Show that rapidities (in the same direction) add
Λ(φ1 )Λ(φ2 ) = Λ(φ1 + φ2 )
Exercise 2.20 (Galilean transformations). Show that for small rapidities Lorentz
boosts reduce to Galilean transformation:
t0 = t + O(c−2 ), x0 = x − vt + O(c−2 )
2.5. 4-VELOCITY 39
2.5 4-Velocity
The proper time dτ is a Lorentz scalar and is non-zero for a clock that travels
slower than light. In this case we can define the velocity as a 4-vector
dxµ
uµ = (2.9)
dτ
The length of u is always -c2 , essentially by definition,
dxµ dxµ
uµ uµ = = −c2 (2.10)
(dτ )2
The 4-velocity is a time-like vector. It lies in the forward light cone. It is related
to the usual velocity by
dxµ dxµ dτ uµ
(c, v) = = =
dt dτ dt γ
Remark 2.21 (Old fashioned velocities). Whereas the 4-velocity transforms like
any 4-vector under Lorentz transformations, the usual velocity has complicated
(bad) transformation properties. This is because both the numerator and the
denominator are components of a vector.
It is convenient to think of the path xµ (τ ) as parametrized by proper time.
Since the 4-velocity is normalized we can always write it as
u = c (cosh φ, n sinh φ), n·n=1
n is the direction, which may depend on τ and φ = φ(τ ). Evidently
γ = cosh φ
from which is follows that φ is the rapidity defined in the previous section as
v = c tanh φ.
Exercise 2.22. Show this
2.5.1 4-momentum
Define the 4-momentum
pµ = muµ = mγ(c, v) = (E/c, p)
The associated scalar is
pµ pµ = −(mc)2
It is a time-like vector for massive particles. This encapsulates the most famous
equation in physics, Einstein equation for the energy in the rest frame, E = mc2 .
Remark 2.23 (Tachyons). Tachyons are (fictitious) particles with whose mass
is (fittingly) imaginary. The 4-velocity is space like, uµ uµ = (mc)2 > 0. A
Tachyon can not be a point particle: there is a Lorentz frame where the particle
is spread on a line at fixed t. If Tachyons would interact with ordinary particles
(tardyons), you could go back in time and kill your grandfather.
40 CHAPTER 2. MINKOWSKI SPACE-TIME
2.5.2 4 Acceleration:
The 4-acceleration can be similarly defined as
duµ
aµ = (2.11)
dτ
It is always Minkowski orthogonal to the velocity
uµ aµ = 0 (2.12)
(Since the Minkowski length of the velocity is −1). It follows that The 4-
acceleration is always a space like vector.
Example 2.24 (Constant acceleration). The 4-velocity along some fixed direc-
tion is
u = c (cosh φ, n sinh φ)
where φ(τ ) is a parametrization of the path and n is a fixed unit vector. The
4-acceleration is then
a = c φ̇(sinh φ, n cosh φ)
Evidently
aµ aµ = c2 φ̇2
Constant acceleration g corresponds to linear dependence of φ on the proper-time
τ
gτ
φ(τ ) =
c
The red lines in Fig. 2.9 show paths of constant acceleration.
Remark 2.25 (Coincidence). An amusing phenomenological coincidence is that
the year, the gravitational acceleration on earth g, and c are simply related:
g × year/c = 1.03
Exercise 2.26. What fraction of the velocity of light would you reach in this
case. Answer: tanh 1.03 = 0.77
Exercise 2.27 (Space travel). You may worry that since c is the ultimate speed
man, living for, say 80 years, can explore at most a smallish neighborhood of 80
light-years around earth. This is wrong. A space traveler who lives for n years,
in a space ship which accelerates at g will travel a distances, cosh n, measured
in light years. The visible universe has radius of about 1011 light years. So you
will get there is about 26 years.
2.6. GPS 41
Example 2.28 (Motion with fixed velocity but changing direction). The 4-
velocity of a particle moving with fixed velocity but changing direction is
uµ = c(cosh φ, n(τ ) sinh φ)
where φ is fixed. The acceleration is
aµ = c sinh φ(0, ṅ)
Exercise 2.29. Find the orbit xµ (τ ) describing a plane circular motion with
constant angular velocity.
2.5.3 Horizons
t t¢
S
x
O
Figure 2.12: An inertial observer that lives long enough sees all the events in
Minkowski space-time
Figure 2.13: An accelerating sees only half the events in Minkowski space-time.
He never sees the black dot on the left. The red line is his horizon.
2.6 GPS
Every time you use your GPS and find, with relief, that the GPS really knows
where you are, you are testing special and general relativity. It took a century
to turn Einstein’s conceptual revolution into a palm gadget.
42 CHAPTER 2. MINKOWSKI SPACE-TIME
GPS works like that: There are about 24 GPS satellites orbiting earth at
a radius of about 26, 000 [km] and period of about half a day. Their orbit are
known (and monitored) with great precision (few centimeters). On each satellite
there is an atomic clock that measures the proper time with great precision.
Each satellite radios the coordinates of the event of transmission: The b satellite
radios the event Xbµ (τb ) and the event of reception if xµ . Since the transmission
is by electromagnetic wave that propagate at c, the two events are light-like. To
determine the four unknown coordinates xµ of teh reception event you need 4
equation. You need to see 4 GPS satellites and record 4 transmission events all
light-like separated from you. This gives 4 equations with 4 unknowns.
The GPS system is sophisticated and involved: It takes into account special
and general relativity; atmospheric effects on the veldocity off light, and the fact
you insist on having your location in a non-inertial coordinate system attached
to a rotation earth. The point I want to make here is that relativistic corrections
are large: Ignoring relativity would degrade the the accuracy to about 10 km
and make GPS useless. Relativity is regularly and routinely tested. If you want
to know more about that, then the article of Neal Ashby in Living Reviews of
General Relativity is a good place to learn. Wikipedia is, as usual, quite good
as well.
Instead, I will consider a caricature of GPS to illustrate one idea. Namely,
how one can use 4-clocks, with known position, and the fact the speed of light
is a constant of nature, to determine the reception event in space time.
• Compute the difference between the coordinate time and the self-time of a
GPS clock after one day. ( ∆t ≈ πRv/c2 ≈ 3.6 × 10−6 [s])
Exercise 2.31 (GPS in 1+1 dimensions). Two satellites with known orbits
, a0b (τ ), a1b (τ ) , b = 1, 2, emit signals at τa and τb respectively. Assuming that
aµ − bµ is space like, show that light-cone intersect at
2x1 = ±(a01 − a02 ) + a11 + a12 , 2x0 = (a01 + a02 ) ± (a11 − a12 )
ct
a
b
Figure 2.14: The world line of the two satellites are the blue lines. The intersec-
tion of the light cones is the event whose coordinates we seek. Since the orbits
of the satellites are known, the events (a0 , a1 ) and (b0 , b1 ) are known given the
proper times τa and τb .
Bibliography
1. B. F. Schutz, A first course in general relativity, geometric exposition of
relativity.
2. Neal Ashby in Living Reviews of General Relativity
Chapter 3
In space-time the electric and magnetic fields (E, B) are amalgamated into a
second rank, anti-symmetric, tensor, Fµν . Fµν is derived from the 4-potential
Aµ . This leads to the homogeneous Maxwell equations and a neat way to write
Coulomb and Lorentz laws.
The electric field is that part of the force which is independent of the velocity
of the particle and the magnetic field is the part of the force is linear in the
velocity. Of course, the partition into electric and magnetic field is different in
different inertial frames. Observers in different inertial frames do not agree on
the partition. E and B mix under change of frame.
Remark 3.1 (SI). Eq. (3.1) appears to say the magnetic forces are a relativistic
correction to the electric forces. This is misleading because it depends on what
values we take for the wo fields. In Gaussian units the unit of electric field is
300 [V /cm] and the unit of magnetic field is Gauss. In SI units the unit of
electric field is much five orders of magnitudes smaller, 1 [V /m], and the unit
45
46 CHAPTER 3. THE ELECTROMAGNETIC TENSOR
f = e (E + v × B)
The force of electric field of 1 [volt/cm] and the magnetic force of 1 [Gauss]
have comparable magnitudes at velocities of 1000 [km/sec].
1
E0 = E + v × B, B0 = B (3.2)
c
The mixing of E and B under the change of inertial frames suggests that
they come from a single entity in space-time. This entity is not a 4-vector since
we need 6 slots and a 4-vector has too few. It is not a general second rank
tensor, since it has too many components–16. It is not a symmetric second rank
tensor since this too has too many components–10. However an anti-symmetric
rank 2 tensor
Fµν = −Fνµ (3.3)
This leaves us with the problem of how to put the two Euclidean vectors
(E, B) in Fµν . We need to identify slots in F that behave like vectors. Assuming
Cartesian coordinates, write
0 x1 x2 x3
−x1 0 y3 −y 2
Fµν = (3.4)
−x2 −y 3 0 y1
−x3 y2 −y1 0
(The identity on the right expresses the fact that Euclidean rotations correspond
to orthogonal transformation). The (covariant) components
F0j = xj
transfrom by
Rotations of Fjk
Now consider the triplet 2y j = εjmn Fmn . The Levi-Civita symbol is invariant
under (proper) rotations since det R = 1. Hence
This gives
j
(2y 0 )j = (Rt )k (2y)k
= Rj k (2y)k (3.8)
One sees that x and y both transform as vectors. More precisely, x is a vector
while y is a pseudo-vector as the transformation rule relied on the use of Levi-
Civita (and det R = 1).
We can use these facts to identify the slots for E and B. The force f is a
vector. By the Coulomb law E must be a vector. In contrast, the Lorentz force
involves a cross product, which uses a hand rule: B is a pseudo-vector.
F = Fµν eµ ⊗ eν
Exercise 3.7 (Coulomb in spherical coordinates). Using the rules of tensor
calculus, show that F for Coulomb field in spherical coordinates has F0r =
−Fr0 = re2 and all other components are 0.
2 And also on the covariant form of Couloms-Lorentz law, see the next section
3.1. AMALGAMATING E AND B 49
Example 3.8. The covariant components of the field tensor in cylindrical co-
ordinates (ct, ρ, φ, z) are
0 −Ex c − Ey s ρ(−Ey c + Ex s) −Ez
... 0 ρBz −By c + Bx s
Fcylind =
...
... 0 ρ(Bx c + By s)
... ... ... 0
Exercise 3.9 (Magnetic field of currnet line). A line of current I along the
z-axis creates a magnetic field B = 2I
cρ θ̂ in cylindrical coordinates. Show that
in cylindrical coordinates Fρz = −Fzρ = 2I cρ and all other componnets vanish.
(Hint: It is simpler to use properties of the basis vectors eρ , ez rather than
compute the transformation matrix.)
Exercise 3.10. Explain why the last identity is true and why it is desirable.
∇·B =0
says that there are no magnetic monopoles. In Euclidean space this statement
is equivalent to the fact that B is derived from a vector potential A:
B=∇×A
and this says that E + ∂tcA is the gradient of a scalar potential −φ. E and B
are therefore derived from potentials
1
B = ∇ × A, E = −∇φ − ∂t A (3.13)
c
Fµν = ∂µ Aν − ∂ν Aµ (3.15)
In components:
1
Ej = Fj0 = ∂j A0 − ∂0 Aj = − ∂t Aj − ∂j φ,
c
3 In SI units c is absorbed in B.
3.2. THE ELECTROMAGNETIC POTENTIAL 51
and
B i = 21 εijk Fjk = (∇ × A)i
We have reproduced Eq. 3.13.
Exercise 3.11. Show that if A is a 4-vector field then F defined through the
4-potential transforms like a second rank tensor under arbitrary (possibly curvi-
linear) coordinate transformation.
xΜ
xΝ
Figure 3.1: Loop and surface element for Stokes.
where dS is the area element spanned by the loop. Both sides are manifestly
Lorentz scalars, and the right hand side is manifestly gauge invariant.
52 CHAPTER 3. THE ELECTROMAGNETIC TENSOR
Exercise 3.13. Prove Stokes for the square planar loop in Fig 3.1.
A familiar, special case of the formula is a closed loop at fixed time, dxµ =
(0, dxk ), where
I I Z Z
k
Ak dx = A · d` = ∇ × A · dS = B · dS = Φ
Exercise 3.14. Suppose that γ is a curve in Euclidean spaceR and consider the
surface S spanned
R by the curve for t ∈ [0, t0 ]. Show that F dS is the emf
action, i.e. Eemf dt.
a b
V
R
Figure 3.2: Voltmeter measures the electromotive force: Eab = − E · d`, which
is gauge invariant but, in general, path dependent (and non-local). In contrast
with A0 which is gauge dependent, but path independent (and local). It is path
independent provided the loop defined by the path (black) and its variation
(gray) enclose no time-dependent magnetic fields.
3.2. THE ELECTROMAGNETIC POTENTIAL 53
R
The emf E = γ E · d` is manifestly gauge invariant and so, in principle,
measurable. In general, it is non-local4 and path dependent. It is not a prop-
erty of a single event but the pair (a, b) and, in general, depends on the path
connecting them. The integral form of Faraday induction law is
I
Φ̇
E · dx = −
c
where Φ is the magnetic field enclosed by a loop (at a given time slice). If the
variation of the path γ does not enclose any time dependent magnetic fields
Faraday induction law, Eq. 3.12, says the emf is (at least locally) path indepen-
dent. We may then define E = −∇φ. The difference in φ is what a voltmeter
measures.
3.2.5 Generalizations
Manifolds
Since we wrote the equations in a tensorial form, they have a natural translation
to space-time manifolds that are only locally Minkowski.
d + 1 dimensions
It may well be that the apparent dimension of space-time as 3+1 is the dimension
we perceive on macroscopic scales whereas the dimension is different (larger) in
the microscopic scale. It is a nice feature of the formalism that one can formulate
electrodynamics in d+1 space time dimensions. The vector potential Aµ has d+1
components, and the field F is still a second rank antisymmetric tensor. It has
d+1
2 components. d components “electric” and the remaining are “magnetic”.
In d = 3 the number of electric and magnetic component coincide. In lower
dimension there are more electric components andd in higher dimensions more
magnetic.
Exercise 3.16 (1+1 dimension). How many components does the electric field
have in one dimension and how many the magnetic field?. Show that the electric
field is a scalar.
QCD, and other nonabelian gauge theories, are the non-commutative general-
izations of electrodynamics in the sense that the (real, commutative) 4-vector
potential Aµ , is replaced by a Hermitian (matrix valued) 4-vector potential.
4 To measure the emf oen needs to measure E simultaneously along the Euclidean path γ.
One could argue that a simple instrument, like a AVOmeter can’t possibly do that exactly. It
is at best an idealization.
54 CHAPTER 3. THE ELECTROMAGNETIC TENSOR
Fµν η µν = Fµ µ = 0
3.3.2 Dual: F ∗
Duality, denoted by ∗, is an operation whose square in the identity: ∗∗ = 1.
This means that taking a dual involves no loss of information. In n dimensions,
the Levi-Civita tensor allows us to define a duality for anti-symmetric tensors of
rank r and anti-symmetric tensors of rank n − r. Note first that anti-symmetric
tensors of rank r make a linear space whose dimension nr (the number of
independent components). The contraction of the Levi-Civita with an anti-
symmetric
n
n−r
tensor of rank r gives an anti-symmetric tensor of rank n − r. Since
r = r the operation can be used to define a duality.
In 4-dimensions the dual of an anti-symmetric second rank tensor is a second
rank tensor. The dual of F is then defined by
p
∗ µν εµναβ ∗ |g|εµναβ αβ
(F ) = p Fαβ ⇐⇒ (F )µν = F (3.17)
2 |g| 2
In Lorentz-Cartesian coordinates,
p g = η and |g| = 1, of course. To avoid writing
ugly formulas involving |g| is shall sacrifice generality for transparency and
write formulas in Lorentz-Cartesian coordinates.
Exercise 3.17. Show that ∗ is indeed a duality. Namely, show first that
1.
εαβγδ εαβµν = 2(δα µ δβ ν − δα ν δβ µ )
3.4. THE HOMOGENEOUS MAXWELL EQUATIONS 55
Similarly,
For example, the field of a charge moving at uniform velocity has E2 − B2 >
0 and E · B = 0. Similarly, The field of a plane electromagnetic wave has
E2 − B2 = E · B = 0, in any frame.
Exercise 3.18 (Alternate form). Show that 4 equations (3.19) can also be
written as
∂α Fβγ + ∂β Fγα + ∂γ Fαβ = 0
Exercise
p 3.19.Show that if F is an anti-symmetric second rank tensor then
∂µ |g|(F ∗ )µν transforms like a 4-vector under curvilinear coordinate trans-
formations.
In mechanics the price for using the a non-inertial coordinate system, such as
earth, leads to the price of the emergence of new forces: Coriolis and Centrifugal.
You may wonder if there is an analog in electrodynamics. The next exercise
explains why there is none.
Exercise 3.20. Consider the coordinate system attached to the rotating earth
introduced in the previous chapter.
1. Compute det g for the earth rotating coordinate system. (Answer: det g =
−1)
2. What does this imply about the homogeneous Maxwell equations in the
earth frame?
Variational principle
Figure 4.1: The blue curve shows S(x) = x2 . The black curve shows the action
under a change of coordinate x → tan x, a change of scale and shift S →
2S − 1/10. The action changes, but the physical point p where the minimum
occurs is the same. In the example also the coordinates of the points are the
same: x = 0.
Requiring that the action be a Lorentz scalar which reduces to the Newtonian
form for non-relativistic motions fixes the action uniquely.
57
58 CHAPTER 4. VARIATIONAL PRINCIPLE
Figure 4.2: The world line is required to have time-like tangents–otherwise the
action is complex. The blue curve represents the variation of a world line with
fixed end points. The black line maximizes the proper-time τ . The red lines are
light-like. The red path between the end points has zero proper-time.
and
r
mc dxα dxβ mc2
q
L=− −gαβ (x) =− −g00 c2 − 2g0j cv j gjk (x)v j v k
γ dt dt γ
4.1.2 Interaction
The action in Mechanics is constructed from the potentials, not the fields. The
electromagnetic potential is a 4-vector, which we can pair with the path element
dxµ to form a scalar. Electromagnetic interaction is, of course, proportional to
the charge. In c.g.s we also need c to fix the units and the overall sign is fixed
by the choice of the metric η = (−1, 1, 1, 1) so
Z Z Z
e µ e
Sint = Aµ dx = −e φ(x)dt + A(x) · v dt (4.2)
c c
This is precisely the terms one adds to the Kinetic energy in classical mechanics
to describe the interaction with E and B.
A0µ = Aµ + ∂µ Λ (4.3)
This leads to
Z
0 e e
(∂µ Λ) dxµ = Sint −
Sint = Sint − Λ(xf ) − Λ(xi ) (4.4)
c c
This means that although Sint changes under a gauge transformation, its vari-
ation δSint does not (so long as the end points are fixed). This guarantees that
the Euler Lagrange equaltions are gauge invariant.
i.e. it is a function of the position and velocity. The end point events xi
and xf are fixed. The action associates a number for every path x(τ ) =
{x0 (τ ), . . . x3 (τ )}. We shall denote the variation of the path by
δx = { δx0 (τ ), . . . δx3 (τ ) }
| {z }
inf initesimal f unctions
The rule of the game is that the end points are fixed: δx vanishes at the end
points events. (The events which are fixed not the proper-times at the end
points.) The strategy is to use integration by parts to bring δS to the form
x
Z xf
δS = hµ (x, u)δxµ xfi + gµ (x, u, u̇)δxµ dτ
xi
Since δxµ vanish at the end points the first term drops. And since δxµ are
arbitrary,1 the only way for the integral to vanish is if
gµ (x, u, u̇) = 0
These are differential equation that the optimal path must satisfy. They are
known as Euler-Lagrange. Lets consider examples.
4.2.1 Variation of Sp
In Minkowski Cartesian coordinates2 , g = η, the variation in the proper-time
with is given by
By assumption, the variation vanishes at the boundary so the first term drops.
since δx is arbitrary, so long as it is small, if follows that δSp = 0 is equivalent
to conservation of 4-momentum:
ṗµ = mu̇µ = 0
1 Viewed as functions of τ the variations satisfy one constraint: (dδxµ )(dδxµ ) = −(cdτ )2
2 The case of a general metric g(x) is treated in a supplement
4.2. VARIATION OF THE ACTION 61
The basic idea in the calculus of variation is to use integration by parts to get
rid of terms of the form δdx. Hence, rewrite the last term
Combining the two expressions and changing summation indices where needed
we find
Hence,
Z
e e
Aµ δxµ |bdry + Fµν uν dτ δxν
δSint = (4.6)
c c
e
m u̇µ = Fµν uν (4.7)
c
Remark 4.2 (Sign conventions). You can use the non-relativistic limit to fix
and verify the signs convention in 3.9
62 CHAPTER 4. VARIATIONAL PRINCIPLE
xn (t) = A sin t
for arbitrary amplitude A. All solve the equation of motion ẍ = −x and so are
local minimizers. For all the action vanishes:
Z π Z π
S= Ldt = A2 (n2 cos2 nt − sin2 nt)dt = 0,
0 0
At the same time there is no nice minimizer connecting the origin and any other
point at half the period.
Figure 4.3: There are infinitely many paths connecting the origin when the time
difference is half the period. But there is no honest minimizer connecting the
origin to any other point on the red line at half the period. The minimizer ”goes
through infinity”.
Note that that the second segment is a solution of the Euler-Lgrange equation
for the Harmonic oscillator Eq. 4.9. Compute the action of the two segments
for A/x0 1. Show that when A → ∞, the total action S(xA ) ≈ −x0 A and
hence diverges to −∞ if x0 6= 0.
For example, the functions in Fig. 4.1 are convex. It is evident that if a function
is convex its minimum is unique. (It may, however, lie at infinity).
In the case at hand x, the argument of S, is itself a function–a path. Func-
tions naturally form a vector space so the notion of convex function also extends
to this case. γ is a convex function of v. This then implies that the minimizer
for a free relativistic particle is unique.
Exercise 4.5 (Charged particle in a radiation field). Show that the equations
of motion of a charge particle in the radiation field of a circularly polarized
plane wave admit solutions that are circular orbit in the plane orthogonal to the
direction of propagation of the light.
4.3 Supplement
4.3.1 Fermat principle
The mother of variational principles is Fermat principle. It formulates geometric
optics at the the minimizer of the time of propagation between two points. The
ray propagates in a medium with index of refraction n(x). The propagation
time dt is c dt = n(x)|dx|. We can think of n as inducing a metric in Euclidean
space–one that measures the propagation time:
dx · dδx
δ(d`) = = t · dδx, δn = (δx · ∇)n
d`
The corresponding time variation is (the integral of)
δ(c dt) = (δx · ∇)nd` + nt · dδx = δx · ∇n d` − d nt + d nt · δx)
The integral of variation vanishes provided the brackets (and boundary terms)
vanish:
d(nt)
= ∇n
d`
In particular, in a region where the refraction index is a constant, ∇n = 0, the
ray keeps it direction of propagation: t is a constant.
Exercise 4.6. Show that the equation of motion is consistent with the t being
a unite vector.
Θ1 n1
n1 sin θ1 = n2 sin θ2
4.3.2 Rainbow
The simplest features of the rainbow can be understood from Snell’s law.
4.3. SUPPLEMENT 65
Φ Α=2Φ-Θ
Θ
Figure 4.6: The blue line shows a ray undergoing one internal reflection in a
drop of water. The impact angle θ is defined in the figure. The outgoing ray
is focused near the maximum 2φ − θ. This partial focusing is called a caustic.
This gives the direction of the rainbow.
Exercise 4.8. Use Snell law and show that a light ray in air (na = 1) hitting a
water droplet, nw > 1, at lattitude θ is reflected back at angle 2α(θ) = 4φ(θ)−2θ,
see figure. The function φ(θ) is defined by Snell law: nw sin φ = sin θ.
A computation gives
dα cos θ
= −1 + 2 √ 2
dθ n − sin2 θ
The derivative vanishes for
3 cos2 θ = n2 − 1
which gives a real value for θ provided 1 < n < 2. This gives the maximal
value of 2α. Evidently, I(2α) = ∞ there. The divergence implies focusing of
the reflected light. This is called caustic in geometrical optics.
Exercise 4.9. Show that for water (n = 1.33) the caustic occurs for 2α = 42◦ .
This is the main angle of the rainbow, first found by Bacon in 1268. (Different
colors have slightly different angles due to the slight frequency dependence of n).
The 4-velocity is
dxµ
uµ =
dτ
66 CHAPTER 4. VARIATIONAL PRINCIPLE
Figure 4.7: The cyan arrows represent light rays from the sun. The two small
light-blue balls represent two water droplets. The red arrow are the reflected
light rays in the direction of the rainbow caustics. The green eye represents the
observer. Pilots sometimes see rainbow that are circular.
We want to find the path that minimizes the action (equivalently, maximizes
the proper time) Z
S=− dτ
Hence
−2δ(dτ ) = (δgµν ) uν dxµ + 2gµν uν δ(dxµ )
Rewrite the first therm as
(δgµν ) uν dxµ = (∂α gµν ) uν dxµ δxα = (∂µ gαν ) uν dxα δxµ
collecting
− δ(dτ ) = d(gµν δxµ uν ) + 1 ν
2 (∂µ gαν ) u dx
α
− (∂α gµν ) uν dxα − gµν duν δxµ
4.3. SUPPLEMENT 67
The first term is a boundary term so vanishes upon integration. The vanishing
of the variation gives the Euler-Lagrange which is of the form
and Γµαβ may be assumed to be symmetric in α, β without loss. One then finds
Γµαβ = 21 g µν ∂α gνβ + ∂β gνα − ∂ν gαβ (4.11)
Maxwell Equations
We stick with the same definition in Minkowski geometry paying the small price
g → |g|, taking care of the fact that the metric is non-definite. We shall denote
by dΩ the space-time volume element.
Lorentz transformations are isometrics of Minkowski space-time. In Carte-
sian coordinates det η = −1. It follows that the Cartesian expression for the
space-time volume is a Lorentz scalar:
p
dΩ = |η|dx0 dx1 dx2 dx3 = dx 0 1 2 3
| dx {zdx dx} = (cdt) dV (5.1)
M inkowski cartesian
in agreement with the fact that time-dilation and space contraction have com-
pensating factors.
5.1.2 Densities
Under a change of coordinates x0 ↔ x a scalar function ϕ(x) has the transfor-
mation rule
ϕ0 (x0 ) = ϕ(x)
1 This is clearly the right notion of volume in orthogonal coordinates where g is diagonal.
The tensorial properties then guarantee that this expression holds in general.
69
70 CHAPTER 5. MAXWELL EQUATIONS
Figure 5.1: The volume of the parallelepiped spanned by the vectors vj is their
triple product.
It retains the value at the image point. Densities, ρ(x) are different. The
associated scalar is ρ(x)dV . It may count the charge in the volume, or the
number of particles there. The rule is therefore
5.1.3 Distributions
We want to introduce Dirac delta function in Minkowsky space-time. The defin-
ing property is Z
f (0) = dΩf (x)δ (4) (x) (5.2)
• Smooth functions form an algebra: You can add and multiply them in the
obvious way. Distributions are a vector space: You can add them but not
multiply. There is not such thing as a square of a delta function, or a root
of a delta function.
• Distributions are naturally viewed as the dual vector space to the space of
smooth and localized functions. With a pair we associate the number given
by integration.
• A distribution D(x) is the zero distribution if for any smooth and localized
function f (x) Z
f (x)D(x)dx = 0
5.2.1 4-current-density
Consider a point particle whose trajectory is given ξ µ (τ ) as a function of its
proper-time. For the sake of simplicity, we work in Minkowski Cartesian co-
ordinates. We can make a 4-current density using only scalars and 4-vectors,
objects that behave nicely under Lorentz transformations:
Z
j µ (x) = ec dτ δ (4) x − ξ(τ ) ξ˙µ (τ )
(5.5)
where dot is a derivative with respect to the proper time. The scalar factor ec
fixes the dimensions to the dimensions of current density.
To relate this expression to Eq. (5.4) integrate over τ , getting rid of one of
the delta functions. Since ξ is a real orbit, there is a 1-1 correspondence between
coordinate time ξ 0 (τ )/c and the proper time τ . Changing variables from cτ to
ξ0
Z Z
dτ
c dτ δ (4) x − ξ(τ ) ξ˙µ (τ ) = c dξ 0 δ (4) x − ξ ξ˙µ 0
dξ
Z
= dξ 0 δ (3) x − ξ δ(ct − ξ 0 )v µ (ξ 0 )
where t0 is the solution of ct0 = ξ 0 (τ ) and v µ = (c, v). The result is a pleasant
surprise because it coincides with
(cρ, j)
5.2. CHARGE DENSITIES AND CURRENTS 73
ct
derived before we knew anything about Lorentz invariance. There are no rela-
tivistic corrections one needs to make to the classical formulas for charge den-
sities and currents. δ (3) is not a density in Minkowsky space, and v µ is not a
4-vector. However, together they conspire to give the 4-vector (density) j µ .
Evidently,
∇ · v(t)ρ(x − ξ(t) = (v · ∇)ρ x − ξ(t)
and thus
0 = ∂t ρ +∇ · (vρ) = ∂µ j µ , j µ = (c, v)ρ
|{z} |{z}
density current
ct
Figure 5.4: Charge conservation expresses the fact that the orbit is a continuous
curve which does not terminate and moves always into the future. If it enters
a box in space-time it also leaves it. If the orbit enters the box at the bottom
leaves it at the top we say that the charge in the box is conserved. If it leaves
and enters on the sides we say that incoming current balances the outgoing
current.
Once this equation holds for one charge, it holds for any number. When
we consider huge numbers of charges with poor spatial resolution we may then
think of j µ (x) as a smooth function on space time, which satisfies the continuity
equation.
interaction is
Z
e
Sint = Aµ (ξ)uµ (τ ) dτ
c
Z
e
= 2 Aµ (ξ)uµ (τ ) δ (3) (x − ξ(τ )) dV d(cτ )
c
Z
e
Aµ ξ 0 , x v µ (ξ 0 ) δ (3) (x − ξ(t)) dV dξ 0
= 2
c
Z
e
Aµ x v µ (t) δ (3) (x − ξ(t)) dΩ
= 2
c
The middle two terms are the 4-current, hence
Z
1
Sint = 2 Aµ (x)j µ (x) dΩ (5.6)
c
describing both smooth and discrete 4-current distributions.
The first term gives a boundary term and vanishes if Λ → 0 at infinity. Sint is
therefore guaranteed to be gauge invariance provided the current satisfies the
continuity equation holds
1 p
p ∂µ |g|j µ = 0 (5.7)
|g|
Figure 5.5: The action associated a number with a given field configuration and
a box in space-time. We allow variation of Aµ inside the box: The variation
vanishes outside the box and on its boundary. This is the analog of what we do
when we vary the path.
This means that the associated action is a boundary term: Its variation vanishes
identically.
We are left with the first candidate. We need first to decide on the sign so
that the action will have a minimum rather than a maximum. Recall that in
Lagrangian mechanics the kinetic energy comes with a positive sign. Now E is
linear in Ȧ so the E2 must come with a positive coefficients.
Z
1
SF = − F ·F dΩ (5.8)
16πc √ Q µ
|{z}
|g| µ dx
The 16π is the choice4 that gives Maxwell equations in c.g.s units and in par-
ticular leads to the Coulomb potential re .
δ(Fµν F µν ) = 2F µν δ(Fµν )
and
δ(Fµν ) = ∂µ δAν − ∂ν δAµ
By the anti-symmetry of F
To get the field equations we consider variations δA for a given source term j.
This variation gives
Z
1
δSint = 2 δAν j ν dΩ (5.10)
c
1 p 4π
p ∂µ |g|F νµ = jν (5.11)
|g| c
variation.
5.4. VARIATION OF THE FIELD: RULES OF THE GAME 79
in accordance wirh Eq. 5.7. If the source j was not current conserving, Maxwell
equations would not form a consistent set of equation.
−Ej = F0j = F j0 j 0 = cρ
4π 0
∇ · E = 4πρ ⇐⇒ ∂µ F 0µ = j (5.12)
c
The spatial components are:
1 4π j
∂µ F jµ = ∂0 F j0 + ∂k F jk = − ∂t Ej − ∂k (εikj Bi ) = j
c c
Using
∂k (εikj Bi ) = −εjki ∂k Bi = −(∇ × B)j
4π 4π k
− Ė + ∇ × B = J ⇐⇒ ∂µ F kµ = j (5.13)
c c
and dot denotes partial derivative with respect to ct. The vector equations are
written in the form of first order evolution equations that allow to propagate E
and B in time, given their initial values and the source J.
In total, there are 8 Maxwell equations for the 6 unknown fields. This looks
like an over constrained system. It is better to view them as two evolution
vector equation for two vectors and view the scalar equations as a constraint on
the initial data. This constraint is preserved by the evolution provided (ρ, J)
satisfy the continuity equation.
Exercise 5.6. Show that the evolution respects the constraint.
Example 5.7 (Current carrying wire). An electrically neutral, infinitely long,
metallic straight wire (along the z-axis) with circular cross section of radius a
carries a stationary current I. Suppose Ohm’s law in the form J = σE with σ a
constant in the wire. Find the profiles of the electric and magnetic fields inside
and outside the wire. Assume cylindrical symmetry, translational symmetry in
the z direction and stationarity. Analyze the problem in cylindrical coordinates
(ρ, θ, z).
By Gauss and the symmetry Eρ = 0 Since the magnetic field is assumed
stationary, by Faraday and the symmetry Eθ = 0. You might then be tempted
to say that outside the wire, one should have Ez = 0. This, however, leads to a
contradiction: Combinig Gauss and Faraday
0 = ∇ × E ⇒ 0 = ∇ × ∇ × E = −∆E + ∇ ∇ · E ⇒ ∆E = 0
Which says that E is harmonic everywhere. Hence, if it is zero outside the wire,
it is zero everywhere. This, together with Ohm’s law, contradicts the assumption
that the wire carries current.
Let us then retreat to the next line of defense and take E = E0 ẑ with E0 a
constant. This is still harmonic By the integral form of Ampere
(
2I 1 ρ>a
B= θ̂ × ρ 2
cρ a ρ<a
where I is the total current. We have used the fact that inside the wire, the
constancy of E implies the constancy of J.
5.5. DIELECTRIC AND MAGNETIC MEDIA 81
It may be a little shocking at first that a neutral current carrying wire bundles
with it an electric field that does not decay as you get far from the wire. This is
a pathology due to the assumed infinite length of the wire.
ρ = −∇ · P, J = ∇ × M − Ṗ
As these are supposed to represent the local sources in the material, we impose
the boundary condition
(We allow for currents and surface charges.) Define the auxiliary fields D and
H
D = E + 4πP, H = B − 4πM
∇ · D = 0, −Ḋ + ∇ × H = 0
Exercise 5.8. Show that the if you treat Gauss law as a constraint, Ampere’s
evolution law respects the constraint.
6 The choice of normalization is unit dependent. The choice made here is standard in cgs.
82 CHAPTER 5. MAXWELL EQUATIONS
The sources respond to the fields and vice versa. If, in the absence of ex-
ternal driving fields, the body in question is in equilibrium, one would expect a
linear relation between cause and effect. Moreover, if we are only interested in
macroscopic length and memoryless response, the relation should be local:
Dj = εjk Ek , B j = µjk Hk (5.16)
This replace the unknown microscopic sources by material specific functions ε
and µ. For a homogeneous system, these are constant matrices. For isotropic
bodies, they are material specific constant numbers. Eq. 5.16 is known as con-
stitutive relation.
Exercise 5.9. Derive Maxwell’s equation from the variation of the action
Z
1
S= dΩ(Dj Ej − B j Hj )
16πc
subject to the constitutive relations.
Exercise 5.10. Explain why one might expect ε to be a positive matrix (i.e.
real symmetric with positive eigenvalues) and why there is no similar expectation
from µ.
Exercise 5.11. Show that the charge distribution of a homogeneously polarized
sphere of radius a is concentrated on the surface with surface density
P · x̂ δ(|x| − a)
Exercise 5.12. Show that the current distribution of a homogeneously magne-
tized sphere of radius a is concentrated on the surface with surface density
M × x̂ δ(|x| − a)
5.6. NEW PHYSICS 83
The diagonal part is the dissipative conductance and the off-diagonal is the
Hall conductance. von Klitzing found that in certain two dimensional systems,
at sufficiently low temperatures, and with sufficiently strong magnetic field, the
system is characterized by non-dissipative topological phases where
e2
σ = 0, σH ∈ Z
h
Planck constant is an indication that the phenomenon is quantum.
A field theory that encapsulates the Hall effect relies on the Chern Simon
action. In 2+1 dimensions we can construct a the scalar F ∗ · A. The Chern
Simon Lagrangian density:
νσ0 νσ0 e2
LCS = − (F ∗ )α Aα = − 2 εαβγ Fβγ Aα , σ0 =
2c2 4c h
α runs over 0, 1, 2. ν is a dimensionless number h is Planck constant. σ0 is
the quantum unit of conductance with dimensions of velocity (in c.g.s) . So
the prefactor the same dimensions as in the corresponding Maxwell term. This
guarantees that the action has dimensions [Energy × time]. Planck constant
naturally appears since the origin of the term is quantum.
Figure 5.7: The phase diagram for the Integer quantum Hall effect for the Hofs-
tadter model on the triangular lattice at T = 0. The vertical axis is the magnetic
flux through the unit cell. The horizontal axis is the chemical potential. Figure
made by Gal Yehoshua for an undergrad project.
As a model for the quantum Hall effect take the Lagrangian density
1
LCS + Aα j α
c2
The corresponding Euler-Lagrange equations are
νσ0 ∗ α 1
− (F ) + 2 j α = 0 =⇒ νσ0 F ∗ = j
c2 c
Unlike Maxwell’s equations, this is not a differential equation, but an algebraic
relation between the fields and sources. In components
We have reproduced some of the known features of the Hall effect, namely
• Ohms law: The current is proportional to the fields
5.6. NEW PHYSICS 85
• There is no dissipation
eiS/~
Now comes a shady trick. We assume that time is a circle we take for Λ the
transformation corresponding to unit of emf-action: A jump in Λ across the
time cut (also, the quantum unit of flux)
hc
∆Λ =
e
The quantization condition is then reduced to spatial integration
Z Z
νσ0 νσ0
F0∗ dS 0 = B dxdy = 0 M od 1
ce ce
Now, by an argument of Dirac, if space is a close manifold, say a torus then the
total flux is quantized
Z
hc
B dxdy = m , m ∈ Z
e
We finally get
h
νm σ0 = mν = 0 M od 1
e2
which quantizes ν to a rational number.
∇ · E = α(∇φ) · B
Ampere law
∂0 E − ∇ × B = αφ̇B + α∇φ × E
Exercise 5.18. Verify.
When φ is a constant one recovers the sourceless Maxwell equations. In
general, ∂µ φ acts like a source term in Maxwell equations.
∇ · E = αδ (2 (x)n · B
The magnetic field on the surface acts as if there was a charge on the interface.
This is something we have already encountered in the CS theory of the quantum
Hall effect.
8 Since E is even and B odd under time reversal the Lagrangian breaks time-reversal unless
φ is also odd under time reversal. The notion of time reversal in the quantum case is subtle.
5.6. NEW PHYSICS 87
Φ=0
Φ=Π
Figure 5.8: The interface between two insulators that are topologically distinct
gives rise to a singular Axion field.
∂0 E − ∇ × B = αδ (2) (x)n × E
This means that electric field on the surface acts as if there were currents at the
interface. In particular in Axion electro-Magneto-statics
B
E
E, B
Φ=0
Φ=Π
E=0
Figure 5.10: A electric charge, (red dot) is placed near a different topological
insulator with zero fields. On the left, the physical setup. On the right the
image method.
The magnetic provides a source term for the electric field. The source term
is precisely the same as the source term in the corresponding electrostatic image
charge problem provided
gα = 2e
Now, if we add to this field the electric field given by electric monopole of charge
e above the x-y plane we obtain the same electric field configuration as in the
electrostatic image charge problem, everywhere, i.e.
x − dẑ x + dẑ
E=e − θ(z)
|x − dẑ|3 |x + dẑ|3
This describes the electric field everywhere. It remains to see what values B
takes in the lower half-space. Now E · n = 0 on the boundary and so we see
that B is the solution of
∇×B=∇·B=0
everywhere subject to the boundary condition that fixes B on the plane z = 0.
We introduce a scalar potential for B in the lower half pace
B = ∇φ, ∆φ = 0
g d
φ(x, y, z = 0) = p , Bz = ∂z φ = g
x2 + y2 + d2 |x2 + y 2 + d2 |3/2
The problem then reduces to solving Laplace equation with two types of bound-
ary conditions.
E B
Image Image
Figure 5.11: The red curve shows the surface charge density that allows the
field to terminate at the surface. On the right one sees the response in the form
of a magnetic field that seems to have a magnetic monopole at the image point.
There is no real magnetic monopole anywhere, of course.
90 CHAPTER 5. MAXWELL EQUATIONS
-1
-2
-2 -1 0 1 2
The energy, momentum and angular momentum of the electromagnetic field are
identified as the conservation laws associated with symmetries of space-time.
Maxwell stress tensor is derived from variation of action due to variations of
the metric.
1 αµ β
T αβ = F F µ − 41 g αβ F · F (6.2)
4π
In the first row and columns you recognize the energy density and Poynting
vector. This explains the 4π normalization. We shall discuss the other terms
below.
93
94CHAPTER 6. CONSERVATION LAWS AND THE STRESS-ENERGY TENSOR
ct
Figure 6.1: The green rectangle represent the field. The red box is a box in
space-time. The symmetries of Minkowski space allow to shift (and rotate the
box) and fields without affecting the action. For infinitesimal shifts of physical
fields, this triviality statement translates to a conservation law.
is conserved. This identifies the first row (and column) with the energy (mo-
mentum) densities.
(6.3)
It follows that for fields that satisfy the Euler-Lagrane equation, the variation
is a boundary term
Z
4πc (δSF ) = − dΩ ∂µ (F µν δAν ) (6.4)
t t t
x x x
Figure 6.2: The action remains the same when the fields and the integration
box are both shifted in space-time. For a small shift the change in action can be
split into two virtual shifts: A shift of the field with the box held fixed, shown
in the middle figure, and a shift of the box with the field held fixed field, shown
on the right.
Z
= dΩ ∂µ F µν Fαν − 14 gαµ (F · F ) δξ α
Z
= dΩ ∂µ T µ α δξ α
Since this is supposed to hold for any (infinitesimal) box and any shift, we get
the conservation law.
tions for the energy and the momentum so each has the appropriate dimensions. A more
convincing normalization procedure will be given when we discuss Maxwell stress tensor.
6.2. T AND VARIATION OF THE METRIC 99
This can be seen ass follows. Under infinitesimal spatial rotation by δθ (a vector)
the coordinates transform like
x0 → x0 , xj → xj + Rkj xk , Rkj = εj km (δθ)m
We are interested in the variation of δAn (x). The rotation acts both on the
coordinates xm and also on the vector Am . Together, we have
δAn = Rj n Aj + (Rj k xk )(∂j An )
= Rj n Aj + (Rj k xk )(Fjn + ∂n Aj )
= Rj k xk Fjn + Rj n Aj + Rj k ∂n (xk Aj ) − Rj k δnk Aj
= Rj k xk Fjn + Rj k ∂n (xk Aj )
By Eq. (6.6) the conservation law is determined by
F 0n δAn = F 0n Rj k xk Fjn + Rj k ∂n (xk Aj )
= Rj k xk F 0n Fjn + Rj k ∂n (xk F 0n Aj )
| {z }
bdry term
dS
Figure 6.4: The momentum flux is a second rank tensor made from teh two
vectors: Momentum and velocity.
m ∂L
L= gij q̇ i q̇ j , pj = = mgjk q̇ k = mq̇j
2 ∂ q̇ j
6.2. T AND VARIATION OF THE METRIC 101
Hence
p p p
|g|F · F = δ( |g|) F · F + |g|F αβ F γδ δ(gβγ gδα )
δ
p p
= δ( |g|) F · F + 2 |g|F αβ F γδ gδα δgβγ
p p
= δ( |g|) F · F + 2 |g|F αβ F γ α δgβγ
where γµ are its (real) eigenvalues and Pµ are orthogonal projections. By defi-
nition Y X
det g = γµ =⇒ log |g| = log γµ
and so
p X δγµ
δ log det g = 21 δ log(|g|) = 1
2 γµ
We want to express the right hand side in terms of g and its variationδg. To do
that observe that, by the functional calculus of operators,
X Pµ
g −1 =
γµ
102CHAPTER 6. CONSERVATION LAWS AND THE STRESS-ENERGY TENSOR
and so
X Pµ (Pν δγν + γν δPν )
g −1 δg =
γµ
Exercise 6.9 (Projections). Show that if Pµ are orthogonal projections, Pµ Pν =
δµν Pν , follows that
T r(Pµ δPν ) = 0
and so finally p p p p
δ |g| = |g| δ log |g| = 12 |g|g γβ δgβγ (6.11)
Using this we can now manipulate (twice the covariant β components) of the
vector in the brackets
2F νµ ∂ν Fβµ − F µν ∂β Fµν
= −F µν 2∂ν Fβµ + ∂β Fµν
= −F µν ∂ν Fβµ + ∂β Fµν +∂ν Fβµ
| {z }
use hom
µν
= −F − ∂µ Fνβ + ∂ν Fβµ
= −F µν ∂µ Fβν + ∂ν Fβµ
| {z }
µ−ν symmetric
=0
Exercise 6.10 (Plane waves). Show that the stress tensor for plane electromag-
netic waves
Aµ = aµ eik·x , k µ Aµ = 0, kµ k µ = 0
is
4πT µν = a · a k µ k ν
e2
δP = −F × T x̂, F =
4d2
to the system. The charge is at rest, so it is not the charge that is absorbing
the momentum. It mus be the field.
Figure 6.5: You need to apply a (non electromagnetic) force to hold two op-
positely charged particles apart. This force can be computed from the surface
integral of T 11 on the relevant surface. The arrow, the force and the surface are
all in the 1 direction.
We now use Eq.6.13 to see where the momentum of the field sits in T jk .
Consider the space-time box associated with the left half space of time duration
T , i.e.
Ω = {t, x|0 < t < T, x1 < 0}
Now consider what Eq. 6.13 implies for this box. Lets focus on the 1 component:
Z Z Z
1 1
∂α T 1α dΩ = jα F 1α dΩ = j0 F 10 dΩ = −F T
Ω c Ω c Ω
6.3. ENERGY MOMENTUM CONSERVATION 105
The right hand side was easy to evaluate since there is a single, stationary a
point charge in the box which only feels the electric field of the other side5 This
is interpreted the momentum Bob transfers to the box.
Now let us see how the field accommodates the momentum. For this consider
the left hand side of the equation
Z Z
dΩ ∂µ T 1µ = dSµ T 1µ
Ω ∂Ω
So
Z Z T Z
1µ 1 2
dSµ T = dt dx2 dx3 T 11 , T 11 = E2 + E32 − E12
∂Ω 0 8π
For the case at hand, E2 = E3 = 0 and E1 6= 0, but this is not crucial. The
point is that we can now identify T 11 with the momentum that is either leaving
or entering the space-time box through its boundary. If we now focus on the
spatial boundary of the box, namely, the plane, we get the interpretation that
T 11 is the force density on the boundary,
Z
F1 = dx2 dx3 T 11
Force per unit area is the standard notion of stress in elasticity. A similar
interpretation applies to the other components by considering more complicated
boxes.
The parallel and perp components come with opposite signs. The stress can be
positive or negative, but the sign has nothing to do with the signs of E.
Figure 6.6: You need to apply a force to hold two oppositely charged capacitor
plates apart (red arrows). If the arrow is in the z-direction, T zz < 0. You get
this sign if you think of the electric field lines as rubber band: As if the pressure
is negative.
Figure 6.7: The magnetic field lines of a solenoid. The stress in the radial
direction T ρρ > 0 inside the solenoid. This is because Bz 6= 0 while Bρ ≈ 0.
You get the right sign if you replaced the field lines by rubber bands: Stretched
rubber bands along the z-axis, that fan out in the radial direction, will lead
to a positive pressure in the radial direction and negative pressure in the axial
direction.
6.4 Applications
6.4.1 Radiation pressure
The luminosity of the sun L = 3.6 × 1026 [W ], giving a stream of 1045 pho-
tons/sec. The radial component of Maxwell energy momentum tensor at a
distance R from the sun is then
L
T0r̂ = (6.16)
4πR2 c
Consider a macroscopic (black) particle of radius r that perfectly absorbs radi-
ation. The force on the particle at a distance R from the sun is then
r2
Fradiation = L (6.17)
4R2 c
The gravitational force on a particle with density ρ is
4πρr3 M G
Fgravity = (6.18)
3R2
6.4. APPLICATIONS 107
Where M = 2 × 1033 [gram] and Newton constant G = 6.7 × 10−8 [cgs]. The
ratio of the two is then
For water ρ = 1 [gm/cm3 ]. For earth, r = 6 × 108 [cm], the ratio is minuscule:
10−10 . However, for very small grains, of radius less than 6×10−2 [cm] ≈ 600 [µ]
radiation dominates.
Radiation pressure cleans the solar neighborhood from fine dust. This could
be a mechanism of transporting viruses from our solar system to distant parts
of the universe6 .
Exercise 6.13 (Comet tails). Can you figure out the shape of a comet tail?
Suppose the tail is associated with a planet in circular non-relativistic orbit.
Hint: Figure out the tail in the rotation frame.
Figure 6.8: A planet encircling a star and the tail of dust it sprays (tail)
A ρAdM G
Fradiation = L > Fgravity = (6.20)
4πR2 c R2
1
L > ρdM G (6.21)
4πc
Up to factors of order unity we get, the same estimate as above. You need very
thin sails to build solar sails.
6 The assumption that the particle is black is not reasonable when the radiation penetrates
Figure 6.9: Halbach array gives a large magnetic field above the array and small
one below it.
∇ · V = 4πρ, ∇ × V = 4πω
∇·ω =0
Radial vector fields are vorticity free. x is vector field with uniform source,
∇ · x = 3.
The converse is also true: The sources ρ and ω with ∇ · ω = 0 determine the
field V. By linearity, we can decompose the problem in to two problems:
V =E+B
∇ · E = 4πρ, ∇×E=0
and B is sourceless
∇ · B = 0, ∇ × B = 4πω
As we shall see the equations for E and b are solved by the same technique.
109
110 CHAPTER 7. POISSON EQUATION, CLOAKING
Figure 7.1: Left: An irrotational field with a source at the origin. A sourceless
field with vorticity along the z-axis
Integrate this identity on a ball at the origin. The last term (on the right) gives
φ(0). Since G is a radial function, the middle term can be written as
Z Z R Z
d−1 0
2∇φ · ∇GdV = ωd r dr G (r) dS · ∇φ
|x|≤R 0 |x|=r
| {z }
0 by Gauss
(φ is Harmonic, the flux through any closed surface of ∇φ vanishes so the integral
on the right vanishes for any r > 0.)
It remains to integrate the term on the left
Z Z
∆ (φG) dV = dS · ∇ (φ(x)G(r))
|x|≤R |x|=R
Z
= dS · (∇φ)G(r) +φ(x) G0 (r)r̂
|x|=R | {z }
0 by Gauss
Z
= G0 (R) dS · r̂ φ(x)
|x|=R
∆E = (d · ∇)∆φ = 0
7.1. VECTOR FIELDS IN 3D: SOURCE AND VORTICITY 113
3. Using the principle of virtual work show that the torque on the dipole is
T=d×E
4. Suppose that at every point x the dipole is co-oriented with the field
d(x) = ± d Ê(x)
and the dipole is placed initially at a point where E(0) = 0 (no sources
near the origin). Show that one of E± = ±d|E(x)| always has a minimizer
at x = 0.
5. What is the form of the minimizer?
6. Can you relate this to the stability of the Levitron?
Figure 7.2: The vector field of a moving charge with rapidity φ = 1. The field
is manifestly radial but not spherically symmetric.
∇ · A = 0, ∆φ = −4πρ
∆Λ = ∇ · A
∇ · A0 = ∇ · A − ∆Λ = 0
E = −∇φ0 − Ȧ0
Remark 7.11 (Causality). The Coulomb gauge is a-causal: The scalar potential
φ is fixed by the instantaneous charge distribution. You move a charge here and
the potential φ changes immediately everywhere. The fact that the potential
changes faster than light has no use for transferring information because the
fields are still causal.
and θ(x) = 1 for x > 0 and 0 otherwise–the standard step function. Now
consider the limit a → 0 and I → ∞ so that that Ia2 is fixed.
Exercise 7.16 (Delta function). Show that
θ(a2 − x2 − y 2 )
lim δ(z) = δ(x)
a→0 πa2
The a → 0 limit represents a point dipole, characterized by a vector
2 2
π Ia
m= ẑ
c
Ampere equation takes the form
To find B we could plug the source into Biot-Savart. However, this is not much
simpler then retracing the derivation. For A we find
(m × ∇)y δ(y)
Z
A(x) = dy
|x − y|
Z
δ(y) δ(y)
= dy (m × ∇)y −(m × ∇)x
|x − y| |x − y|
| {z }
bdry term
Z
δ(y)
= −(m × ∇)x dy
|x − y|
1
= −(m × ∇)
|x|
m×x
=
|x|3
To compute B we need a version of the vector identity
a × (b × c) = b(a · c) − c(a · b)
∇ × (m × c) = m(∇ · c) − (m · ∇)c
118 CHAPTER 7. POISSON EQUATION, CLOAKING
∇ · (xr−3 ) = 4πδ(x)
Hence x
B(x) = 4πmδ(x) − (m · ∇)
r3
It follows that the magnetic field of a dipole is
−m + 3(m · x̂)x̂
B= + 4πmδ(x)
|x|3
Remark 7.18 (Singularity). The magnetic field has a bad (non-integrable) sin-
gularity at the origin. One way to see this is to consider the total flux through
the origin. Take the plane oriented with m through the dipole. The flux through
such a plane is
=0
z }| {
(m · m)(x · x) − 3((m · x))2
B·m=− + 4πm2 δ(x)
|x|5
(m × x)2
=− + 4πm2 δ(x)
|x|5
The first term has an non-integrable singularity at the origin. At the same time,
we know that the total flux through any surface must be zero
Exercise 7.19 (Vanishing flux). Show that the total flux through any such plane
at distance ε from the origin vanishes.
We look for a vector potential whose flux is the monopole charge em , and
with nice Coulombic field
x
B = em 3
|x|
There is no smooth A that does it. So imagine we allow A with a singularity
along the negative z axis. (But B is still smoth.) Consider the sphere minus
the south pole. Then, by Stokes,
Z Z
4πem = ∇ × A dS = A · d`
`
Aφ 2em
Aφ̂ (θ = π) = √ =
gφφ r sin θ
Exercise 7.20. Show this. (Recall that Aφ eφ = Aφ̂ φ̂, with eφ · eφ = gφφ )
1 − cos θ
A = em φ̂
r sin θ
is nicely behaved along the positive z axis. This singularity is called the Dirac
string. A computation you are asked to do in the next exercise shows that A is
the vector potential of a monopole
x
∇ × A = em
|x|3
Exercise 7.21 (Monopole). Show that the covariant component of the vector
potential is Aφ = g(1 − cos θ). Show that
em
Br = B r = Br̂ =
r2
You may think of the string along the negative z-axis as a flux tube that
brings in the magnetic flux that emanates from the monopole. By the Aharonov-
Bohm effect, such a string is invisible if the flux satisfies Dirac quantization rule.
If n 6= 0 the loops link. The converse is, however, not always true.
7.4. CLOAKING 121
7.4 Cloaking
7.4.1 Maxwell equation in a dielectric medium
In a dielectric medium, the homogeneous equations, Faraday’s law and no-
monoples, are the same as in free space1
∇ · B = 0, Ḃ + ∇ × E = 0, (homogenous)
∇ · D = 0, Ḋ − ∇ × H = 0 (inhomogeneous)
Dot stands for derivative with respect to x0 and D and H are defined through
the constitutive relations2
D = εE, B = µH
where ε and µ are tensors. In a fixed, narrow, band of frequencies, these tensors
can be viewed as functions of the spatial coordinates alone. This will be assumed
from now on. In free (Euclidean) space
ε=µ=g
1 E and B then represent averages over a macroscopically small, but microscopically large,
ball.
2 Hopefully no confusion will arise between ε as dielectric constant and ε as Levi-Civita
tensor.
122 CHAPTER 7. POISSON EQUATION, CLOAKING
1 √ εijk
∇ · E = √ ∂j ( gE j ), (∇ × E)i = √ ∂j Ek
g g
We write the equations in terms of Ej and Hj both with lower indexes. In the
case of vacuum µ = g.
The inhomogeneous equations (without sources) are
√ √
∂j ( gDj ) = ∂j ( g εjm Em ) =0, (Gauss)
| {z }
Em 6=Dm
√ j ijk √
g Ḋ − ε ∂j Hk = g ε Ėm −εijk ∂j Hj =0
jm
(Ampere)
| {z }
Dm 6=Em
This shows that in a dielectric material where the dielectric and permeability
tensors are the same (and time independent) hen we can reinterpret the consti-
tutive relations as metric:
εjm = µkm ⇐⇒ g jm
∂(x0 )i ∂(x0 )j
(g 0 )ij = g αβ (7.24)
∂xα ∂xβ
This translates a dielectric into empty space in curved coordinates. This is the
basis of cloaking.
Exercise 7.22. Consider the scaling (x0 )j = αxj . What is the value of ε and
µ that behaves like vacuum? (Note that time has not been scaled.).
Exercise 7.23. Show that if ε(x) is known, then the field on the surface, En (x)
is a linear functional of φ(x) on the surface.
(The metric g is known.) This is a tensorial equation, it retains its form under
coordinate change. Consider a coordinate transformation that only affects the
interior of the body. This will not affect anything you can do or measure, , but
it will scamble ε and φ inside the body. It follows that you can not determine
ε from (static) boundary data.
Figure 7.6: Straight lines, geodesics in the Euclidean metric, look like deformed
curves when plotted in the coordinates x0 associated with the deformed metric
g 0 . In cloaking you use this fact, and Fermat idea, to force the light avoid the
hiding place.
h(r0 ) is a nice, monotonic, 1-1. This introduces a dielectric medium inside the
unit ball. We want to find the dielectric media.
Using the covariant version of Eq. 7.24:
2 !
h2 (r0 )
0 0 1 2 0 2 0 2
p 0
(g )r,θ,φ (x ) = , h (r ), h (r ) sin θ , g 0 (x ) = 0 0 sin θ
h (r0 )
0 h (r )
sin θ 2
(hh0 (r0 ) , 1, sin−2 θ
h0 (r0 )
viewed as a function of (r0 , θ). Replace the argument r0 → r gives the function
sin θ 2
(hh0 (r) , 1, sin−2 θ
0
h (r)
Plugging in Eq. 7.23 gives for the contravariant components of the (diagonal)
tensors !
2
h(r)h0 (r)
r,θ,φ r,θ,φ 1 1 1
µ =ε = 0 , 2, 2 2
h (r) r r r sin θ
The dielectric tensor looks nicer, and may have clearer physical meaning, in
normalized components
2 !
h(r)h0 (r)
1
µr̂,θ̂φ̂ = εr̂,θ̂φ̂ = 0 , 1, 1
h (r) r
h(r) = r for r > 1 one gets µ = ε = 1 outside, as one must. However, inside the
ball we have a non-trivial dielectric medium. Such a medium is invisible.
Exercise 7.24. Explain why the tensor represents and isotropic medium.
7.4. CLOAKING 125
7.4.6 Cloaking
In cloaking you want more than an invisible ball. You want to use its interior to
hide something. This can be achieved with an h that creates a protected cavity
inside ball of radius 1/2.
1+r
r0 = θ(1 − r) + rθ(r − 1)
2
r'
2.0
1.5
1.0
0.5
Figure 7.7: Coordinate change which creates a hiding cache in a ball of radius
1/2. The origin The surface of the sphere r0 = 12 is mapped to the origin r = 0.
The interior of the sphere is mapped into a different world r < 0. The real world
with a hole r0 > 12 is mapped into a fictitious world without a hole which looks
empty.
r = h(r0 ) = 2r0 − 1
Plugging in the equation for the dielectric functions we get in the coordinates
of physical space (with the standard spherical coordinates)
2 !
1 2(2r − 1)
µr̂,θ̂,φ̂ = εr̂,θ̂,φ̂ = , 1, 1 , 1/2 < r < 1
2 r
Electromagnetic waves
Maxwell equations in vacuum are reduced to the wave equation for the poten-
tials. We discuss the notions of polarization, Green’s function, retarded and
advanced solutions.
∂ µ Aµ = 0
This is known as the Lorenz gauge condition (see section 9.1.2). Here we derived
it as a special case of the Coulomb gauge in the absence of charges. In fact, one
can always impose the Lorentz gauge condition (even when there are charges),
but the proof that one can do that shall only be given later. It is manifestly
gauge invariant.
In the Lorentz gauge the potentials satisfy the wave equation
1
Aµ = 0, = ∂ µ ∂µ = − ∂tt + ∆
c2
Remark 8.1 (Lorentz invariance). Since the Dalambertian, and the Lorenz
gauge conditions are manifestly a Lorentz invariants, Lorentz transformations
of electromagnetic waves are electromagnetic waves.
127
128 CHAPTER 8. ELECTROMAGNETIC WAVES
Ė + ∇ × B = 0, Ḃ − ∇ × E = 0
Ë + ∇ × (∇ × E) = 0, B̈ + ∇ × (∇ × B) = 0
∇ · E = 0, ∇·B=0 (8.2)
Combining we get that the electric and magnetic fields satisfy the wave equation
Ë − ∆E = 0, B̈ − ∆B = 0 (8.3)
Waves that satisfy the wave equation, Eq. (8.3), and the divergence-less con-
straint, Eq. (8.2) are called transverse waves.
aµ → aµ + λkµ (8.5)
k · k = 0, k·a=0
This says that k is a light-like vector and fixes the dispersion relation
ω = ±c|k|
8.2. PLANE WAVES 129
In the Coulomb gauge, the amplitudes are orthogonal (in Euclidean space) to
the direction of propagation.
Figure 8.1: The triad of E, B, k for a plane wave. The wave propagates in the
k̂ = Ê × B̂ direction
130 CHAPTER 8. ELECTROMAGNETIC WAVES
8.2.2 Doppler
Since k · x is a Lorentz scalar
k · x = k 0 · x0
Lorentz transformation of a plane wave is a still a plane wave. The wave has,
in general, different wave vectors and amplitudes in different frames:
kµ0 = Λµ ν kν , a0µ = Λµ ν aν
Longitudinal Doppler
Consider a plane wave propagating in the z-direction, Eq. (8.7). Boosting the
wave with rapidity φ in the same direction is the same as viewing the wave from
an inertial frame boosted in the opposite direction. The associated Lorentz
transformation is
This is linear in the velocities for small speeds. a1,2 are not affected by the
boost.
Transverse Doppler
Consider, as before, a wave propagating in the z-direction, but a boost in the
x-direction so that
ω 0 = ω cosh φ = ω γ
Let us compute the threshold for particle production. The total energy-
momentum of a proton with rapidity φ and counter-propagating photon in the
plane is
pµ = mP (cosh φ, sinh φ) + ~ω(1, −1) = (p0 , p1 )
The energy in the center of mass frame, Ecm , is the scalar
q q
Ecm = p20 − p11 = m2P + 2mP ~ωeφ = mP + mπ
and the equality on the right expresses The threshold for pion production.
Exercise 8.3. Can you figure out why the estimate is too big?
where M is the mass of the atom. This leads to low temperatures whenerevr
the energy level has long life time so Γ is small. Indeed,
2
Γ
= O(α6 ) = O(10−12 )
E
132 CHAPTER 8. ELECTROMAGNETIC WAVES
8.3 Polarization
8.3.1 Amplitude and phase
Scalar plane waves are simply characterized by their frequency ω, wave vector
k, amplitude and phase. Electromagnetic waves, being vector valued, are more
complicated. In addition to the amplitude and phase they are also characterized
by their polarization.
The electric field of an electromagnetic wave propagating is the real part of
E0 eiφ , φ = k · x − ωt (8.12)
Figure 8.2: Four Stokes parameters describe elliptically polarized light. Three
numbers identify the size of the ellipse, its tilt to the axes, its eccentricity. A
fourth number gives the purity (the coherence) of the light. The plane of the
ellipse is perpendicular to the direction of propagation k.
8.3.2 Polarization
Let x̂ and ŷ denote orthogonal unit vectors in the plane perpendicular to k.
Write √
E0 = E+ z+ + E− z− , 2z± = x̂ ± iŷ
We may formally identify E± with the components of a an (un-normalized)
spin 1/2 namely, |ψi ⇐⇒ (E+ , E− )t . Quantum mechanics provides us with a
canonical procedure for factoring out the normalization and the overall phase
in of a quantum state: The density matrix:
∗
|E+ |2 E+ E−
1
ρ = |ψi hψ| = ∗ (8.13)
|E+ |2 + |E− |2 E+ E− |E− |2
8.3. POLARIZATION 133
p
ρ does not care about the overall amplitude, |E+ |2 + |E− |2 , and overall phase
of the wave. Since T rρ = 1 while det ρ = 0 the two eigenvalues of ρ are 1 and
0: ρ is a projection ρ2 = ρ
Figure 8.3: The Poincare sphere associates with every point on the sphere a
polarization. The north and south poles represent right and left circularly po-
larized light and the equator with linearly polarized light.
Exercise 8.4. Show that antipodal points on the Poincare sphere represent
orthogonal polarization in the sense that ρŝ · ρ−ŝ = 0.
are
s3 = cos2 χ − sin2 χ = cos 2χ,
s1 + is2 = 2e−iψ cos χ sin χ = e−iψ sin 2χ
This makes 2χ and ψ the standard spherical coordinates.
Circular polarization
The north poles correspond to cos χ = 1 so that E0 =⇒ z+ . The electric field
is–up to an overall amplitude and phase–
√
2E = x̂ cos φ − ŷ sin φ, φ = k · z − ωt (8.17)
As φ increases from 0 to 2π the vector E describes a circle in the x-y plane which
is turning clockwise. At a fixed z and as a function of time the field rotates
counter clockwise while propagating: It behaves like a left handed screw and so
is called left circularly polarized.
Exercise 8.5 (South pole). Show that the south pole represent right circular
polarization.
Linear polarization
The equator is s3 = 0 =⇒ cos2 χ = sin2 χ. As the overall
√ phase does not play a
role we may take for the amplitude (e−iψ/2 , eiψ/2 )/ 2. The corresponding E0
is:
E0 = 12 (x̂ + iŷ)e−iψ/2 + 12 (x̂ − iŷ)eiψ/2
= x̂ cos(ψ/2) + ŷ sin(ψ/2)
As φ increases from 0 to 2π this describes a line element in the x-y plane at
angle ψ/2 to the x axis. ψ = 0 corresponds to x̂ polarized wave and ψ = π to
ŷ polarized wave.
(Schwartz inequality was used here). By 8.16 this implies that the vector |s| ≤ 1;
the vector lies in the unit ball. The light we get from the sun is completely
unpolarized. It is associated with s = 0, the center of the Poincare ball.
8.3.6 3D glasses
When you view a 3D movie, the 3D glasses transmit a picture with right circular
polarization to, say, the right eye and left circular polarization to the left eye.
Exercise 8.8. Can you give an (ergonomic) argument why spectators would
prefer circular to linear polarization?
Exercise 8.9. Define quarter wave plate as the rotation of the Poincare sphere
that turns circular polarization to linear. Show that it is represented by Hadamard
gate H √
2H = σ3 + σ1
136 CHAPTER 8. ELECTROMAGNETIC WAVES
Exercise 8.10. Explain why the filtering associated by linear polarizers can be
described by the projections
1 ± σ1
PH,V =
2
It follows from the two exercises above that the right and left glasses can be
represented by 2 × 2 matrices
g1 = PH H, g 2 = PV H
PV σ 3 = σ 3 PH
φ = 4∂uv φ = 0
Figure 8.5: If the initial data, φ0 and φ̇0 are localized in the red interval, the
solution at later times lives in the cone. This is called the domain of influence
of the initial (red) data.
This allows for reconstructing f and g from the initial data by integration. As
we shall discuss in more detail below, the localization of φ0 and φ̇0 does not
imply that f and g are localized, see Fig. 8.6.
Any solution can be thought of a a linear combination of the solution with
vanishing initial data for φ̇ = 0 and the complementary case, where the initial
data for φ vanish. We are, of course interested in the case where the initial data
are localized bump functions.
Figure 8.6: The initial data is φ0 = 0 and bump function φ̇0 = 2 cosh−2 x shown
on the left. Also shown are f (x) = tanh(x) and g(x) = tanh(−x). On the
right you see the initial data φ̇(0) again, and the wave φ(x) at time 2. This is
supposed to illustrate that wave lingers near the origin forever and the failure
of Huygens principle in one dimension.
φ0 = 0 =⇒ f = −g =⇒ 2g 0 = φ̇0
Now, the localized initial data φ̇0 do not guarantee that f and g are localized
(only that f 0 and g 0 are bump function). See fig. 8.6. As a consequence Huygens
principle fails in one dimension.
138 CHAPTER 8. ELECTROMAGNETIC WAVES
φ̇2 + (∇φ)2
2. Explain why the failure of Huygens principle in one dimension does not
lead to conflict with energy conservation.
Figure 8.7: The domain of dependence of the initial red data. When Huygens
principle holds, the value of the wave at the space time point is determined by
the intersection of the backward light-cone with the initial data. In general, it is
determined by the intersection of the interior of the cone with the initial data.
Explicit integral representation for the initial value problem of the wave
equation in three dimensions, is given e.g. Wikipedia see also Pinchover and
Rubinstein.
The forward light cone is associated with out going (retarded) waves. Similarly
Z
1 dk
φ< (t, x) = φ̃(−|k|, k) e−i(|k|t+k·x) (8.24)
(2π)2 2|k|
φ = c(u)e−c(u)zz̄ eiv/λ
whose solution is
1
c(u) =
`2 + iλu
k · k = k02 (8.26)
The smallest wave length that such a wave can accommodate is 2π/k0 : The
frequency limis the spatial resolution.
140 CHAPTER 8. ELECTROMAGNETIC WAVES
The noteworthy fact about this waves is that the frequency k0 , does not limit
anymore the spatial resolution 1/k. Near x = 0 one can find waves with k k0 .
E = E0 e−κx eikz
κ(E0 )1 + ik(E0 )3 = 0
Ė + ∇ × B = 0, Ḣ − ∇ × D = 0
∇ · D = 0, ∇·B=0 (8.27)
ωε−1 D + k × µH = 0, ωH − k × D = 0, k · D = 0, k · µH = 0
where µ and ε are the constitutive relations. We assume that µ, ε are positive
matrices. (Possibly functions of ω.) Substitution gives for D
ω 2 = εj µ k 2
√
The wave propagates at different speeds ej µ along the principal directions of
ε. This is birefringence.
This equation merits some discussion regarding the existence and the uniqueness
of the solution. The issue of existence has to do with the fact that the source
terms is singular: A delta function (a distribution). What regularity properties
shall we require of the solutions? When we studied Poisson’s equation–an elliptic
equation–a singular source had a Greens function that was a pretty regular
function (the Coulomb singularity is integrable). This is a feature of elliptic
equations. We will not have this luxury in the case of the wave equation which
is hyperbolic: We shall have to allow for solutions G that are distributions if we
insist on causality.
The second issue is uniqueness. As usual, we have the freedom of adding
solutions of the free wave equation. We fix the solution by imposing causality:
We shall denote by G> the solutions which is created by the source. This is
also called a retarded solution. Formally, the wave equation with a source also
admits advanced solutions where the wave is fully absorbed by the source (and
solutions of mixed type). We select the retarded solutions because it is causal.
As we shall show below, the retarded Green function in 3+1 dimensions is
δ(|x| − ct)
G3+1
> (x) = 2θ(t)δ(x · x) = , x = (ct, x), (8.30)
|x|
142 CHAPTER 8. ELECTROMAGNETIC WAVES
G> lives on the light-cone–rather than in its interior. This is the Huygens
principle.
x0
@s = 0 6 s=0
@ s<0
@
@
@
@ s>0
s>0 @
@ - x1
@
@
@
@
s<0 @
@
@
Gd+1
> (x) = θ(t)f (s), s = xµ xµ (8.31)
We need to ascertain that G> solves the homogeneous wave equation for t > 0.
To do so, let us compute the two derivatives in the Dalambertian one at a time:
It follows
1 µ
2 ∂ ∂µ f (s) = ∂ µ (xµ f 0 (s))
= (∂ µ xµ )f 0 (s) + 2(xµ xµ ) f 00 (s)
= (d + 1)f 0 (s) + 2s f 00 (s)
= (d − 1)f 0 + 2(s f 0 )0
Hence
This is a simple, second order ordinary differential equation which can be inte-
grated directly to give
(d − 1)f + 2(s f 0 ) = c1
To compute the constant consider the left hand side outside of the forward light
cone. Since we insist on causal solutions both f and f 0 must vanish outside the
light cone.
Exercise 8.20. Why?
This says that c1 = 0. We are left with a first order equation
(d − 1)f + 2s f 0 = (d − 3)f + 2(sf )0 = 0 (8.33)
This simple differential equation can be integrated in any dimension d.
Exercise 8.21 (Regular solutions). Show that the regular solutions of the dif-
ferential equation 8.33 are (the real and imaginary parts of )
1−d
f (s) = c2 (−s) 2 d 6= 1,
The solution found in the exercise distinguished d even and d odd. For even
d it looks ok for the (real part) lives in the forward light cone. However, for
odd d the solution is not causal: Causality forces c2 = 0 and we are left empty
handed. What have we missed? We missed solutions that are distributions.
Example 8.22. In 2+1 dimensions
1
G2+1
> = λ2 θ(t) Re √
−s
is causal but does not satisfy the Huygens principle.
Back to 3 + 1
Let us now focus on the case d = 3. The differential equation 8.33 reduces to
(sf )0 = 0
which is easily integrated to
sf (s) = c2
with c2 the integration constant above. We use causality to conclude that
c1 2 = 0 and we are left with
sf (s) = 0
If you treat this as an equation for a function f then f = 0. Have we done
something wrong?
The subtle point here is that f need not be a function. After all the source
was a distribution so we should allow also distributional solutions. Now we can
find a non-trivial solution, namely
sδ(s) = 0
144 CHAPTER 8. ELECTROMAGNETIC WAVES
We conclude that
G3+1
> (x) = λθ(t)δ(s)
and Huygens principle holds.
Exercise 8.23 (Solution in 1 + 1). Using the similar arguments show that
G1+1
> (x) = λ1 θ(t)θ(−s)
The solution lives in the forward light-cone rather than its boundary. It does not
satisfy Huygens principle.
It remains to determine λ3 in 3+1 dimensions. We do that by Gauss law.
We integrate the wave equation with a source at the origin of a space-time box
Ω 2 as in fig. 8.6
Z Z
G> dΩ = 4π δ (4) (x)dΩ = 4π (8.34)
Ω Ω
The left-hand side can be converted to an integral on the boundary of the box
Z Z
G> dΩ = (∂µ G> ) dS µ
Ω ∂Ω
Exercise 8.24 (Signs-Sigh). On the early and late face of the box, determine
the sign in dS 0 = ±dV .
t
@ 6
@
@
T
' @ $
@
@
@
x
@ -
@
& %
@
@
@
@
@
@
Inspecting the figure one realizes that only the intersection of the top face of
the box with the forward light cone contributes. Everywhere else the retarded
solution vanishes. Using dS 0 = dV and Eq. 8.32
Z Z Z
µ
(∂µ G> ) dS = dV ∂0 G> = λ3 dV (−2cT ) δ 0 (s)
∂Ω t=T t=T
2 Minkowski cartesian coordinates with |g| = 1 are assumed.
8.6. GREEN’S FUNCTION FOR THE WAVE EQUATION 145
With s = r2 − c2 T 2
Z Z ∞
dδ dr
dV δ 0 (s) = 4π dr r2
t=T dr ds
Z0 ∞
dδ
= 2π dr r
0 dr
Z ∞
d(rδ)
= 2π dr −δ
0 dr
Z
= −2π drδ(r2 − c2 T 2 )
π
=−
cT
This and Eq.8.34 fixes λ3
π
λ3 (2cT ) = 2λ3 π = 4π =⇒ λ3 = 2
cT
Exercise 8.25. Compute the normalization constants λ1+1 and λ2+1 .
Bibliography
F. John, Partial differential equations,
Y. Pinchover and Y. Rubinstein, —Introduction to Partial Differential Equa-
tions, Cambridge
146 CHAPTER 8. ELECTROMAGNETIC WAVES
Chapter 9
Radiation
Retarded solution of the wave equation. Retarded potentials, retarded electro-
magnetic fields. Lienard Wiechert potentials and fields; Dipole radiation.
147
148 CHAPTER 9. RADIATION
t
zHtL
x0
y0
Figure 9.1: The world like z(t) of the point charge is the blue line. The wave at
the point x is determined by the intersection of the backward light cone with
the orbit at time y 0 . There is one such point since the velocity is time like.
The θ function guarantees causality: The past influences the present. The delta
function says that signals propagate with the speed of light–Huygens principle
holds. If the orbit z(z 0 ) is that of a physical particle, a single point contributes:
a single particle has a single image (in the absence of mirrors) .
To compute the remaining integral use
Z
1
δ f (y) dy = 0 , f (y0 ) = 0. (9.9)
|f (y0 )|
Here f = R · R whose derivative is related to the the 4-velocity of the source
u = −Ṙ:
df df dτ R · Ṙ R·u
= =2 = −2
dy 0 dτ dy 0 γ γ
We obtain a simple looking formula for the wave φ(x) at the observing point x:
γ(y 0 )
φ(x) = , R = x − y, R·R=0 (9.10)
|R · u(y 0 )|
which is manifestly causal and satisfies Huygens principle.
The formula looks simple but at the price of being implicit. The right hand
side is not an explicit function of the argument x. To compute γ(y 0 ), u(y 0 ) and
R = x − y on the right hand side you need first to evaluated at the earlier time
y 0 (see Fig. 9.1.3), which is not explicitly given. This time is determined as the
solution of the equation
2
(x0 − y 0 )2 = x − z(y 0 )
which may be arbitrarily complicated if the orbit z(z 0 ) is complicated.
As expected, the amplitude of scalar waves decays like 1/R. However, you
may argue if you want to call the general solution we have found a wave in all
cases. For example you would probably not call the solution for a source at rest
a wave.
The equations are coupled through the Lorenz gauge condition. If the current
j µ is not conserved then the derivation is inconsistent with the Lorentz gauge.
Conversely, if current is conserved, then Lorentz gauge condition follows for all
times provided the initial data for Aµ and Ȧµ satisfy it.
Exercise 9.1. Show that if one imposes the Lorentz gague condition as initial
data for the wave equation, then the Lorenz gauge condition holds for all times
provided current is conserved.
Comparing with Eq. (10.6) for scalar waves we see that the retarded potentials
are:
uν uν (y 0 )
Aν (x) = e = −e (9.15)
|R · u| R · u(y 0 )
where removed the absolute value by taking into account that R is forward
light-like and u forward time-like so R · u < 0.
The result admits the following interpretation: The vector potential, being
a 4-vector, must be of the form
(scalar)(vector)µ
We have (at least) two 4-vectors at our disposal: R and u. Between these
we can form 3 scalars: Two uninteresting u · u = −c2 and R · R = 0and one
interesting R · u. This, plus dimension analysis and the limit case of a charge
at rest determines Eq. (9.1).
The result can also be viewed as the covariant form of Coulomb law:
e e uν
Aν (x) = (1, 0, 0, 0) = − (−c, 0, 0, 0) =⇒ −e
|x| c|x| | {z } R·u
uν
with
e
A0 = −
|x|
when you change t of the event x the distance x does not change because the
particle is at rest. Hence
∂ 0 A0 = 0
A sneaky argument: Suppose in the distant past the source was a rest. The
coulomb solution then satisfies the Lorenz gauge in the past. By conservation
of charge, the Lorenz gauge is preserved by the evolution.
An honest Computation: The reason for doing also an honest computation
is that this will force us to derive the identity that describes how the retarded
time depends upon variation of the observation event x:
Rµ
∂µ τ = (9.16)
R·u
This follows by differentiating R · R = 0
0 = 21 ∂µ (R · R) = Rα ∂µ (xα − z α ) = Rµ − R · u (∂µ τ )
uµ
µ
0 = ∂µ A = −e∂µ
R·u
?
uµ
z}|{
2 µ µ
(R · u) ∂µ = (R · u) ∂µ u − u ∂µ (R · u) = 0 (9.17)
R·u
We have verified that the solution indeed satisfies the Lorenz condition.
152 CHAPTER 9. RADIATION
Figure 9.2: The self time τ parametrizes the blue orbit. It can be extended to
a function on space time by pushing the value of τ to the forward light-cone.
The figure illustrates how τ changes when the point of observation x changes.
The red lines are light-like.
and the right hand side is a convenient notation. The word formal above refers
to the fact that in taking the partial derivatives we need to remember that τ ,
the retarded time, is a function of the point of observation x. If we want to
treat x and τ as independent variables, ∂µ needs to be interpreted as
∂ ∂τ ∂ ∂ Rµ ∂
∂µ =⇒ + = + (9.19)
∂xµ ∂xµ ∂τ ∂xµ R·u ∂τ
9.4. RETARDED FIELDS 153
The xµ differentiation sees only the first term in R, the location of the observer,
while the τ differentiation only sees the second term, the location of the charge.
Using the explicit form of the potential:
u
ν uµ uν Rµ u
ν
− ∂µ Aν = e ∂µ = −e 2
+ e ∂τ (9.20)
R·u (R · u) R·u R·u
Because F is anti-symmetric, the first term in Eq. (9.20) drops upon anti-
symmetrization and only the second term contributes
The field F depend on the location, velocity u and the acceleration u̇ of the
charge at the early time. It does not depend on any higher derivatives, e.g. the
jerk ü. Now compute:
u u̇ν uν
ν
∂τ = − ∂τ (R · u) (9.21)
R·u R · u (R · u)2
u̇ν uν uν
= + 2
u·u− R · u̇
R·u (R · u) (R · u)2
u̇ν uν uν
= − c2 − R · u̇
R·u (R · u)2 (R · u)2
Consequently
Rµ u
ν Rµ u̇ν Rµ uν Rµ uν
∂τ = − c2 − R · u̇ (9.22)
R·u R·u (R · u)2 (R · u)3 (R · u)3
We get F by anti-symmetrizing:
R[µ u̇ν] R[µ uν] R[µ uν]
Fµν = −e 2
− 3
R · u̇ +e c2 (9.23)
(R · u) (R · u) (R · u)3
| {z } | {z }
radiation ”Coulomb”
Since u is normalized to c the last term, is order O(c0 ). It decays with distance
like R−2 . This is, essentially, the Coulomb term. The first two terms are
proportional to the acceleration and so formally of order O(c−2 ). They decay
more slowly at large distance, like R−1 . These are the radiating terms.
9.4.1 Interpretation
The formula is complicated and at first also opaque. It may be useful to view
it from general principles.
We have three vectors in the problem: R, the light-like vector connecting
the point of observation and the source, u the particle 4-velocity and u̇ its
4-acceleration. From these we can make three interesting scalars
R · u, R · u̇, u̇ · u̇
R · R = 0, u · u = −c2 , u · u̇ = 0 (9.24)
154 CHAPTER 9. RADIATION
F must be linear in the acceleration. This is because the potential did not
depend on the acceleration at all. As a consequence, that the scalar u̇ · u̇ should
not appear and the scalar R · u̇ can only appear in the numerator. Given this
we can reconstruct all the three terms in F , up to sign, just by the fact that F
is a tensor and dimension analysis. From the tensorial properties of F it must
be of the form
(tensor)µν = (scalar) (vector)µ (vector)ν
Since F has dimension of [charge][length−2 ] and u has the dimension of c,
one possible term is
R[µ uν]
ec2
(R · u)3
which gives the last term in Eq. (10.5). You can even get the numerical factor
(and the sign) by looking at the limiting case of the Coulomb field of a particle
at rest where uµ = (−c, 0, 0, 0).
Looking at the terms proportional to the acceleration u̇ one possibility is
R[µ u̇ν]
e
(R · u)2
which gives the first term up to sign and numerical factors. The middle term is
obtained similarly.
e R[i u̇j]
= −
c2 |x|2
e (a(0) × x)k
= (9.25)
c2 |x|2
where a, the 3-vector of acceleration u̇µ = (0, a), is orthogonal to u = (c, 0).
9.5. PARTICLE INSTANTANEOUSLY AT REST 155
e2
E · B = 0, E2 − B2 = −
|x|4
Exercise 9.2. Show that the Poynting vector is
c e2
E×B= x̂
4π 4πc3 |x|2
156 CHAPTER 9. RADIATION
r/c
t-r/c
t-r/c+r.z/c
Figure 9.3: A space time diagram showing the geometric meaning of the succes-
sive approximations leading to Eq. (9.36). The world-line of the moving charge
is the blue path. It is confined to a narrow corridor between the two red lines
a distance 2` apart. The retardation time is the intersection of the back light
cone (green line) with the blue orbit. The first approximation t − r/c is the
intersection of the back light cone with the origin (gray vertical line). The next
correction in Eq. (9.36) is the corner of the small green square.
(We have use the homogeneity of Minkowski space time to shift y 0 = 0 to any
other time.) and similarly for E.
x0 − y 0 = |x − z(y 0 )|
9.6. RETARDATION FROM A DISTANT SOURCE 157
where z(y 0 ) is the obit of the charge. What we seek is an explicit, approximate,
formula for y 0 (x).
In general, there is little one can say for arbitrary xµ and general orbits
0
z(y ). In practice, one often is interested in the field far from the source: The
light from a star; from radiating atom. In these cases the observer is far from
the source:|x| |z|. Assume the orbit z is confined to a ball of radius ` |x|.
Under this condition we can write
2 !
2 2 2 2 x·z `
|x − z| = |x| + |z| − 2x · z = |x| 1 − 2 2 + O
|x| |x|
Hence the implicit equation for the retardation for a distant source reduces to:
2
0 0 0 `
x − y = |x| − x̂ · z(y ) + O (9.29)
| {z } |{z} | {z } |x|
large O(x) O(x) O(`) | {z }
negligible
z(y 0 ) = ` n̂ cos(ky 0 )
So, if
ky 0 = |123456789
{z } .987654321 × (2π)
irrelevant
and we want an accuracy of 1%, then the all we care are the three digits in blue,
just after the decimal point. In other words, cos is a periodic function, we need
only the fractional part of ky 0 . In particular we need to compute the fractional
part of the retardation. We can now write Eq. (9.29) in dimensionless form
2
0 0 0 k`
k(x − z ) = k|x| − kx̂ · z(z ) + O (9.30)
| {z } |{z} | {z } |x|
1 1 O(k`) | {z }
negligible
Whether the term O(k`) can be neglected or not does not depend anymore on
how far the source is but rather on how big k` is.
example, the case for the light emitted by a single atom where the characteristic
wave length is thousands of times larger than the atom. In this case the adequate
approximation it to keep only the big terms in Eq. (9.30) to get the an explicit,
approximate, expression for the early-time y 0 :
Harmonic motion
As an application consider a charge undergoing non-relativistic harmonic motion
with acceleration
0
a(y 0 ) = a0 eiky (9.32)
√
a is a vector with complex amplitudes. Thus, for example ao = `(1, i, 0)/ 2
describes circular motion with radius ` in the plane. Non-relativistic means
ω` c ⇔ k` 1 which is the condition for the dipole approximation to apply.
This sets
ky 0 = k(x0 − r), r = |x|
Using Eq. (9.28) for the magnetic field of a charge moving non-relativistically
we get
0
eik(x −r)
e a0 × x̂
B(x) = (9.33)
c2 | {z r }
spehrical wave
Many particles
The radiation fields from many particles with prescribed orbits is, by linear-
ity of the Maxwell equation, the sum of the radiation of the individuals ones.
In general, each particle will have its own retarded time, and the formulas re-
main implicit. A simplification occurs in the dipole approximation for “ small
antenna” where all the charges share the same retardation.
The dipole moment of a large collection of charges is:
X
d(t) = ej zj (t) (9.34)
and we assume that all the orbits zj are such that the dipole approximation
applies. In this case all the charges have the same retardation and the magnetic
field is simple
d̈(t − r/c) × x̂
B(r, t) = (9.35)
c2 |x|
9.7. POWER 159
9.7 Power
B
E
a
Figure 9.4: Dipole oriented along the z-axis and the associated fields. φ is the
angle between the z-axis and the blue arrrow.
Both E and B lie in the plane perpendicular to the line of sight so P is parallel
to r̂. Suppose that a = aẑ. The magnitude of P in the direction φ relative to
the z-axis is
E2 e2 a2 sin2 φ
P (φ) = c = (9.38)
4π 4πc3 r2
The power through a spherical shell of radius r is then
Z
2
PT = 2πr dφ sin φ P (φ) (9.39)
Evidently
Z Z
4
dφ sin φ sin2 φ = − d(cos φ) (1 − cos2 φ) = 2 1 − 13 = (9.40)
3
Figure 9.5: Polar plot of the power radiated by a dipole antenna as function
of the angle. The maximal power is radiated in the plane perpendicular to the
dipole.
A dipole antenna is not isotropic: It does not radiate at all in the directions
of the dipole.
Remark 9.3. You can not make an isotropic antenna. No matter how com-
plicated an antenna you make the Poynting vector must vanish it at east two
directions. This is a consequence of topology: The vector B is tangent to the
sphere. It is a fact that every vector field tangent to the sphere must vanish at
two points, at least. This is sometimes expressed as you can not comb a tennis
ball. Hence P must vanish at two points at least.
1 e2
E=−
2 |x|
2 e2 2
−Ė = a
3 c3
We can now eliminate a and obtain a differential equation for the energy
25
Ė = −kE 4 , k=
3(me)2 c3
In other words, the charge would collapse on the nucleus in finite time 1/γ.
The ground state energy of hydrogen-like atom is,
2
e2
2
2E0 = −mc = −mc2 α2
~c
and its period is 2π~/E0 . Therefore, the decay time, counted in periods, is
E0 1
= × α−3 = 6400
2π~γ 128π
This means that the electron in hydrogen would fall on the on the proton in
2 × 10−12 seconds. The classical world is unprotected against collapse to the
nucleus.
The apparent instability of the atoms in classical physics is one of the reasons
for quantum mechanics.
E
t
Figure 9.6: Blowup at finite time: The energy of a charged particle encircling
the nucleus goes to −∞ in finite time.
Chapter 10
Radiation reaction
Electrodynamics is a practical theory that proved itself in huge number of appli-
cation. It is a consistent theory when the sources are continuous distributions.
However, it is not a self-consistent fundamental theory of point particles.
163
164 CHAPTER 10. RADIATION REACTION
which is small when the particles move slowly. The problem of interacting
charges is then encapsulated by the Hamiltonian
X 1 1 X ej ek
H= vj2 +
2mj 2 |xj − xk |
The electromagnetic field has been integrated out of the problem. This is the
starting point of musch of atomic physics.
10.1.3 Infinities
Electrodynamics of point-like charges is not a fully consistent theory. The source
of trouble can already be seen by looking at the the field energy of a point charge:
e2
Z
1
dx = ∞
8π |x|4
a2 = (a ·˙ v) − ȧ · v = −ȧ · v
1 Thermodynamics needs to be thrown in to close the equations.
10.2. RADIATION REACTION 165
2 e2 2 2 e2
P = a = − ȧ · v = −FAL · v
3 c3 3 c3
If the particle is moving at constant speed, an opposite (non-electromagnetic)
force must be applied to feed back the radiated energy. This means that the
radiation acts back on the charge and produces a force
2 e2
FAL = ȧ
3 c3
The jerk ȧ is opposite to the velocity, Fig 10.1. The force acts like a friction
force. −FAL is then the non-electromagnetic force that we need to apply to keep
the orbit stationary. It is known as the Abraham-Lorentz force. It is a somewhat
unusual force, in two ways. Unlike the usual friction it is not proportional to the
velocity but rather to the jerk. Second, it is ultimately a force that a particle
applies on itself. Something Newton would not approve of.
v a
Exercise 10.1 (Covariant form of Abraham Lorentz force). Using the fact that
Newton law
maµ = fµ
is consistent with u · u = −c2 provided f · u = 0 and a · a + u · ȧ = 0, show that
the covariant form of Abaraham-Lorentz force is
2e2 a·a
maµ = ȧ µ − uµ
3c3 c2
mc3
ω
e2
If you take for m the mass and charge of the electron, then the left hand side
is of the order of 1023 [Hz]. For most practical purposes, radiation reaction is
negligible.
10.2.2 Friction
Small forces can still do something if they act for long time. Consider the
equation of motion in a force field f with radiation reaction like friction:
2 e2
ma = f (x) + k ȧ, k= (10.1)
3 c3
From a conceptual point of view, this is a major modification of Newton law,
since the order of the equation changed: It is not enough to specify initial
conditions in the form of position and velocity, one also needs to specify the
initial acceleration. This is contrary to much evidence we have.
However, since k is so small, a reasonable strategy for interpreting the equa-
tion is look at it iteratively. Namely, replace the equation by
k k
ma = f (x) + ḟ = f (x) + (v · ∇)f
m m
In this form the equation of motion is still second order and the radiation reac-
tion looks like a friction term.
As an example consider the harmonic oscillator where f (x) = −κx. The
equation of motion is now
κk
ma = −κx − v
m
which is solved by x(t) = x(0)e±iω± t where ω is a solution of the quadratic
equation
2 κ κk
ω± − ∓ iω± 2 = 0
m m
Since k is very small, this is solved by
r
κ κ
ω± = 1 ± ik
m m
This describes reasonable solutions of slowly decaying oscillations.
Exercise 10.2. Write the equations of motion for the Kepler problem with
radiation reaction friction term.
10.2. RADIATION REACTION 167
hhhh
hh
x
The two charges communicate when x± (t + τ ) − x∓ (t) is light like. The time
delay τ = O(ε) since the dumbbell is small.
R± (τ ) = x± (τ ) − x∓ (0) is a light like vector. We choose a Lorentz frame so
that the dumbbell is at rest at time zero, i.e. q(0) = q̇(0) = 0 and so
We want to find the mutual forces on the dumbbell at time τ and express
them in terms of the acceleration and jerk at time τ , not at time 0. This gives a
funny situation where we specify the position and velocity at time 0 but specify
the acceleration at time τ . (We take the jerk to be constant.)
Let a and ȧ be the acceleration and jerk at time τ . Clearly
a(τ ) = a = a(0) + ȧτ
From this it follows that
q̇(τ ) = a(0)τ + 21 ȧτ 2 = aτ − 21 ȧτ 2 (10.3)
Integrating that gives the position
q(τ ) = 12 a(0)τ 2 + 16 ȧτ 3 = 12 aτ 2 + 1 1
ȧτ 3 = 12 aτ 2 − 31 ȧτ 3
6 − 2 (10.4)
This fixes the function q(τ ) in R± .
τ and ε are related by the condition that R is light like:
(cτ )2 = ε2 + q 2 (τ )
which is a polynomial equation for τ of order 6. However, as q(τ ) is quadratic
in τ to leading order
cτ ≈ ε
The retarded field at time τ is determined by the velocity and acceleration
of the particle at time 0:
R[µ aν] R[µ uν] 2 R[µ uν]
Fµν = −e − R·a−c (10.5)
(R · u)2 (R · u)3 (R · u)3
The four velocity
is uµ (0) = (−c, 0, 0, 0), and the four acceleration is aµ (0) =
0, a(0), 0, 0 . We therefore have
R · u = −c2 τ
The electric field in the direction of motion on one of the charges due to the
other at time τ is then
!
R0 a1 R1 u0 2 R1 u0
− Ex = F01 = −e + R · a +c
(c2 τ )2 (−c2 τ )3 (−c2 τ )3
| {z }
O(τ )
Disregarding sel-forces, we can now write Newton law assuming bare dumbbell
mass mb , in an external force Fex and radiation self-force as
e2
2
2e
mef a = mb + 3
a = ȧ + Fex
2τ c 3c3
The effective mass gets a contribution from the electromagnetic interaction.
This is very nice except that as we send τ → 0 we need to take the bare mass
large and negative.
Remark 10.3. One can get the self radiation reaction using the following trick
Amos Ori taught me. Let f (e) be the self force. Then, in the limit τ → 0 the
total force on the dumbbell is
4ε2
Ft = 2f (e) + ȧ = f (2e) = 4f (e)
3c3
2e2
This says that the radiation reaction force is f (e) = 3c3 ȧ as we have seen before.
When is it important? Small forces are important in two cases. The first
is when they are all that there is and the second if they operate for very long
time.