Lecture Notes On Electrodynamics 118120
Lecture Notes On Electrodynamics 118120
J.E. Avron1
November 8, 2024
1
Comments and typos welcome. Send to [email protected].
2
I thank my TAs Dr. Dana Levanony, Mr. Yaroslav Pollak and Barak Katzir
and Prof. Amos Ori and Dr. Oded Kenneth for all they taught me. The
students the class in 2012 pruned many typos and Daniel Klein showed me how
to query Wolfarm α.
November 8, 2024
Contents
1 Tensor calculus 9
1.1 The geometry of space time . . . . . . . . . . . . . . . . . . . . . 9
1.1.1 The metric tensor . . . . . . . . . . . . . . . . . . . . . . 10
1.1.2 Einstein summation convention . . . . . . . . . . . . . . . 12
1.1.3 Coordinate transformations . . . . . . . . . . . . . . . . . 12
1.1.4 Curvature: you may skip this . . . . . . . . . . . . . . . . 13
1.2 Vectors: Contravariant components . . . . . . . . . . . . . . . . . 14
1.2.1 Covariants components . . . . . . . . . . . . . . . . . . . 15
1.2.2 Contraction makes scalars . . . . . . . . . . . . . . . . . . 18
1.2.3 Orthogonal coordinates . . . . . . . . . . . . . . . . . . . 18
1.3 Scalars, vectors, tensors . . . . . . . . . . . . . . . . . . . . . . . 18
1.3.1 Symmetric and anti-symmetric tensors . . . . . . . . . . . 19
1.3.2 Densities and Weights . . . . . . . . . . . . . . . . . . . . 19
1.3.3 Volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.3.4 Levi-Civita tensor and symbol . . . . . . . . . . . . . . . 20
1.4 Tensors and pseudo-tensors . . . . . . . . . . . . . . . . . . . . . 22
1.5 Isometries of Euclidean space . . . . . . . . . . . . . . . . . . . . 22
1.6 Tensorial equations are coordinate free . . . . . . . . . . . . . . . 23
1.7 Differential operators . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.7.1 Grad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.7.2 Div . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.7.3 Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.7.4 Laplacian . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.8 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3
4 CONTENTS
4 Variational principle 61
4.1 Physics is where the action is . . . . . . . . . . . . . . . . . . . . 61
4.1.1 Action for a free massive particle . . . . . . . . . . . . . . 62
4.1.2 Interaction with the electromagnetic field . . . . . . . . . 63
4.1.3 Gauge invariance . . . . . . . . . . . . . . . . . . . . . . . 64
4.1.4 Euler-Lagrange . . . . . . . . . . . . . . . . . . . . . . . . 64
4.2 Variation of the action . . . . . . . . . . . . . . . . . . . . . . . . 64
4.2.1 Variation of the action of a free particle . . . . . . . . . . 65
4.2.2 Variation of the action associated to interaction . . . . . . 66
4.2.3 Euler-Lagrange equation . . . . . . . . . . . . . . . . . . . 66
4.2.4 The non-relativistic limit . . . . . . . . . . . . . . . . . . 66
4.2.5 The minimiser of the action . . . . . . . . . . . . . . . . . 67
4.3 Geodesics in Curved space-time. (You may want to skip this) . 68
4.3.1 Relativistic Kepler law . . . . . . . . . . . . . . . . . . . . 70
CONTENTS 5
4.4 Supplement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.4.1 Fermat principle . . . . . . . . . . . . . . . . . . . . . . . 72
4.4.2 Rainbow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5 Currents 75
5.1 Charge densities and currents . . . . . . . . . . . . . . . . . . . . 75
5.1.1 4-current-density . . . . . . . . . . . . . . . . . . . . . . . 75
5.1.2 Charge conservation . . . . . . . . . . . . . . . . . . . . . 77
5.1.3 Current conservation and gauge invariance . . . . . . . . 78
5.1.4 Gauge invariance and the continuity equation . . . . . . . 79
5.1.5 The continuity equation in curvilinear coordinates . . . . 79
8 Cloaking 105
8.1 Dielectric media . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
8.2 Invisible dielectrics . . . . . . . . . . . . . . . . . . . . . . . . . . 106
8.3 Cloaking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
12 Radiation 163
12.1 Wave equation with arbitrary source term . . . . . . . . . . . . . 163
12.1.1 Scalar wave generated by a moving point source . . . . . 163
12.2 Maxwell equation in the Lorenz gauge . . . . . . . . . . . . . . . 165
12.3 Lienard-Wiechert: Retarded potentials . . . . . . . . . . . . . . 166
12.3.1 The Lorenz Gauge condition . . . . . . . . . . . . . . . . 167
12.4 Lienard Wiechert formula for retarded field . . . . . . . . . . . . 168
12.4.1 Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . 169
12.5 Accelerating particle in its rest frame . . . . . . . . . . . . . . . . 170
12.5.1 The Magnetic field: . . . . . . . . . . . . . . . . . . . . . 170
12.5.2 The electric field . . . . . . . . . . . . . . . . . . . . . . . 171
12.5.3 Magnetic field in the far field region . . . . . . . . . . . . 172
12.6 Retardation from a distant source . . . . . . . . . . . . . . . . . 172
12.6.1 The dipole approximation . . . . . . . . . . . . . . . . . 173
12.6.2 Dipole approximation: Successive approximations . . . . 174
12.6.3 Radiation from a charge in Harmonic motion . . . . . . . 175
12.6.4 Many particles . . . . . . . . . . . . . . . . . . . . . . . . 176
12.7 Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
12.8 Classical instability of atoms . . . . . . . . . . . . . . . . . . . . 178
8 CONTENTS
Tensor calculus
Tensor calculus allows us to write equation without making a commitment to
a coordinate system.
9
10 CHAPTER 1. TENSOR CALCULUS
invented lad surveying tested this. Later Einstein taught us that space-time is
actually curved and there are many physical tests of this. This is a another
story.
0
−3 −2 −1 0 1 2 3 x
−1
−2
−3
The distance between two neighboring points using a standard meter is de-
noted d`: It does not depend on the choice of coordinates
X
(d`)2 = (dx)2 + (dy)2 = (dr)2 + r2 (dθ)2 = gij dxi dxj
ij
g is called the metric tensor, also known as the Riemann metric tensor. It is a
second rank tensor: It has two indices and is symmetric gij = gji . When g is a
diagonal, the coordinates are called orthogonal.
Note that the coordinates have upstairs indices while the metric has them
downstairs.
A one to one mapping
{x1 , x2 } 7→ {ξ 1 , ξ 2 }
x = ρ cos θ, y = ρ sin θ,
1.1. THE GEOMETRY OF SPACE TIME 11
90◦
120◦ 60◦
150◦ 30◦
180◦ 0◦
1 2 3 4
210◦ 330◦
240◦ 300◦
270◦
The mapping is 1 − 1 except for the origin where the point (x = 0, y = 0) maps
to a line (ρ = 0, θ). You see this in the pictures where the θ coordinates collapse
at the origin. This is a coordinate singularity, which does not reflect any bad
physical features, the space is still nice and smooth there.
The metric in the ξ coordinates, which we denote by γ is, the chain rule,
(and Pythagoras),
X X ∂xi
(d`)2 = (dxi )2 = γjk dξ j dξ k , γ = Λt Λ, Λi j = (1.1)
∂ξ j
where i is the row index and j the column index of the matrix Λ. For polar
12 CHAPTER 1. TENSOR CALCULUS
coordinates we get
cos θ −ρ sin θ 1 0
Λ= , γ=
sin θ ρ cos θ 0 ρ2
Note that
det Λ = ρ ≥ 0
so the map is invertible (and orientation preserving) except for ρ = 0. The
vanishing of the metric at ρ = 0 is a reflection of a coordinate singularity.
Remark 1.4. A choice of coordinates for the sphere, which has no finite coor-
dinate singularity is the spherical projection, shown in the figure. The image of
the south pole is at the origin, and the north pole at infinity.
The metric on the plane induced from the standard metric of the sphere is
(dx)2 + (dy)2
(ds)2 = 4R4
(R2 + x2 + y 2 )2
4R4
gjk = δjk (1.2)
(R2 + x2 + y 2 )2
Note that the coordinates (x, y) are dimensionless and det g = 0 at aa single
poiint: infinity.
If g is the metric tensor of some space (not necessarily Euclidean) in the
coordinate x and ξ is different coordinate system of the same space, then the
new metric γ is
where
∂xi
t i
γ = Λ gΛ, Λ a = (1.3)
∂ξ a
dx = vdt (1.4)
e2
dx2
dx
e1
dx1
Figure 1.4: The contravariant components are the incremets dxj of the coordi-
nates.
Exercise 1.5. In a polar coordinate system the covariant basis vectors and the
normalized unit vectors are related by
eρ = ρ̂, eθ = ρ θ̂
Remark 1.6. Normalized unit vectors are defined only for orthogonal coordi-
nate systems.
Under a change of coordinates xj 7→ ξ j , the transformation of the contravari-
ant components dxj of a vector
ei · ej = δi j (1.10)
and δ is the Kronecker symbol. The vector dx can be represented in two different
ways
dx = dxa ea = dxa ea (1.11)
Clearly
dxj = dx · ej (1.12)
16 CHAPTER 1. TENSOR CALCULUS
e2
dx2
e2
dx
e1
dx1
e1
Figure 1.5: The covariant components are the incremets dxj of the orthogonal
projections on the coordinates (schematic).
From the definitions of the metric tensor, and the notion of duality we get
dxk = dxa ea · ek = dxa ea · ek = gka dxa (1.13)
The metric tensor allows us to push indexes down.
Exercise 1.9. Show that
ej = (ej · ea ) ea
The length of the vector dx is given by
dx · dx = dxa dxb g ab = dxa dxb gab , g jk = ej · ek (1.14)
Taking the scalar product of exercise 1.9 with ek we conclude that gjk and g jk
are inversely related
δkj |{z}
= ej · ek |{z}
= (ej · ea )(ea · ek ) = g ja gak =
|{z} gj k
duality ex.10 index gym
The transformation rules for the covariant components of the position vector
follow:
dξa = γab dξ b
= γab (Λ−1 dx)b
= Λi a Λj b gij (Λ−1 )b k dxk
= Λi a Λj b (Λ−1 )b k gij dxk
= Λi a (ΛΛ−1 )j k gij dxk
= Λi a gij dxj
= Λi a dxi
= dxi Λi a
dxa = Λa i dξ i
Since all these indices can make one dizzy, let me write the two rules in a way
that make the comparison simple. If you write the covariant components as a
row vector and contravariant as a columns, the transformations can be written
in matrix and vector notation as
dξ = Λt x , dξ = Λ−1 dx (1.17)
| {z } | {z }
covariant components contravariant components
νk = uj Λj k , uk = Λk j ν j (1.18)
Note the interchange u ↔ ν and since the summation indexes j are adjacent uj
should be considered as a row vector and ν j as column.
18 CHAPTER 1. TENSOR CALCULUS
eρ · eρ = 1, eθ · eθ = ρ2 , eρ · eθ = 0
ρ ρ θ θ −2
e · e = 1, e ·e =ρ , eρ · eθ = 0
ρ̂ · ρ̂ = 1, θ̂ · θ̂ = 1, ρ̂ · θ̂ = 0
v = (rθ̇) θ̂ = θ̇
|{z} eθ = (r2 θ̇) eθ (1.21)
|{z} | {z }
velocity angular velocity “angular momentum00
one tensors because their components have a single index. The rule of transfor-
mation under coordinate change, as we have seen is:
νk = uj Λj k , uk = Λk j ν j (1.22)
Tensors are multi-index objects that transform as the product of vectors. The
number of indices of the tensor is called each rank. Each index transforms
according to whether it is up or down.
The metric tensor is an example of a symmetric second rank tensor. In
particular, the metric of a coordinate transformation of the Euclidean plane
discussed in section 1.1.1, follows this rule:
1.3.3 Volume
det g enters into the volume element dV . For the sake of simplicity, let us
consider the case of two dimensions where volume is area. Recall that the
(signed) area of the parallelogram associated with the two vectors u and v is
u×v =u∧v
e2
dx2
θ
e1
dx1
|e1 |2
|e1 ||e2 | cos θ
g=
|e1 ||e2 | cos θ |e2 |2
we conclude that p
dV = z det g dx1 dx2
Objects with such a rule of transformation are called weights. det g has weight
−2.
It is always anti-symmetric, but its numerical value is not, in general, the iden-
tity except if Λ is a rotation so det Λ = 1. For example, in polar coordinates,
the Levi-Civita tensor would be
and
1
ερρ = εθ,θ = 0, ερθ = −εθρ =
ρ
In any dimension, the Levi-Civita symbol is defined as the highest rank of com-
pletely anti-symmetric tensor with components ±1 and 0. In particular in three
dimensions
123 231
|ε = ε{z = ε312} = −ε321 = −ε213 = −ε132 = 1 (1.24)
| {z }
even permutations odd permutations
The Levi-Civita symbol is a tensor with density. The relation between the
Levi-Civita tensor (bold) and symbol is,
p i...j 1 i...j
εi...j =
|{z}
det g εi...j ,
|{z} |{z} = √det g ε|{z}
ε (1.25)
tensor symbol tensor symbol
ε1...d = 1
Remark 1.16. In even dimensions cyclic permutations are odd, while in odd
dimensions cyclic permutations are even.
εij... εij... = n!
(x0 )j = xj + aj =⇒ Λ = 1 =⇒ g 0 = g = 1,
Λt Λ = 1
det Λ2 = 1 =⇒ det Λ = ±1
Example 1.20. In a two dimensional Euclidean space there are two second rank
tensors that are invariant under rotations (up to multiplication by a scalar): The
identity and Levi-Civita
1 0 0 1
, (1.28)
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
1.7.1 Grad
The chain rule
∂ ∂xk ∂ ∂
j
= j k
= Λk j k
∂ξ ∂ξ ∂x ∂x
says that partial derivatives of scalar functions behave like covariant components
of a vector. That is, if φ(x) is a scalar field then ∇φ give the components of a
covariant vector field.
24 CHAPTER 1. TENSOR CALCULUS
1.7.2 Div
In any coordinate system
1 √
∇ · E = √ ∂j ( gE j ) (1.29)
g
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√
f
= dx2 dx ge1 · 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
The last line is in terms of the normalized coordinates.
1.7. DIFFERENTIAL OPERATORS 25
1.7.3 Curl
An important differential operator we shall need to discuss is the curl:
εijk εijk
(∇ × E)i = √ ∂j Ek = √ (∂j Ek − ∂k Ej ) (1.30)
g 2 g | {z }
anti-symmetric tensor
The term on the right is the contraction of the Levi-Civita tensor with an anti-
symmetric tensors so the object is a bona-fide vector.
Eq. 1.30 is evidently the standard definition in Cartesian coordinates. To
see why it is true in general we take Stokes law as defining property of the curl:
Z Z
dS · (∇ × E) = d` · E
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
= (dx2 e2 ) · (E2 e2 ) − ...
i
f
= (dx2 e2 ) · E − ...
i
= d` · E
1
Example 1.23. The φ components of curl in spherical coordinates (recall
Remark1.15 ) 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 1.24. Compute the (∇ × E)r and (∇ × E)r̂ .
Exercise 1.25 (Vector identities). Show the vector identities
∇ × (∇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)
1.8 Bibliography
• S. Weinberg, Gravitation and Cosmology, gives all a physicist needs to
know about tensors.
• B. Schutz
• Flanders
Chapter 2
Remark 2.1 (c is large). Practical units, like MKS, are chosen to be of O(1)
on human scale. The period of a pendulum of a meter length (about the length
of a leg) is about 2 [sec] and the second is about a heart beat. However, a light
second is about the distance to the moon. On human scale c ≈ 3 × 108 [m/s]
is essentially infinite. It is likely that ancient natural scientist entertained the
thought that c is finite rather than infinite. (Maimonides cautions against logical
pitfalls resulting from the use of infinities.) But as c is so large it is difficult
to devise an elementary experiment to measure it. The first to estimate c from
irregularities in motion of the moons of Jupiter was the Danish astronomer
27
28CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
Rømer (1644-1710). He essentially used the Doppler effect (two hundred years
before Doppler) to measure the ratio v/c where v is the velocity of earth around
the sun.
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
(t, x). It is convenient to measure space and time in the same units, e.g. measure
distances in light second, or alternatively, measure time in light-meter. With
the latter choice we write
(ct, x1 , x2 , x3 ) = {xµ }, µ = 0, . . . 3
Index convention
Greek indices µ, ν run on 0, 1, 2, 3. Roman indices j, k on 1, 2, 3
t
e
ur
t
Fu
x
t
s
Pa
Figure 2.1: Space-time: The red lines represent the light cones. The blue dot
at the origin is the event where light was emitted (and absorbed).
It allows to associate a scalar with a vector. In the case of a vector dxj relating
two nearby events in space-time the scalar is called interval
Note that (d`)2 is indefinite. When (d`)2 > 0 it has a spatial character and has
a real root measured in [cm]. When (d`)2 < 0 it has time-like character and it is
2.2. SPACE-TIME 29
then convenient to consider instead the real root of (dτ )2 > 0 which is measured
in [sec].
• A vector v µ is called space like if vµ v µ > 0;
• v µ is called time-like if vµ v µ < 0
• v µ is light-like if vµ v µ = 0
• The 1-dimensional line x = 0 is the time axis, and the clock attached to
the origin measures time τ related to the interval by (cτ )2 = −(d`)2 ≥ 0.
• The 3-dimensional portion of space time
ηµν xµ xν = 0
is a 3-dimensional light cone.
Example 2.2. In spherical coordinates (ct, r, θ, φ) Minkoski metric is
−1 0 0 0
0 1 0 0
ηµν = 0 0 2
(2.3)
r 0
2 2
0 0 0 r sin θ
t t¢
S
x
O
Figure 2.2: The backward light cone of a given point is the collection of the
events in your past that you can observe. An inertial observer that lives long
enough eventually sees all the events in Minkowski space-time.
b
S
O
a
2.3 Simultaneity
Two events X1 and X2 occur simultaneously in the lab if
∆X = X1 − X2 = (0, x1 − x2 )
(∆X)µ T̂µ = 0
(c, v)
T̂ T̂ 0
x0
Figure 2.4: The vector T̂ is the time direction in the lab and the black dot is
simultaneous with the red dot in the lab. The vector T̂ 0 is the time direction
in a moving inertial frame and the blue dot is simultaneous with the red dot in
this frame.
Two events are simultaneous in the moving frame if the 4-vector of their
difference ∆X is Minkowski orthogonal to T̂ 0 , i.e.
(∆X)µ T̂µ0 = 0
32CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
In particular, the events simultaneous with the event at the origin (0, 0, 0, 0) are
given by the points (x0 , x) that satisfy
cx0 − v · x = 0
The two straight lines marked x and x0 in Fig. 2.4 show the events simultaneous
with the origin in the two frames.
The notion of now in the lab frame and the notion of now in the inertial
frame of the clock are incompatible. Now is a well defined concept at a point in
space-time.
t Interval
x x
Figure 2.5: The Minkowski “distance” (d`)2 between the origin and an event
t = 1 in the lab, the blue line, takes negative values inside the light cone and
positive values outside. The minimum (negative value) is taken at the red dot.
This corresponds to the proper time of a clock which is at rest at the lab. A
moving clock, as shown by the green line, registers cdτ = |d`|. The intersection
with the blue line will give a smaller value for the proper time of the moving
clock.
2.3. SIMULTANEITY 33
t t0 Interval
x0
x t
Figure 2.6: The blue and red dots are the ends of a ruler of length 1 at rest in
the lab. The straight blue line is the world line of the right end. The Minkowski
distance between the origin and the blue line takes its maximum at the red dot.
The green frame is Lorentz frame where the rod is moving. x0 is the time 0 slice
in the moving frame. It intersects the blue line at a point whose Minkowski
distance is smaller than one. The proper length is largest in the rest frame.
Exercise 2.4. Suppose that you have a factory at the origin that makes identical
clocks. Explain how you can distribute the clocks while keeping them (approxi-
mately) synchronized in the Lab. (Hint: What happens to time delay if you half
the speed and double the travel time?)
Remark 2.5. In the 1970’s, when Paul Krugman, A Nobel Laureate in eco-
nomics, was a young assistant professor he wrote an amusing article about the
consequences of time-dilation in economics. It can be found here and is worth
reading.
Exercise 2.6 (Metric in light cone coordinates). Show that the Minkowski met-
ric 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
34CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
u v
Figure 2.7: Light-cone coordinates. This is not a Cartesian frame since u·v = 1.
2.4 4-Velocity
The proper time dτ is a Lorentz scalar. It is non-zero and real for a clock that
travels subluminally. We can then define the velocity as a 4-vector
dxµ
uµ = (2.5)
dτ
The length of u is always −c2 , essentially by the definition of the interval, Eq.
2.2,
dxµ dxµ (d`)2
uµ uµ = = = −c2 (2.6)
(dτ )2 (dτ )2
The 4-velocity is therefore a time-like vector which lies in the forward light cone.
It is related to the usual velocity by
dxµ dxµ dτ uµ
(c, v) = = = (2.7)
dt dτ dt γ
The direction, n = n(τ ), and so is the rapidity, φ = φ(τ ). Note that, in spite of
the notation, φ ∈ R is not an angle. Comparison with Eq. 2.7 gives the relation
between rapidity and the Newtonian velocity
vc
Rapidity
2.4.1 4 Acceleration:
The 4-acceleration can be similarly defined as
duµ
aµ = (2.10)
dτ
It is always Minkowski orthogonal to the velocity
uµ aµ = 0 (2.11)
(Since the Minkowski length of the velocity is constant). It follows that The
4-acceleration is always a space like vector.
The world line is a hyperboloid that is eventually tangent to the light cone, as
shown Fig. 2.9.
Figure 2.9: An accelerating observer who lives forever, will still see only half
the events in Minkowski space-time. He will never see the black dot on the left.
The red line is his horizon.
Exercise 2.10. What fraction of the velocity of light would you reach in this
case. Answer: tanh 1.03 = 0.77
The acceleration is
It follows that
det Λ = ±1
This divides Lorentz transformations into two classes: The proper Lorentz trans-
fromations where det Λ = 1. which contain the identity and the improper ones
where det Λ = −1 that contain the reflection.
(x0 )µ = xµ + aµ
This gives
∂(x0 )µ
Λµ ν = = δµν ⇐⇒ Λ = 1
∂xν
Translations expresses the homogeneity of Minkowski space time.
38CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
Λ(t) = eLt
2.6.3 Rotations
Rotations by θ about the x-axis and are generated by
0 0 0 0
0 0 0 0
Lyz = 0 0 0 1
(2.16)
0 0 −1 0
Remark 2.15. An airplane has three rotation controls: Stick, for pitch, rudder
for yaw, and ailerons for roll. The three are linearly independent, but by the
commutation relation you can always generate the third from the first two.
2.6.4 Boosts
A boost in the x direction is generated by
0 1 0 0
1 0 0 0
Ltx =
0
(2.20)
0 0 0
0 0 0 0
To check that φ is indeed the rapidity, consider the world-line of the origin.
Since
Λtx (φ)(ct, 0, 0, 0)t = ct(cosh φ, sinh φ, 0, 0)t
its velocity is c(cosh φ, sinh φ, 0, 0). Comparison with Eq. 2.7 shows that φ is
the rapidity.
40CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
Exercise 2.18. Check Eq. 2.20 in 1-1 dimensional space-time. (Hint: use the
formula for eφσx from quantum mechanics.
One similarly defines the generators of the boost along the y and z axis. A
boost by φ in an arbitrary direction n is then given by
The addition law for velocities in special relativity is a mess. It is, however,
simple for the rapidities: Let n be the direction of motion and L = (nx Ltx +
ny Lty + nz Ltz ). Since L commutes with itself
t0 = t + O(c−2 ), x0 = x − vt + O(c−2 )
2.6.5 Commutators
The commutators of the generators of rotations are given by cyclic permutations
of Eq. 2.19. The commutators of the generators of rotations with the generators
of boosts follow the same rules, since the three generators of boosts are naturally
associated with a vector, i.e.
t0 = t, ρ0 = ρ, z 0 = z, φ0 = φ + Ωt
2.7. ROTATING FRAMES IN MINKOWSKI SPACE 41
Exercise 2.21 (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 = ± 2 , A = πRe2
c
where Re is the earth radius and the ± depends on whether the trip was
towards the east or towards the west.
4. Compute ∆τ . (Answer: 207 [ns])
5. Explain why the result implies that one can not synchronize clocks on earth.
6. Is it still true that uµ uµ = −c2 in a rotating frame? (Yes)
7. What does get modifies is γ. Show that
γ −2 = 1 − (v/c)2 + 2Ωφ̇ρ2 /c2 − Ω2 ρ2 /c2
42CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
Figure 2.10: Minkowski space in rotating earth coordinates. It takes light dif-
ferent time, in the rotating frame, to complete a clockwise round-trip and coun-
terclockwise round-trip.
x = ρ cosh τ, t = ρ sinh τ
Since cosh τ ≥ sinh τ the Rindler coordinates cover 1/4 of space time.
2.8 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 revolution into a useful gadget.
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 its proper time with great precision1 .
Each satellite radios at specific interval, a message that contains its identity and
the reading of its clock. Since the orbit of the satellite is known, the data specify
the transmission event Xbµ of satellite b. The corresponding event received by
the user is xµ . Let us focus of the ideal case when both the transmitter and
receiver are in empty space so that electromagnetic wave propagate at c. It
follows that
(Xbµ − xµ )((Xb )µ − xµ ) = 0 (2.26)
This is an equation for the 4 unknowns xµ . To determine the four unknown
coordinates xµ of the reception event you need 4 equation. You need to receive
simultaneously, at least 4 signals from 4 GPS satellites and record 4 transmission
events all light-like. One expects that these equations are typically independent
and to have a unique solution which is the position and time of thee lost tourist.
1 To locate a point with a precision of 1 [m] you need to measure time to a precision of
O(10−9 )[sec].
44CHAPTER 2. REVIEW OF SPECIAL RELATIVITY: MINKOWSKI SPACE-TIME
Eq. 2.26 clearly incorporates special relativity as it uses the fact that light
propagates at c irrespective of the motion of the satellite and the receiver. It
turns out that for GPS to be practically useful, special relativity is not enough.
One needs to take account of the slight deviations of space-time from Minkowski.
In particular, the self-time of atomic clocks depends, not only on their velocity,
but also on the gravitational field that they see and moreover, one needs a
better approximation for the metric, that takes into account the gravitational
field of eaarth, see section 2.9.) Yet another complication is that in practice one
does wants to know the coordinate of the event in the non-inertial coordinate
system that rotates with earth (as in section 2.7.) There are many additional
complications that need to be taken care of: Atmospheric effects on the velocity
of light etc.
Ignoring special or general relativity would degrade the the accuracy of GPS
to about 10 km and make it useless. You can then satisfy yourself that special
and general relativity has been tested many billions of times.
If you want to know more about GPS, 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.
Exercise 2.23 (Orders of magnitudes).
• What is a typical velocity of GPS satellite? (Anser: 3.8 [km/s])
• 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])
• What is the resulting positioning inaccuracy?
ct
a
b
Figure 2.12: 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 .
Far from the star, when Φ = 0, the clock rate coincides with the coordinate
time rate. However, close to the star dτ < dt: the clock ticks more slowly than
the coordinate time. This means that if you moved near a very massive star,
you will outlive your friends who stayed far from it. Kip Thorn reinterpret
gravitational attraction as our wish to live longer. On earth, the difference the
gravitational metric and Minkoiwski is very small:
Example 2.25. Φ at the surface of the earth is dominated by the pull of the
sun, Φ = 2 × 10−8 and the pull of the earth is about an order of magnitude
smaller due to its pull is Φ = −1.4 × 10−9 . This adds about 1 [sec] to our life
expectancy.
Bibliography
Remark 3.1 (SI). In the (defunct) Gaussian (c.g.s.) units the unit of electric
field is rather large, 300 [V /cm], being three orders of magnitude larger than the
field near an A battery, and the unit of magnetic field is Gauss which is rather
small, on the scale of the earth magnetic field. The more practical SI units which
involve arbitrary constants such as the permittivity 0 = 8.85 × 10−12 [F/m] and
permeability µ0 = 4π × 10−7 [H/m], the unit of electric field is almost five orders
of magnitudes smaller, 1 [V /m], and the unit of magnetic field is four orders of
magnitude larger: T esla = 104 Gauss, which is comparable to the field near a
strong toy magnet. In SI units2 the Coulomb-Lorentz law is
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 Moreover, since f , v are vectors so is E and under inversion E 7→ −E. But B is a pseudo-
47
48 CHAPTER 3. THE ELECTROMAGNETIC FIELDS
The partition into electric and magnetic field is, of course, different in dif-
ferent inertial frames3 .
Exercise 3.2 (Galilei invariance). In Newtonian mechanics the force is Galilean
invariant: f = f 0 (Why). Show that the transformation rules for the fields under
Galilean transformations with relative velocity v is
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 16 components which are too many. It is not even a
symmetric second rank tensor since it has 10 components, still too many. An
anti-symmetric rank 2 tensor
Fµν = −Fνµ (3.3)
has 6 components, which is just right.
Exercise 3.3 (Symmetry is Lorentz invariant). Symmetry is a tensorial invari-
ant, e.g. if Fµν is anti-symmetric so is (F 0 )µν = Λµ α Λν β Fαβ under arbitrary
change of coordinates (and Lorentz transformation in particular). As a conse-
quence if F is anti-symmetric in Cartesian coordinates it is also anti-symmetric
in spherical coordinates.
This leaves us with the problem of how to put the two vector fields (E, B)
in Fµν . The quickest way to do this is to write the Coulomb-Lorentz law in a
way that is manifestly Lorentz invariant, i.e. as an identity between 4-vectors,
which reduces to its non-relativistic form in the approximation γ ≈ 1. Since the
Coulomb-Lorentz force in linear in the fields and in the velocity, we can get a
4-force vector by contracting the field tensor with the 4-velocity
e e
fµ = Fµν uν =⇒ f1 = F10 u0 + u
F11 1
+ F12 u2 + F13 u3 (3.4)
c c
Since u = γ(c, v) in the non-relativistic approximation γ ≈ 1 we identify the
electric field with the first row of F
F10 = E1 =⇒ Fj0 = Ej (3.5)
and the magnetic field with
F12 = B 3 =⇒ Fjk = jki B i (3.6)
In conclusion the identification of F with E and B is given by
0 −Ex −Ey −Ez
Ex 0 B z −B y
Fµν = Ey −B z
(3.7)
0 Bx
Ez B y −B x 0
3 Einstein was lead to the discovery of special relativity y considering the symmetry of
Maxwell’s equations.
3.1. ELECTROMAGNETIC FIELDS IN MINKOWSKI SPACE 49
Exercise 3.4 (Mixed components). The matrix associated with the mixed tensor
F is neither symmetric nor anti-symmetric. Verify that
0 Ex Ey Ez
Ex 0 Bz −By
F µν = Ey −Bz
(3.8)
0 Bx
Ez By −Bx 0
F = Fµν eµ ⊗ eν
Eq. 3.4 gives the relativistic Coulomb-Lorentz force 4-vector.
Euclidean Rotations
Consider a 3×3 rotation matrix R of Euclidean space so R−1 = Rt . In Euclidean
space contravariant components are the same as covariant components. The
components of matrix R are
column row
z}|{ z}|{
R j k = (Rt ) k
j
|{z} |{z}
row column
(x0 )j = F0j
0
= Λ0 µ Λj ν Fµν
= Rj k F0k
= Rj k xk (3.11)
50 CHAPTER 3. THE ELECTROMAGNETIC FIELDS
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).
Example 3.6. 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 3.7 (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 components vanish.
(Hint: It is simpler to use properties of the basis vectors eρ , ez rather than
compute the transformation matrix.)
3.2. THE FIELD OF A UNIFORMLY MOVING CHARGE 51
In the frame where we see a moving charge, everything depends on time. So let
us compute everything at t0 = 0 when the charge is at the origin. We have
x = Cx0 , y = y0 , z = z0 (3.19)
In particular
Hence, at t0 = 0
x x0
Ex (x) = e = eγ = Ex0 (x0 ) (3.22)
r3 r3 (r0 )
where r(r0 ) is the ugly expression Eq. (3.20).
For the transverse directions
Figure 3.1: The vector field of a moving charge with rapidity φ = 1. The field
is manifestly radial but not spherically symmetric.
and so
y y y0
Ey (x) = e , = Ey0 (x0 ) = γe = γe (3.24)
r3 r3 r3 (r0 )
The formula is the same but for different reasons. In one case γ came form the
field transformation and in the other from the coordinates.
It now follows that in both frames the field is radial, because
x Ex (x) E 0 (x) x0
= = x0 = 0 (3.25)
y Ey (x) Ey (x) y
Remark 3.8. This is a bit surprising. One could have argued that since the
field is radial in the rest frame, you may expect it to point in the direction of
the particle at the retarded time, not now.
2 γ2
E 0 = e2 = γ2E2 (3.26)
r4
It is stronger in the frame where the charge is seen moving (computed for the
same event).
Fµν η µν = Fµ µ = 0
3.3. LORENTZ SCALARS 53
You might be worried that we have made a sign error: E2 +B2 is proportional to
the energy density of the field. Why the minus sign? Actually, it is a fortunate
that we did not get the energy density, for, as we shall see the energy density is
a component of a second rank tensor.
The minus sign is reminiscent of the minus sign you find in Lagrangian
mechanics: The Lagrangian is the difference of the kinetic and potential energies,
not their sum. As we shall see, this is not a coincidence.
1 jk
ω∗ = ε ωjk ⇐⇒ (ω ∗ )jk = εjk ω (3.28)
2
and in 3 dimensions between vectors and anti-symmetric 2-rank tensors:
1 ijk
(ω ∗ )j = ε ωjk (3.29)
2
In 4-dimensions it is between anti-symmetric tensors and corresponds to ex-
changing the two 3-vectors associated to the tensor:
1 µναβ
(F ∗ )µν = ε Fαβ (3.30)
2
Remark 3.9. In any dimension there is a duality between anti-symmetric ten-
sors of rank r and anti-symmetric tensors of rank n−r. Anti-symmetric tensors
of rank r make a linear space whose dimension nr and since nr = n−r
r the
two linear spaces are isomorphic. The contraction of the Levi-Civita with an
anti-symmetric tensor of rank r gives an anti-symmetric tensor of rank n − r.
1.
εαβγδ εαβµν = 2(δα µ δβ ν − δα ν δβ µ )
4 For the sake of simplicity, we shall take | det g| = 1 so that Levi-Civita tensor and symbol
(F ∗ )0i = 12 ε0ijk Fjk = εijk Fjk = B i , {ijk} ∈ even permutation of {1, 2, 3}.
Similarly,
Exercise 3.11. Write the formulas in this section without assuming | det g| = 1.
For example, if there is electric field and no magnetic field in one frame, then in
any other frame E2 − B2 > 0 and E · B = 0 etc. Similarly, the field of a plane
electromagnetic wave has E2 − B2 = E · B = 0, in any frame.
3.4.1 No monopoles
The absence of magnetic monopoles is expressed by
∇ · B = ∂j (B j ) = 0 (3.35)
∂µ (F ∗ )µν = 0 (3.38)
3.5 Potentials
The homogeneous Maxwell’s equations can be rephrased as the statement that
the fields are derived from potentials. B is derived from the vector potential A:
∇ · B = 0 ⇐⇒ B = ∇ × A (3.39)
56 CHAPTER 3. THE ELECTROMAGNETIC FIELDS
You normally see this equation rearranged so the field is on one side and the
potentials on the other side.
Aµ = (−φ, A) (3.42)
∂α (F ∗ )αβ = 0 (3.43)
Show that
It follows that if k µ is light like the first term in the brackets on the left vanish.
The second term in the brackets vanishes if k and A are Lorentz orthogonal.
Since the (linear) mapping A 7→ F has a large kernel–the pure gauge fields,
it has no inverse and F does not determine the 4-potential A uniquely. For any
(scalar) function Λ of space-time
Aµ → Aµ + ∂µ Λ (3.48)
give the same field F . Fields have a direct physical meaning: they can be mea-
sured as forces. Potentials are tools for computations and no physical instrument
measure potentials. In particular, a voltmeter does not measure φ.
xΜ
xΝ
Figure 3.2: Loop and surface element for Stokes.
58 CHAPTER 3. THE ELECTROMAGNETIC FIELDS
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.
A familiar, special case of the formula is a closed loop at fixed time, dxµ =
(0, dxk ), where
I I Z Z
Ak dxk = A · d` = ∇ × A · dS = B · dS = Φ
and define the dual so that F ∗ is a tensor, (not a density). This means that we
replace Eq. 3.30 by
1
(F ∗ )µν = p εµναβ Fαβ (3.52)
2 |g|
From the rhs of Eq.3.44 we have the Bianchi-like identity
p
0 = ∂µ ( |g|(F ∗ )µν ) (3.53)
This gives Faraday law. In the last line we used the fact that curvilinear co-
ordinates for Maxwell equations in an inertial frame, have a g which is time-
independent.
It is quite remarkable that Electrodynamics in curvilinear coordinates can
be formulated with no reference to Christoffel symbols.
Bibliography
1. F. Hehl and Y. Obukhov, A gentle introduction into the foundations of
electrodynamics
60 CHAPTER 3. THE ELECTROMAGNETIC FIELDS
Chapter 4
Variational principle
• It is elegant
The property of being minimizer does not depend of the the choice of co-
ordinates and so the Lagrangian formulation guarantees the tensorial character
of the equations. Since Lorentz transformations are special coordinate trans-
formations of Minkowski space, a theory is guaranteed to be Lorentz invariant
once the action is a Lorentz scalar.
In this section, we shall see how one formulates relativistic mechanics using
Lagrangian formalism. The algorithm for doing that is:
61
62 CHAPTER 4. VARIATIONAL PRINCIPLE
Figure 4.1: The minimum at the point p does not depend on what coordinates
you choose for the x axis)
• Verify that in the non-relativistic limit the action reduces to its Newtonian
form
This does not fix the relativistic action uniquely. This is a feature that allows
for creativity to be involved.
Figure 4.2: The world line is required to have time-like tangents–for the action
to be real. 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.
The Lagrangian
mc2
L=−
γ
= −mc2 (1 − v2 /c2 )−1/2
m
mc2 +
≈ − |{z} v·v (4.2)
const |2 {z }
kinetic energy
Use this to drive the equations of motion of a free particle in a rotating frame.
A0µ = Aµ − ∂µ Λ (4.6)
This leads to
Z
0 e e
(∂µ Λ) dxµ = Sint −
Sint = Sint − Λ(xf ) − Λ(xi ) (4.7)
c c
Although Sint changes under a gauge transformation, the change is fully de-
termined by the end points of the orbit. Variations of the orbit that keep the
endpoints fixed do not affect the the action. This guarantees that the Euler
Lagrange equations are gauge invariant.
4.1.4 Euler-Lagrange
S is a Lorentz scalar, of course. If we want to use the ready made Euler-Lagrange
equations, familar from mechanics,
d ∂L ∂L
= (4.8)
dt ∂v ∂x
we nee to write Z
S= Ldt, L = L(x, v) (4.9)
In fact
e
f (x, u) = −mc2 + Aµ (x)uµ
c
4.2. VARIATION OF THE ACTION 65
δx = { δx0 (τ ), . . . δx3 (τ ) }
| {z }
inf initesimal f unctions
δx vanishes at the end points. The strategy one uses to derive the Euler-
Lagrange equations is to use integration by parts to bring δS to the form
Z xf
x
δS = hµ (x, u)δxµ xfi + gµ (x, u, u̇)δxµ dτ
xi
Since δx vanish at the end points the boundary term drops. And since δxµ are
µ
gµ (x, u, u̇) = 0
Euler-Lagrange equation follow by setting the boundary term and the variation
to zero. This gives
mu̇µ = 0 (4.14)
Recall that the momentum in classical mechanics is defined as the rate of change
of the action due to change of the end-point of a classical path. This means that
we look at the boundary term when we set mu̇ = 0 in the integral in Eq.4.13,
namely
δSp = pµ δxµ = m uµ δxµ (4.15)
The covariant components of the momentum are the gradient of the action with
respect to the end points.
1 Viewed as functions of τ the variations satisfy the constraint: (dδxµ )(dδxµ ) = −(cdτ )2
66 CHAPTER 4. VARIATIONAL PRINCIPLE
The basic idea in the calculus of variation is always 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 = (4.16)
c c
where we introduced the second rank tensor Fµν to describe the electromagnetic
fields fields E and B.
The action has a positive contribution from the kinetic energy and a negative
contribution from the potential energy. Each of the two can be arbitrarily large.
This suggests that S is actually unbounded below when considered as a function
of paths. This is indeed the case.
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”.
The action is
Z π Z π
−
S= Ldt = A2 (ω 2 cos2 ωt − sin2 ωt)dt −−−−→ 2x0 A
0 0 A→∞
For example, the function in Fig. 4.1 is convex. It is evident that if a function
is convex its minimum is unique. (It may, however, lie at infinity).
The notion of convexity extends to the case that x, the argument of S, is
itself a function–a path. γ is a convex function of v. It follows that the action
is a convex function of the path. This then implies that the minimizer for a free
relativistic particle is unique.
d µ
0= (u gµν uν ) = 2u̇µ uµ + (∂α gµν )uµ uν uα = 0 (4.22)
dτ
4.3. GEODESICS IN CURVED SPACE-TIME. (YOU MAY WANT TO SKIP THIS) 69
We want to find the path that minimizes the action (equivalently, maximizes
the proper time) Z
S = −mc2 dτ
δ(cdτ )2 = 2c2 (dτ )δ(dτ ) = −(δgµν ) dxµ dxν − 2gµν δ(dxµ )dxν
Hence
− 2c2 δ(dτ ) = (δgµν ) uν dxµ + 2gµν uν δ(dxµ ) (4.23)
Rewrite the first term on the right as
(δgµν ) uν dxµ = (∂α gµν ) uν dxµ δxα = (∂µ gαν ) uν dxα δxµ
The first term is a boundary term which does not contribute to the variation of
the action. Dividing by dτ the brackets on the right, and renaming the dummy
index ν, β give
gµν u̇ν + ∂α gµβ − 12 ∂µ gαβ uβ uα = 0 (4.24)
where Γ is linear in the derivatives of g. Since only the symmetric part of Γµαβ
contributes, we define Γµαβ so it is explicitly symmetric
Γµαβ = 12 g µν ∂α gνβ + ∂β gνα − ∂ν gαβ (4.26)
Exercise 4.5. Show that great circles on the sphere are geodesics.
uρ = ρ̇ = 0, uθ = θ̇ = ω, ut = cṫ = cγ (4.30)
− c2 = uµ uµ = (cγ)2 gt + ω 2 ρ2 (4.31)
Now, since u̇µ = 0 for a stationary geodesic, the geodesic (differential) equations
reduce to constraints relating ρ, ω, γ. As uµ has only two non-zero components
the 4 geodesic equations are
The (non-zero) Christoffel symbols are Γρtt and Γρθθ so the geodesic equation
gives a single constraint relating ω, γ and ρ
Combined with Eq. 4.31 which relates γ and ω, we get a relation between ρ and
ω for circular (time-like) stationary orbits
c2
ωτ2 = (4.36)
gtt Γρθθ /Γρtt − ρ2
2 When g is diagonal, we denote its diagonal elements by gµ
4.3. GEODESICS IN CURVED SPACE-TIME. (YOU MAY WANT TO SKIP THIS) 71
where we used the metric and orbit to relate self and coordinate time.
1 1 1
Γρtt = − g ρ ∂ρ gtt , Γρθθ = − g ρ ∂ρ gθθ = − g ρ ∂ρ ρ2 = −ρg ρ
(4.40)
2 2 2
Hence
Γρθθ 2ρ
=
Γρtt ∂ρ gt
2GM
gt = −(1 + Φ), Φ=− (4.42)
c2 ρ
GM
ωτ2 = (4.43)
ρ3
3GM
<1
ρ
72 CHAPTER 4. VARIATIONAL PRINCIPLE
4.4 Supplement
4.4.1 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:
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 4.7. Show that the equation of motion is consistent with the t being
a unite vector.
n1 sin θ1 = n2 sin θ2
4.4.2 Rainbow
The simplest features of the rainbow can be understood from Snell’s law.
Exercise 4.9. 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.
4.4. SUPPLEMENT 73
Θ1 n1
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.
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 4.10. 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).
74 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.
Chapter 5
Currents
The sources of electromagnetic fields are the charges and currents. To describe
the sources in a Lorentz invariant framework we need to amalgamate the non-
relativistic notions of charge and currents into a single notion in space-time the
4-current.
5.1.1 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 coor-
dinates. We can make a 4-current density using only scalars 4-vectors and in
75
76 CHAPTER 5. CURRENTS
where dot is a derivative with respect to the proper time. The scalar factor ec
fixes the dimensions to the dimensions of current density.
Remark 5.1. dτ is a scalar, and ξ˙µ a 4-vector. The delta function is a density,
i.e. under a coordinate transformation it is multiplied by a power of det g.
Indeed, under scaling x0 = λx, the metric transforms as λ2 gµν
0
= gµν , so in d
2d 0 d 0 d −d
dimensions λ det g = det g. Since δ (x ) = δ (λx) = λ δ(x) we see that δ
is a density as it transforms by √1 0 δ(x0 ) = √1 δ(x). The current density is a
|g | |g|
density as its name suggest.
ct
To relate this expression to Eq. (5.2) integrate over τ . This gets 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 cdτ to
dξ 0
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(ξ 0 ) δ(ct − ξ 0 )v µ (ξ 0 )
ct
Figure 5.2: 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.
The inhomogeneous
Maxwell’s equations
81
82 CHAPTER 6. THE INHOMOGENEOUS MAXWELL’S EQUATIONS
Figure 6.1: 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.
We can rule out the action Ssc by the following observation: The homogeneous
Maxwell equation
This means that the associated action is a boundary term. Since the rules of
variation keep the boundary terms fixed, the variation of Scs vanishes identically.
We are left with the first candidate. We need first to justify the sign chosen
so that the action will have a minimum rather than a maximum.
2 So we can add it to Sp and Sint
6.2. VARIATION OF THE FIELD: RULES OF THE GAME 83
In Lagrangian mechanics the kinetic energy comes with a positive sign. Since
E is linear in Ȧ while B = ∇ × A, it is the E2 that plays the role of kinetic
energy and as
F · F = −2(E2 − B2 ) (6.1)
we must have Z
1
SF = − F · F dΩ (6.2)
16πc
The 16π gives Maxwell equations in c.g.s units and in particular leads to the
Coulomb potential in the form3 re . In the MKS system where Coulomb law is
e/4π0 r one needs to replace 16π by 40 .
δ(Fµν F µν ) = 2F µν δ(Fµν )
where
δ(Fµν ) = ∂µ δAν − ∂ν δAµ
By the anti-symmetry of F
Since det |g| is a function of the coordinates, and not a function of A, it is not
affected by the variation. From Eq. 6.3
p p p
|g|δ(Fµν F µν ) = 4∂µ ( |g|F µν δAν ) − 4∂µ ( |g|F µν ) δAν
3 Replacing 16π by 4 in the action, gives for the Coulomb potential e/4πr.
84 CHAPTER 6. THE INHOMOGENEOUS MAXWELL’S EQUATIONS
p
The first term is the divergence of the vector field |g|F µν δAν which can be
converted to a 4-surface integral on the boundary of the box where δA = 0.
Hence, Z
1 1 p
δSF = dΩ p ∂µ ( |g|F µν ) δAν (6.5)
4πc |g|
| {z }
divergence of anti-symmetric tensor
Maxwell’s equations in free space follow from δSF = 0 for any δAν . This is the
case if
1 p
p ∂µ ( |g|F µν ) = 0 (6.6)
|g|
The variation will vanish for arbitrary δA provided the brackets vanish. In the
case of curvilinear coordinates where det |g| is not necessarily 1 we have4
1 p 4π
p ∂µ |g|F νµ = jν (6.9)
|g| c
These are the 4-inhomogeneous Maxwell equations in a neat and concise form.
in accordance with Eq. 5.11. 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
Ė = ∇ × B − J, Ḃ = −∇ × E (6.13)
c } | {z }
F araday
| {z
Ampere
and dot denotes partial derivative with respect to ct. The vector equations are
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.
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.
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.
6.3. NEW PHYSICS 87
and they are consistent if the source satisfies the continuity equation.
-1
-2
-2 -1 0 1 2
6.4.1 Axion
In Maxwell’s theory the source term is a the vector field of currents j µ . In 1+1
dimensions there is a different option for a source term, namely a scalar field
φ(x) 5 :
1 1
L = − Fµν F µν + φ(x)εµν Fµν
4 2
This looks first like a different theory, but it is actually equivalent to Maxwell’s.
The variation of A gives, up to boundary terms,
δL = (∂µ F µν )δAν − ∂µ φ(x)εµν δAν
The Euler-Lagrange equations for this model are then
∂µ F µν = j ν , j ν = (∂µ φ)εµν (6.15)
Note that ∂ν j µ = 0 so the current is conserved. We have recovered Maxwell
theory except that the current is interpreted as the gradient of a scalar.
∂B (cρ) = σH (6.18)
Figure 6.3: 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.
• It has broken time reversal and space inversion, symmetries that are bro-
ken by the large external magnetic field
0 = ∂µ j µ , j = (cρ, j 1 , j 2 ) (6.20)
field. To make contact with the Hall effect we identify the vector field with the
electromagnetic gauge field A
j α = kαβγ ∂β Aγ (6.21)
k αβγ k αβγ
LCS 7→ LCS + ε F βγ ∂α Λ = LCS + ε ∂α ∂β A γ Λ (6.24)
8πc2 4πc2
The CS Lagrangian is that it breaks both time reversal symmetry and parity.
You can see this either from the fact that it is first order in the derivatives, or
from the Levi-Civita tensor. This is a reflection of the symmetry of the Hall
effect where the external magnetic field breaks both symmetries.
To find the equations of motion for CS consider first the variation of the
action
The interaction term is the same as in Maxwell theory, and the full Lagrangian
is
1
LCS + 2 Aα j α
c
The Euler-Lagrange equations are
k ∗
F +j =0 (6.25)
2π
Unlike Maxwell’s equations, this is not a set of differential equation, but an
algebraic relation between the fields and sources. Comparing with Eqs. 6.19
gives
k = −2πσH (6.26)
It follows that Z Z
1
2 dx3 Fµν F µν = dSα εαβγ Aβ Fγδ (6.29)
Quantization
To explain why k must be quantized one needs to input quantum mechanics and
also assume that the two dimensional space of the system is a closed mmanifold,
e.g. a two dimensional torus.
In quantum mechanics one allows the system to explore all configurations.
A configuration is weighted by a complex phase
eiS/~
To examine this condition, we need to discuss how Λ enters into quantum me-
chanics. Λ affects the phase of the wave function. Since [Λ] = [e] and phase is
92 CHAPTER 6. THE INHOMOGENEOUS MAXWELL’S EQUATIONS
∇ · E = α(∇φ) · B
Ampere law
∂0 E − ∇ × B = αφ̇B + α∇φ × E
Exercise 6.6. Verify.
When φ is a constant one recovers the sourceless Maxwell equations. In
general, ∂µ φ acts like a source term in Maxwell equations.
φ is also odd under time reversal. The notion of time reversal in the quantum case is subtle.
94 CHAPTER 6. THE INHOMOGENEOUS MAXWELL’S EQUATIONS
Φ=0
Φ=Π
Figure 6.4: The interface between two insulators that are topologically distinct
gives rise to a singular Axion field.
similar to the one in CS theory of the quantum Hall effect. By definition, the
two insulators are topologically distinct if the constant is different in each. This
means that ∇φ = δ (2) (x)n. Gauss law is replaced by
∇ · 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.
Ampere law is replaced by
∂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
∇ · E = αδ (2 (x)n · B, −∇ × B = αδ (2) (x)n × E
B
E
Exercise 6.7. Use Stokes theorem for the (blue) rectangle shown in the figure
to show that
B0 = αE0
x + dẑ
B(x) = g θ(z) + (yet unknown f unction)θ(−z)
|x + dẑ|3
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
96 CHAPTER 6. THE INHOMOGENEOUS MAXWELL’S EQUATIONS
E, B
Φ=0
Φ=Π
E=0
Figure 6.6: 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.
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
The problem then reduces to solving Laplace equation with two types of bound-
ary conditions.
Bibliography Xiao-Liang Qi, et al, Inducing a Magnetic Monopole with
Topological Surface States, Science 323, 1184 (2009);
6.6. SUPPLEMENT: AXION ELECTRODYNAMICS 97
E B
Image Image
Figure 6.7: 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.
98 CHAPTER 6. THE INHOMOGENEOUS MAXWELL’S EQUATIONS
Chapter 7
So far we considered the electric field and the magnetic induction (E, B), and
derived Maxwell equations from Lorentz invariance. Let us now turn to (D, H)
known as the displacement field and magnetic field.
One good reason to introduce two different notions of electric fields, (E, D)
and two different notions of magnetic fields (B, H) is that they are associated
with a-priory different measurements. (E, B) are defined by measuring the
force and then using Coulomb-Lorentz law to determine (E, B). D is defined
by measuring the charge on a pair of metallic plates, as in Fig. 7.1 and H by
measuring the surface current on a thin superconductor, as in Fig. 7.2.
A second good reason is that they are associated with different equations:
(E, B) are associated with the homogeneous Maxwell equations
∇ · B = 0, Ḃ + ∇ × E = 0 (7.1)
+
−
Figure 7.1: Put two thin metallic plates in contact in the field. The surface
charge density on the plates is proportional to field strength. Separate the
plates and measure the charge on the top plate. Define the field D as the
maximal charge per unit area over all initial orientations of the plates.
99
100 CHAPTER 7. MAGNETIC FIELDS AND MAGNETIC INDUCTION
Figure 7.2: A thin superconducting cylinder expels the magnetic field by creat-
ing surface currents. Measuring the current gives H.
4π
∇ · D = 4πρ, −Ḋ + ∇ × H = J (7.2)
c
In a stationary case D = 0 inside a metal and has vanishing tangential com-
ponent, Dk = 0, on the surface. Hence by Eq. 7.2, on the surface charge on
a metallic surface D⊥ is, up to a factor 4π, proportional to the surface charge
density. This explains why measuring the charges on the plates in Fig. 7.1 is a
measure of D. Similarly, in a stationary case H = 0 inside a superconductor,
and Hk is proportional to the surface current. This explains why measuring the
current in Fig. 7.2 is a measurement of H = 0.
The in-homogeneous equations follow from charge conservation: Given ρ,
define D by the solution of the differential equation
∇ · D = 4πρ (7.3)
x−y
Z
D(x, t) = d3 y ρ(y, t) (7.4)
|x − y|3
Now, use this Eq. 7.3 in the equation for charge conservation
c 4π
0 = ∂t ρ + ∇ · J = ∇ · Ḋ + J (7.5)
4π c
Since the brackets have zero divergence it is the a curl of something. This
something is H
4π
Ḋ + J=∇×H (7.6)
c
This is the inhomogeneous Maxwell’s equations.
7.1. CONSTITUTIVE RELATIONS 101
and D and H in D
0 Dx Dy Dz
−Dx 0 Hz −Hy
Dµν =
−Dy
(7.8)
−Hz 0 Hx
−Dz Hy −Hx 0
4π µ
∂µ (F ∗ )µν = 0, ∂µ Dµν = j (7.9)
c
To close the set we need a constitutive relation between F and D. The general
form of such a relation is
Dµν = εµναβ
0 Fαβ (7.10)
εµναβ
0 is the permittivity (not to be confused with Levi-Civita). It is a tensor
of rank 4, which is anti-symmetric in the first pair of indices and the last pair
and so has, in principle, 36 components. It is a property of the material. In the
rest frame of the material it is a function of position, reflecting the composition
of the material, but not a function of time if the material is in equilibrium. It
is a generalization of the familiar relation
describing the relation between the electric displacement field and the electric
field in a dielectric and between the magnetic induction and the magnetic field
in a magnetic materials. Under Lorentz transformations ε0 transforms like a
4-th rank tensor. This reflects the fact that a constitutive relation normally
breaks Lorentz invariance since it is a property of a medium that has a rest
frame.
In the case of free space the constitutive relation is the same in all Lorentz
frames. Indeed, since in vacuum F µν = Dµν we have
Dµν = εµναβ
0 Fαβ = η µα η νβ Fαβ (7.12)
It follows that
1 µα νβ
εµναβ η η − η να η µβ
0 = (7.13)
2
102 CHAPTER 7. MAGNETIC FIELDS AND MAGNETIC INDUCTION
and we wrote ε0 so that it is explicitly unti-symmetric in the first and last pair of
indices. The tensor ε0 plays the role of the identity for 4-th rank anti-symmetric
tensors. It is invariant under Lorentz transformations, since η is.
This relation for E and D takes the form
E j = Dj = (ε0 )jk Ek = η jk Ek (7.14)
This gives the Minkowski metric plays the interpretation of a dielectric constant
of the vacuum. This observation is the starting point of the theory of cloaking.
• The homogenoues Maxwell equations Eq. 7.16 for the six fields (E, B)
should be thought of as one vector valued evolution for Ḃ and a constraint
on the initial data.
• The in-homogenoues Maxwell equations Eq. 7.20 for the six fields (D, H)
should be thought of as one vector valued evolution for Ḋ and a constraint
on the source term: The continuity equation.
• The missing 6 equations are the constitutive relations
where the tensor ε characterized the material. For the vacuum ε is given
by Eq. 7.13.
• In the general ε must be viewed, in general, as a tensorial linear operator.
If the material is homogeneous and translation invariant, then a transla-
tion invariant linear operator is a convolutions in space time. In the case
that the material is also memory-less and has zero-range correlations, ε
reduces to a constant. This is the case for the vacuum1 .
Since D, E are even under time reversal while B, H are odd, when time-reversal
is a symmetry the constitutive relation takes the form
In the case of isotropic media εjl = ε0 δ jk and µjk = µ0 δjk are proportional to
the identity.
Exercise 7.1. Show that the microscopic charge distribution of a homogeneously
polarized sphere of radius a is concentrated on the surface with surface density
P · x̂ δ(|x| − a)
M × x̂ δ(|x| − a)
P P
Cloaking
Dot stands for derivative with respect to x0 . In the absence of external sources,
the inhomogeneous Maxwell equations, Gauss and Ampere laws, in a dielectric
are
1 √ εjk`
∇ · D = √ ∂j ( gDj ) = 0, Ḋj − √ ∂k H` = 0 (inhomogeneous)
g g
There are twelve unknown fields, and six evolution equations and two con-
straints. The missing equations are the constitutive relations2
Dj = εjk Ek , B j = µjk Hk
where ε and µ are tensors. We can then write for Maxwell equations for di-
electrics, in the absence of sources as equations for E, H
√ εjk`
∂j ( gµjk Hk ) = 0, µjk Ḣk + √ ∂k E` = 0 (8.1)
g
and
√ εjk`
∂j ( gεjk Ek ) = 0, εjk Ėk − √ ∂k H` = 0 (8.2)
g
1 In Landau Lifshitz, E and B represent averages over a macroscopical small, but micro-
tensor.
105
106 CHAPTER 8. CLOAKING
The vacuum may then be thought of as a dielectric where the metric tensor is
the dielectric constant, in c.g.s, is:
ε=µ=g
where g is the (Euclidean) metric. In a Euclidean metric g is proportional to the
identity and in any metric g j k = δkj . For example, in a two dimensional space
described by cylindrical coordinates the covariant components of the metric and
contravariant components of the dielectric tensors are
1 0 1 0
g= , ε=µ= (8.3)
0 ρ2 0 1/ρ2
In practice, it is often the case that µ ≈ g. Then Maxwell’s equations take the
(approximate) form
√ √ jk
∂j ( gg jk Hk ) = 0, gg Ḣk + εjk` ∂k E` = 0 (8.4)
and
√ √
∂j ( gεjk Ek ) = 0, gεjk Ėk − εjk` ∂k H` = 0 (8.5)
Figure 8.1: Reflection from a mirror can be minimized when the reflection from
the front and back interfere destructively. This works for specific directions and
wavelengths.
This means that as far as Eq. 8.5 is concerned the dielectric behaves as if its is
a coordinate transformation of free space. This is also true for Eq. 8.4 which is
independent of ε, since it holds in any curvilinear coordinate system. If ε 6= g
in a finite ball, the coordinate transformation is restricted to the ball and one
gets a picture such as in Fig. 8.3. We have thus created an invisible dielectric.
Figure 8.2: The figure illustrates a local coordinate transformation of the Eu-
clidean plane, so that the image of straight lines become the curves in the
figure.Such a coordinate transformation can be implemented by a choice of a
suitable dielectric tensor ε. The corresponding dielectric, the reddish disk, is
invisible.
ρ = r + e−r − 1, θ 7→ θ (8.9)
8.3 Cloaking
Now that we know how to make dielectrics that are invisible3 the next challenge
is to engineer dielectrics that can cloak arbitrary objects.
The basic idea behind cloaking is a singular coordinate transformation, that
maps the exterior of a ball in physical space to the Euclidean space (with the
origin removed).
This is best illustrated by working out through an example in the plane.
Let r be the radial coordinate of physical space, which hosts a dielectric tensor
ε(r). Consider the mapping from the plane with radial coordinates (ρ, θ) into
the plane with radial coordinates (r, θ)
r 2 = ρ2 + 1 (8.10)
This maps the entire plane ρ ≥ 0 to the r-plane minus a disk, namely r ≥ 1.
Physical space (r, θ) is the Euclidean plane with the standard Eclidean polar
metric a g. The induced metric γ on the (ρ, θ) plane is, by Eq. 8.7
2
∂r ρ 2
γρρ = = , γθθ = r2 , det γ = ρ2 (8.11)
∂ρ r
Maxwell’s equation 8.5 is fully determined by the functions
2 2
√ ρρ r r2 1 √ θθ 1 ρ
γγ = ρ = =ρ+ , γγ = ρ = 2 (8.12)
ρ ρ ρ r ρ +1
The corresponding equation in the physical space with a dielectric is
√ rr √ θθ
gε = rεrr gε = rεθθ (8.13)
The differential equations in the two spaces are the same equations up to re-
naming ρ ↔ r if the functions are the same, i.e.
1 r
rεrr = r + , rεθθ = , r>1 (8.14)
r r2 + 1
This is the equation that makes ε a cloaking material. After simplification
1 1 1 1
εrr = 1 + 2
= g rr + 2 , εθθ = = g θθ − 2 2 , for r > 1 (8.15)
r r r2 +1 r (r + 1)
The second term makes it manifest that for r 1 the dielectric is surrounded
by the vacuum. The dielectric ε for r < 1 can be arbitrary and it does not affect
the solutions of Maxwell’s equations outside the ball r > 1. This shows that the
enveloping dielectric for r > 1 is cloaking the inside.
Of course, a difficult issue we have not addressed is how to engineer ε. In
fact, you’d expect a cloaking dielectric to be a wiered material. Landau and
Lifshits argue that ordinary materials at thermal equilibrium have
εjk ≥ g jk (8.16)
8.3. CLOAKING 109
Bibliography :
J. B. Pendry et. al .“Controlling electromagnetic fields”, Science 312, (2006)
Figure 9.1: The stress tensor represents the flow of momentum (red arrow)
through a cross section in space-time (blue line)
111
112 CHAPTER 9. THE STRESS-ENERGY TENSOR
single tensor. Therefore energy density and momentum density will be members
of the same tensor, and so would be the density of energy currents and density
of momentum currents. T 00 wil stand for the energy density and T 0j for the
energy current in the j-th direction. T j0 stand for the j-th component of the
momentum density and T jk for the j-th component of the momentum current
in the k-th direction.
The second observation we make is that T µν must be quadratic in the field
F . This is because the energy density and Poynting vector have this form.
There are three second rank tensors that we can make that have these prop-
erties, using F , η and ε. T should therefore be a linear combination of
Explicitly, T is
0 B×E
E2 + B2
1
4πT = −
B×E
2 |{z} Ei Ej + Bi Bj
relative sign
You recognize the energy density and Poynting vector. We shall discuss the
other terms below.
∂µ T µν = ∂µ T νµ (9.7)
since T is symmetric. The loss of momentum of the field is the gain of mo-
mentum of the particle, in the case of continuous charge distribution, we must
have,
1
∂µ T µν = jµ F µν (9.8)
c
T has dimension of energy density (in space) and ∂T has dimension of energy
density (in space-time). This gives the rhs of Eq. 9.8 the interpretation of source
of energy per unit of space-time volume. The identity follows from Maxwell’s
equations. Since this is a subtle and computation let us first observe that since
T is bi-linear in F and Maxwell’s equations are linear with j its source term the
rhs has the structure one expects.
As preparation let us fist compute the divergence of the first term in T
∂µ (F µα F ν α ) = (∂µ F µα )F ν α + F µα ∂µ F ν α
4π
= − j α F ν α + F µα ∂µ F ν α
c
4π
= jµ F µν + Fµα ∂ µ F να
c
4π
= jµ F µν + Fαβ ∂ α F νβ
c
We have used the the in-homogeneous Maxwell Eq. 6.8 in the second line, moved
indexes up and down in the third and renamed µ 7→ α and α 7→ β in the last
line. For the second term we have
∂µ (η µν F · F ) = 2Fαβ ∂ ν F αβ
= −2Fαβ (∂ α F βν + ∂ β F να )
= −2Fαβ ∂ α F βν − 2Fβα ∂ α F νβ
= 4Fαβ ∂ α F νβ
where I have used the homogeneous Maxwell equation in the second line in the
form of Exercise 3.15. In the third line I replaced α ↔ β. Putting these in the
equation for T µν we get Eq. 9.8.
Exercise 9.1 (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
ct
Figure 9.2: The field is represented by the green ellipse and the space-time box
by the red rectangle
is large enough to embrace all the fields at any given time. Then
Z Z Z t2
Z t2
0 = dΩ ∂µ T µν = dSµ T µν = dS0 T 0ν = dV T 0ν
t1 t1
E2 + B2 B×E
T 00 = , T 0j = (9.10)
8π 4π
The Poynting vector is the momentum density up to factor c.
It may be worthwhile to note that even though T is symmetric the physical
interpretation of T µν is different from the interpretation of T νµ . For exam-
ple T 0j is interpreted as the momentum density (up to factor c) while T j0 is
interpreted as the energy flux.
9.3. STRESS TENSOR 115
Figure 9.3: You need to apply an external force to hold the capacitor plates
from collapsing on each other.
The term T k3 gives the density of the k-th momentum in the space-time volume
element
dP k = T k3 dx1 dx2 dt (9.12)
Since the force is the rate of momentum the k-th component of the force dF k
acting on the 3-surface dx1 dx2 is
dF k = T k3 dx1 dx2
This gives T jk the same meaning as the stress tensor in the theory of elasticity.
E2
Z
1
dzj0 F 03 = 12 σEz = − z (9.15)
c 8π
The force per unit area pulls the upper plate down, indicated by the minus sign,
Ez2
− (9.16)
8π
Using the lhs of Eq. 9.8: The stress T zz is
E2
(
zz − 8πz bottom plate
T = (9.17)
0 top plate
The rate of flow of z-momentum out of the green box per unit area is
Ez2
(9.19)
8π
Since the box is loosing z-momentum, the force in the z-direction is negative.
As the green box can be shrunk to hug the top plate, the force on the plate
computed in both ways agree.
Exercise 9.3. Compute the components of T for the example 6.2.Show that the
current carrying wire pumps energy into the electromagnetic field at a constant
rate. Where does the energy come from and where is it dumped?
9.5. THE STRESS TENSOR AS VARIATION OF THE METRIC 117
Figure 9.4: 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 9.5: The magnetic field lines of a solenoid run parallel to the solenoid
axis. As a consequence the stress in the radial direction T ρρ > 0 inside the
solenoid. 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.
When Maxwell constructed his theory the queen of science was, of course, me-
chanics. In particular, he understood well elasticity and fluid mechanics. In
elasticity theory the concepts of stress and strain are important, and it was
natural for Maxwell to ask what is their analogs in electrodynamics. One can
think of a strain as a deformation of the metric. For example, the strain shown
in the figure 1 can be represented by deformation of the Euclidean metric
1 0 1+ 0
g= →g=
0 1 0 1−
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 variation δg. To
do that observe that,
X Pµ
g −1 =
γµ
and so
X Pµ (Pν δγν + γν δPν )
g −1 δg =
γµ
and so finally
p p p p
δ |g| = |g| δ log |g| = 12 |g|g γβ δgβγ (9.21)
t t t
x x x
Figure 9.7: 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.
in a a fixed box, has been computed in Eq. 6.5 and was found to be
Z Z
µν
4πc (δSF ) = − dΩ ∂µ (F δAν ) + dΩ(∂µ F µν )δAν
| {z } | {z }
bdry term Euler-Lagrange
(9.23)
For fields that satisfy the Euler-Lagrange equation, only the left term matter–
the variation is a boundary term
Z
4πc (δSF ) = − dΩ ∂µ (F µν δAν ) (9.24)
= ∂µ F µν Fαν δξ α + F µν ∂µν Aα δξ α
= ∂µ F µν Fαν δξ α
We see that the change in action due to shifting the field is:
Z
4πc (δSF ) = dΩ ∂µ F µν Fαν δξ α
(9.26)
Z
= dΩ ∂µ F µν Fαν − 41 gαµ (F · F ) δξ α
Z
= dΩ ∂µ T µα δξα
Since this is supposed to hold for any (infinitesimal) box and any shift, we get
the conservation law (in the absence off source currents).
1 αµ
T αα = F Fαµ − 41 η α α F µν Fµν = 0
4π
9.7 Applications
9.7.1 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 = (9.28)
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 (9.29)
4R2 c
The gravitational force on a particle with density ρ is
4πρr3 M G
Fgravity = (9.30)
3R2
Where M = 2 × 1033 [gram] and Newton constant G = 6.7 × 10−8 [cgs]. The
ratio of the two is then
Fradiation 3L 0.06 [mgr/cm2 ]
= ≈ (9.31)
Fgravity 16πrρcM G ρr
9.7. APPLICATIONS 123
For water ρ = 1 [gm/cm3 ]. For earth, r = 6 × 108 [cm], the ratio is minuscule:
10−13 . However, for very small grains, of radius less than 6×10−5 [cm] 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 universe3 .
Exercise 9.6 (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 9.8: A planet encircling a star and the tail of dust it sprays (tail)
Figure 9.9: Halbach array gives a large magnetic field above the array and small
one below it.
Chapter 10
Electrostatics and
magnetostatics
Now, that we have Maxwell equations, it remains to look at solutions for inter-
esting physical problems. We start with time independent problems.
∇ · E = 4πρ, ∇ × E + |{z}
Ḃ = 0
=0
and dot is a derivative with respect to x0 = ct. In the static case Ḃ = 0 and
the equations reduce to
∇ · E = 4πρ, ∇×E=0
E = −∇φ
∆φ = −4πρ (10.1)
which is a partial differential equation for the potential (if the dimension n ≥ 2).
Remark 10.1 (Time dependent ρ). From a formal mathematical point of view,
one may also consider Poisson’s equation with ρ that is time dependent (e.g. a
moving point charge). However, time dependent ρ would normally entail Ḃ 6= 0.
1 And the assumption that space is Minkowski and the fields vanish at infinity
125
126 CHAPTER 10. ELECTROSTATICS AND MAGNETOSTATICS
Show that ∇ · E = 0 implies T r e(2) = 0. In particular the matrix e(2) can not
have all eigenvalues of one sign.
E = −d · E + M g · x
How can we turn this setting to one where the equilibrium is stable? Clearly,
this can only occur if we let d be a function of position. Suppose that, for some
reason, the dipole wants to orient itself opposite to the local field,
d(x) = −d Ê(x)
This language also works for any real vector in 2-D, which can be represented
by a single complex number
E = E · x̂ + i E · ŷ
2∂E = ∇ · E − i∇ × E (10.9)
Hence, for any real vector field, ∇ · E is the real part of 2∂E.
128 CHAPTER 10. ELECTROSTATICS AND MAGNETOSTATICS
Write the potential inside the disc as a sum of a holomorphic and anti-holomorphic
functions
∞
X ∞
X
φ = f + ḡ + a0 , f (z) = fn z n , g(z) = gn z n
n=1 n=1
Evidently, for n ≥ 1
fn = an , ḡn = a−n
f and g are analytic in the unit disc since the corresponding series are absolutely
convergent.
We can now use Riemann mapping theorem to generalize this result for
the unit disc to any open, simply connected domain D in the plane. Riemann
mapping theorem states that if D is an open, simply connected domain in R2
then there is a holomorphic (and invertible) function ζ = ϕ(z) that maps D to
the unit disc |ζ| < 1.
n1 nd
This is explained in the figure. The Harmonic polynomial are the kernel of the
Laplacian3
∆ : Vn,d → Vn−2,d (10.12)
3V n+d−3
n−2,d = 0 for n = 0, 1 since d−1
= 0 for n < 2.
130 CHAPTER 10. ELECTROSTATICS AND MAGNETOSTATICS
∂2
L2 = −
(∂θ)2
and the “spherical harmonics” are
dim Hn,3 = 2n + 1
Spectrum(L2 ) = n(n − 1)
d2
∆= + 4∂ ∂¯ (10.17)
dz 2
It follows that (x + iy)n and (x − iy)n are Harmonic. In particular
n
x + iy
n
z =r n
=⇒ Yn,n (θ, φ) = eiφ sinn θ (10.18)
r
∆ r2−d−n Yn = r−d−n (2 − d − n) 2) − L2 Yn
2−d −n+ −
(d
= r−d−n n(n − 2 + d) − L2 Yn
=0 (10.23)
1 x x · Qx
H2 (x) = x · Qx ⇐⇒ H2 2 = (10.25)
r r |x|5
φ + Harmonic (10.27)
G = ∆−1
∆G(x − y) = δ(x − y)
sd is the area of the d-dimensional unit sphere. This clearly satisfies Gauss law
for unit source Z Z
E · dSd = ∇ · E} dVd = 1
| {z (10.34)
R
=δ(x)
for any sphere of radius R. The source is a delta function. For d > 2 we have
1 d−2
∇ = − d−1 r̂
rd−2 r
it follows that the Green function is
1 1 r r
Gd (r) = − , G2 (r) = log , G1 (r) = (10.35)
(d − 2)sd rd−2 2π a 2
a > 0 is an an arbitrary (length) scale factor.
In 1 and 2 dimensions the Green functions diverges at infinity and in d ≥ 2
it diverges at the origin.
Sanity check:
2π 4π
v3 =v1 =
3 3
Here is an amusing observation about spheres: Evidently
(2π)d
v2d+1 = v1
(2d + 1)!!
vd → 0 super-exponentially.
Exercise 10.9. Compute sd using the Gaussian integral
Z Z ∞
−x2 2
d
d xe = sd rd−1 dre−r
0
(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
Z
1
= dS · r̂ φ(x)
sd Rd−1 |x|=R
and Eq. 10.35 for G(R) has been used. This is precisely the average of φ over
the sphere of radius R.
10.8. STATIONARY MAGNETIC FIELDS 135
Ampere’s law then implies ∇ · J = 0 (and then also ρ̇ = 0). To solve Ampere’s
differential equation we are free to use any gauge we please. In particular, in
the Coulomb gauge:
B = ∇ × A, ∇ · A = 0
Using the identity
∇ × (∇ × A) = −∆A + ∇(∇ · A)
we can write Amper’s equation as Poisson’s equations for (the vector valued) A
4π
∆A = − J
c
Remark 10.10 (Consistency). The equation is consistent with the gauge con-
dition ∇ · A = 0 since ∇ · J = 0.
B(x) = ∇ × A(x)
Z
1 1
= ∇x ×J(y) dy
c |x − y|
| {z }
Coulomb
J(y) × (x − y)
Z
1
= dy
c |x − y|3
ẑ × x
B = 2I
|x|2
A = a × x + (b · x)c
Show that
B = 2a + b × c, ∇·A=c·b
and θ(x) = 1 for x > 0 and 0 otherwise is the standard step function. Consider
the limit a → 0 and I → ∞ so that that Ia2 is fixed.
θ(a2 − x2 − y 2 )
lim δ(z) = δ(x)
a→0 πa2
π 2 Ia2
m= ẑ
c
To find B we could plug the source into Biot-Savart. However, this is not much
10.8. STATIONARY MAGNETIC FIELDS 137
Exercise 10.16 (Vanishing flux). Show that the total flux through any such
plane at distance ε from the origin vanishes.
Figure 10.3: An infinitesimal loop around the positive z-axis carries flux of 4πem
while the same loop around the negative z-axis carries zero flux.
This can be given two interpretation: As the total flux through the sphere
minus a tiny hole near the north pole, or as minus the total flux through the
hole. 4πem is indeed the total flux of a monopole. Hence the Dirac string can
be thought of as the solenoid the feeds the flux of the monople. Now, how shall
we think of this Dirac string. On the one hand it is a gauge dependent property.
We could have chosen our coordinate system so that the string would go from
the origin to infinity along any direction we wish. On the other hand the Dirac
string carries magnetic field with flux 4πem . Magnetic fields are physical!
The problem would go away if we could use quantum mechanics to argue that
there are certain invisible magnetic fluxes. Indeed, this is what the Aharonov-
Bohm effect tells us: In an interference experiment involving a flux tube, one
can not distinguish no flux from a flux with integer number of quantum flux
quanta. This says that the Dirac string becomes invisible if
hc
4πem = Integer × Φ0 , Φ0 = (10.43)
e
140 CHAPTER 10. ELECTROSTATICS AND MAGNETOSTATICS
Figure 10.4: The interefrence pattern from a flux tube is periodic in the flux
with period Φ0 = hc
e
We can try to glue the two half space by taking U = e−inφ . This is possible
provided
~c ~c
AN φ − AS φ = i ∂j log U ∂φ φ = n (10.50)
e e
Comparing with Eqs. 10.45, 10.46 we see that
~c
AN φ − AS φ = n = 2em (10.51)
e
We recovered Eq. 10.44. This gives Dirac quantization rule.
∇ · E = 4πρ, ∇×E=0
and B is sourceless
∇ · B = 0, ∇ × B = 4πω
As we have seen, the equations for E and B are solved by the same technique.
Figure 10.5: Left: An irrotational field with a source at the origin. A sourceless
field with vorticity along the z-axis
If n 6= 0 the loops link. The converse is, however, not always true.
Figure 10.6: Linking circles. The linking number, here 1, can be computed from
Eq. 10.52
Chapter 11
Electromagnetic waves
Aµ 7→ Aµ + ∂µ Λ (11.4)
1I shall show later that this is always possible.
143
144 CHAPTER 11. ELECTROMAGNETIC WAVES
This leads to the wave equation with a source them that is proportional to
∇ × J:
4π
B = − ∇ × J
c
In particular, when ∇ × J = 0 we get the free wave equation for the magnetic
field.
Similarly for the electric field we have from Ampere’s law
Ë + c∇ × Ḃ = 4π J̇
Substituting Faraday’s law and making use of Gauss law gives
Ë = −c2 ∇ × (∇ × E) + 4π J̇ = c2 ∆E − c2 ∇(∇ · E) + 4π J̇
= c2 ∆E − 4π c2 ∇ρ − J̇
The electric field satisfies the wave equation with a source term proportional to
∇ρ − J̇/c2 !
J̇
E = 4π ∇ρ − 2
c
11.3. PLANE WAVES 145
In the absence of source terms, the electric field satisfies the free wave equation.
This was one of Maxwell’s great discoveries, namely, that even in the absence
of sources, the equations admit interesting wave-like solutions. This allowed
him to interpret light coming from the stars and propagating in vacuum, as
electromagnetic waves and eventually lead to the discovery of the radio.
In the Coulomb gauge, the amplitudes are orthogonal (in Euclidean space) to
the direction of propagation.
146 CHAPTER 11. ELECTROMAGNETIC WAVES
Figure 11.1: The triad of E, B, k for a plane wave. The wave propagates in
the k̂ = Ê × B̂ direction
11.3.2 Doppler
Since k · x is a Lorentz scalar
k · x = k 0 · x0
kµ0 = Λµ ν kν , a0µ = Λµ ν aν
Longitudinal Doppler
Consider a plane wave propagating in the z-direction, Eq. (11.9). 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
Transverse Doppler
Consider, as before, a wave propagating in the z-direction, but a boost in the
x-direction so that
Λ0 1 = Λ1 0 = − sinh φ, Λ0 0 = Λ1 1 = cosh φ, Λ2 2 = Λ3 3 = 1 (11.13)
The wave vector kµ = ω(1, 0, 0, 1) is transformed to a light like wave vector
kµ0 = ω(cosh φ, − sinh φ, 0, 1) in the x − z plane. The new frequency is
ω 0 = ω cosh φ = ω γ
This is quadratic in the velocities for small speeds.
11.4 Polarization
11.4.1 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 plane wave propagating is the real
part of
E0 eiφ , φ = k · x − ωt (11.14)
The amplitude, E0 , is a complex vector in the plane perpendicular to k since
k · E0 = 0. E0 therefore has 4 real amplitudes. What is their physical interpre-
tation? Suppose we scale
E0 7→ λE0 , λ = |λ|eiγ (11.15)
One would then say that the amplitude has been scaled by |λ| and the phase
shifted by γ. We have now identified 2 of the 4 parameters in E0 : An amplitude
and a phase. It remains to identify the remaining two parameters hidden in E0 .
We are now in a situation that is reminiscent of quantum mechanics: The wave
function is a complex vector with an equivalence relation by an overall complex
number.
148 CHAPTER 11. ELECTROMAGNETIC WAVES
Figure 11.2: 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.
11.4.2 Polarization
Let x̂ and ŷ denote orthogonal unit vectors in the plane perpendicular to k. Let
me introduce basis vectors, with Dirac ket notation
x̂ ± iŷ
|z± i = √ (11.16)
2
The states are normalized and orhtogonal
which does not care about the overall normalization and phase. Since T rρ = 1
while det ρ = 0 the two eigenvalues of ρ are 1 and 0: ρ is a projection
ρ2 = ρ (11.20)
11.4. POLARIZATION 149
Figure 11.3: The Poincare sphere associates with every point on the sphere
a polarization. The north and south poles represent right and left circularly
polarized light and the equator with linearly polarized light.
Exercise 11.3 (South pole). Show that the south pole represent left circular
polarization.
In the case that all these sources emit plane waves all sharing the same direc-
tion k̂ the polarization of the mixture is naturally defined as the mixture of
polarizations X
ρ= p j ρj (11.26)
11.5. THE WAVE EQUATION 151
It is still true that T rρ = 1. But now hρi need not be a projection since the
averages of a unit vectors is shorter, in general, than a unit vector: | hsi | ≤ 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.
Remark 11.5 (Combing a tennis ball). One amusing, essentially topological,
property of the transverse nature of electromagnetic waves is that it is not pos-
sible to have a fully spherically symmetric electromagnetic wave. The point is
that a spherical wave, with k pointing radially, has E tangent to the sphere. It
is a basic fact in topology that any vector field on the sphere must vanish at (at
least) two points. The field can not be “the same” everywhere.
φ = 4∂uv φ = 0
This allows for reconstructing f and g from the initial data by integration. One
can verify that, the solution in terms of the initial data, is
Z x+ct
1 1
φ(x, t) = (φ0 (x + ct) + φ0 (x − ct)) + φ̇0 (y)dy (11.28)
2 2c x−ct
ct
Figure 11.4: If the initial data, φ0 and φ̇0 are localized in the red interval, the
solution at later times lives in the forward light cone of the initial data. This is
called the domain of influence of the initial (red) data. This statement holds in
any inertial frame.
¯
(∂uv + ∂∂)φ(u, v, z, z̄) = 0, z = x1 + ix2 , z̄ = x1 − ix2
This ansatz solves the wave equation provided the function c(u) satisfies
The solution of this equation has an integration constant giving for c(u)
1
c(u) =
`2 − iλu
` is the minimal waist of the beam at u = 0, and the waist disperses with the
law
(`4 + λ2 u2 )1/4
where Gd+1 is the Green function in space-time, i.e. a solution of the equation
Y
Gd+1 (x) = δ d+1 (x), δ d+1 (x) = δ(xµ ) (11.30)
µ
Eq. 11.30 alone does not fix a unique is not unique Gd+1 , as we can add to
Gd+1 any solution of the free wave equation. Any solution Gd+1 satisfies the
free wave in the past t < 0. We can therefore subtract from Gd+1 the free wave
that agrees with the free wave in the past to have a solution with Gd+1 in the
past. We can repeat the procedure for different rest frames and then we arrive
at the conclusion that we can impose on Gd+1 the condition that it vanishes
outside the forward light cone. It turns out that Gd+1 has different qualitative
properties depending on whether d is even or odd. Thus although we mostly
care about d = 3, we shall consider the general case.
It follows that Z
dd x ∂t G
is a conserved quantity for t > 0 (and trivially so for t < 0). It has a jump at
t = 0: Z (
1 t>0
dd x ∂t G =
0 t<0
154 CHAPTER 11. ELECTROMAGNETIC WAVES
ct
Figure 11.6: Gd+1 = 0 outside the forward light cone. In particular Gd+1 = 0
for xµ xµ > 0.
This is seen by integrating Eq. 11.30 on the space-time slice − < t < .
s = −xµ xµ
ct
dGd (s)
= πGd+2 (s) (11.32)
ds
So if you know the green function in 1 dimension, you can find it for all odd
dimensions by differentiation, and if you know it for 2 dimensions you get it for
all even dimensions.
d2 G1
= c2 δ(ct) (11.33)
dt2
Integrating once, taking into account the light-cone condition, gives
dG1
= cθ(t) (11.34)
dt
156 CHAPTER 11. ELECTROMAGNETIC WAVES
1
G2+1 (s) = √ (11.37)
2π s
In even space dimensions Gd+1 fill the forward light cone. The Green function
does not quite satisfy the Huygens principle as the wave does not live on the
boundary of the light-cone, but rather fills it up. This is a feature of all even
spatial dimensions.
(Note that once again the solution fills the light cone.) From the recursion
relation we now find for G3+1
(
1
δ(s) in the forward light cone
G3+1 (s) = 2π (11.41)
0 otherwise
Note that G3+1 lives on the surface of the light cone. This is the Huygens
principle. This continues to hold in all higher odd dimensions.
δ(|x| − ct)
G3+1
> (x) = θ(t) , x = (ct, x), (11.42)
|x|
∇ · A = 0, ∆Φ = −4πρ (11.43)
Φ is the solution of Poisson’s equation the gauge field A satisfies the wave equa-
tion with a source term:
4π
− ∆A − Ä = J − ∇Φ̇ (11.44)
c
Suppose ∇ · A 6= 0. Let Λ be a solution of the Poisson’s equation
∆Λ = ∇ · A
∇ · A0 = ∇ · A − ∆Λ = 0
Φ0 It is determined by
E = −∇Φ0 − Ȧ0
Taking the divergence of this we see that Φ0 as a solution of Poisson’s equation:
The wave equation for A0 follows from Maxwell and the Coulomb gauge condi-
tion:
4π
J = ∇ × B − Ė
c
= ∇ × (∇ × A) + ∇Φ̇
= −∆A + ∇(∇ · A) + ∇Φ̇
= −∆A + ∇Φ̇ (11.46)
Remark 11.10 (Causality). The Coulomb gauge is a-causal: The scalar poten-
tial φ is fixed by the instantaneous charge distribution. You move a charge here
and the scalar potential φ changes immediately everywhere. The fact that the
scalar potential changes faster than light can not be used to transfer information
faster than light because the fields are still causal, and only fields are measurable.
Remark 11.11 (Free space). When ρ = 0 we may take φ = 0 together with
∇ · A = 0.
158 CHAPTER 11. ELECTROMAGNETIC WAVES
11.8 Appendices
11.8.1 Cosmic rays: GZK limit
The cosmic microwave background (CMB) provides a shield that screens ultra
high energy cosmic rays: A high energy charge proton can collide with a photon
to produce a neutral pion converting much of the high kinetic energy of the
proton to the pion mass. This makes the CMB a screen for high energy cosmic
rays. The GZK limit says that protons with energies above 5 × 1013 M eV are
screened by the 3◦ K thermal photons of the CMB.
Let us compute the threshold for particle (pion) production. The total
energy-momentum of a proton with rapidity φ and counter-propagating pho-
ton in the plane is
and the equality on the right expresses the fact that the two particles are at
rest. This gives the threshold for pion production as
Exercise 11.12. Can you figure out why the estimate is too big?
The forward light cone is associated with out going (retarded) waves. Similarly
Z
1 dk
φ< (t, x) = 2
φ̃(−|k|, k) e−i(|k|t+k·x) (11.50)
(2π) 2|k|
k · k = k02 (11.52)
The smallest wave length that such a wave can accommodate is 2π/k0 : The
frequency limis the spatial resolution.
The noteworthy fact about this waves is that the wave number k in the z-
direction can be much larger than the wave number associated with the fre-
quency ω . Near the boundary x = 0 one finds waves with short wave lengths:
k k0 = ω/c.
E = E0 e−κx eikz
κ(E0 )1 + ik(E0 )3 = 0
Ė + ∇ × B = 0, Ḣ − ∇ × D = 0
∇ · D = 0, ∇·B=0 (11.53)
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
11.8.7 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 11.15. Can you give an (ergonomic) argument why spectators would
prefer circular to linear polarization?
Exercise 11.16. 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
Exercise 11.17. 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
162 CHAPTER 11. ELECTROMAGNETIC WAVES
PV σ 3 = σ 3 PH
Radiation
1 δ(ct − r)
G> (x) = θ(x0 )δ s = , s = x · x, r = |x| ≥ 0 (12.1)
2π 4πr
The θ function guarantees causality: The past influences the present. The
delta function says that in 3+1 dimensions signals propagate on the light-cone.
Huygens principle (in the strong form) holds.
By the homogeneity of Minkowski space, for source term located at the
space-time point y
G> (x − y) = δ(x − y)
By the linearity of the wave equation the retarded solution of the wave equation,
φ, generated by an arbitrary1 source ρ
φ = ρ(x) (12.2)
is Z
φρ (x) = d4 y G> (x − y)ρ(y) (12.3)
Olber’s paradox: If you assume constant density of stars, and that intensity of radiation falls
like r−2 the night sky shoudl be as bright as the sun.
163
164 CHAPTER 12. RADIATION
moving on a world line z = ct, z(t) , −∞ < t < ∞. The motion is assumed to
be that of a real particle so the 4-velocity is time-like. We take the source to be
4π (3)
ρ(x, t) = δ x − z(t) (12.4)
c
x
x0
R
y
y0
Figure 12.1: The blue line is the world line of the point source. The wave is
observed at the black dot x. The backward light cone from x intersects the
(blue) world line at the black dot y. Since the velocity is time like the point of
intersection is unique. R is the 4 vector x − z(y 0 ).
Since the orbit z(z 0 ) is time-like, a single point contributes: a single particle
has a single image2 . To compute the remaining time-integral use
Z
1
δ s(y) dy = 0 , s = R · R, s(y0 ) = 0. (12.5)
|s (y0 )|
2 Mirrors and lenses can create multiple images, of course.
12.2. MAXWELL EQUATION IN THE LORENZ GAUGE 165
Clearly, the time derivative of s is related to the velocity of the source, and it
is natural to express it in terms of the 4-velocity
ds dRµ dz µ
= 2Rµ = −2R µ
dy 0 dy 0 dy 0
µ
dz dτ Rµ uµ
= −2Rµ 0
= −2
dτ dy cγ
R·u
= −2
cγ
The wave φ(x) at the observing point x is therefore given by the deceptively
simple formula
γ(y 0 )
φ(x) = , R = x − y, R·R=0 (12.6)
|R · u(y 0 )|
t u
R
y
x
Figure 12.2: The blue 4-velocity is time-like and the 4-vector R is light like.
Their Minklowsky scalar product is negative and can be close to zero.
The formula is simple but implicit. The right hand side is not an explicit
function of the argument x: You need to know γ(y 0 ), u(y 0 ) and R = x − y and
in particular, the earlier time y 0 (see Fig. 12.1.1). This time is determined as
the solution of the equation
2
(x0 − y 0 )2 = x − z(y 0 )
Comparing with Eqs. (12.4,12.6) for scalar waves we see that conveniently γ
disappears, and the retarded potentials are:
uν uν (y 0 )
Aν (x) = e = −e (12.11)
|R · u| R · u(y 0 )
The absolute value was removed by taking into account that R is forward light-
like and u forward time-like so R · u < 0.
The result admits the following (a-posteriori) 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: One interesting R · u and two uninteresting u · u = −c2 and
R · R = 0. This, plus dimension analysis and the limit case of a charge at rest
determines Eq. (12.11).
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
12.3. LIENARD-WIECHERT: RETARDED POTENTIALS 167
0 = 12 ∂µ (R · R) = Rα ∂µ (xα − z α ) = Rµ − R · u (∂µ τ )
We have verified that the solution, Eq. 12.11, indeed satisfies the Lorenz con-
dition.
168 CHAPTER 12. RADIATION
Figure 12.3: 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.
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) (12.17)
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 = 2
− c2 3
− R · u̇ (12.18)
R·u R·u (R · u) (R · u) (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 (12.19)
(R · u) (R · u) (R · u)3
| {z } | {z }
radiation ”Coulomb”
Since u = O(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 their velocity dependence is of order O(c−2 ). They decay
like R−1 . These are the radiating terms.
12.4.1 Interpretation
The Lienard-Wiechert 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.
• u̇ the 4-acceleration.
From these we can make three interesting scalars
R · u, R · u̇, u̇ ·
u̇ (12.20)
R · R = 0, u · u = −c2 , u · u̇ = 0 (12.21)
This is because the potential did not depend on the acceleration at all. As a
consequence, the scalar u̇ · u̇ should not appear and the scalar R · u̇ can only
appear in the numerator.
We can now reconstruct all the three terms in F 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 is the last term in Eq. (12.19). 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).
One possible term proportional to the acceleration is u̇ is
R[µ u̇ν]
e
(R · u)2
which gives the first term up tod numerical factor. The middle term is obtained
similarly.
e R[i u̇j]
= −
c2 |x|2
e (a(0) × x)k
= (12.23)
c2 |x|2
12.5. ACCELERATING PARTICLE IN ITS REST FRAME 171
where a, the 3-vector of acceleration u̇µ = (0, a), is orthogonal to u = (c, 0).
It follows that, 3 vector of magnetic field on the light-cone emanating from
the origin, (|x|, x) is
e a(0) × x
B(|x|, x) = 2 (12.24)
c |x|2
The main conclusions we draw from this are
1. The field decays like the inverse distance from the source.
2. The field is perpendicular to both the line of sight and the acceleration
vector.
Since
(x · x)a − (x · a) x = − x × (x × a)
The longitudinal part is Coulomb and the transversal part is radiation. E and
B are mutually orthogonal. The two Lorentz scalars are
e2
E · B = 0, E2 − B2 = −
|x|4
c e2
E×B= x̂
4π 4πc3 |x|2
172 CHAPTER 12. RADIATION
In the far field |z| is small compared to |x| and can approximate
(We have used the fact that R is light-like.) If, in addition, the charge is non-
relativistic
uµ ≈ (c, 0) (12.28)
and we obtain
e R[i u̇j] e xi aj − xj ai
εijk B k (x) ≈ − ≈− 2 , r = |x| (12.29)
c2 r2 c r2
We have essentially recovered the results Eq. 12.24 in the far-field also for
particles that need not be stationary and at the origin
e a(y 0 ) × x
B(x) ≈ (12.30)
c2 |x|2
Note that the result now implicitly depends on the retarded time y 0 .
at the time time y 0 when the radiation has been emitted. This time is deter-
mined as the solution of the equation
x0 − y 0 = |x − z(y 0 )|
where z(y 0 ), the obit of the source (charge), is given. This is an implicit equation
for y 0 . A geometric solution is given in Fig. 12.1.1. In general, one can not hope
to find and analytic solution: A-priori z(y 0 ) could be an arbitrarily complicated
function that can not be inverted explicitly. The best we can hope for is to find
approximate solutions in simple cases when there is a small or a large parameter
that we can use in making successive approximations.
12.6. RETARDATION FROM A DISTANT SOURCE 173
R = (x − y 0 , x − z(y 0 )) (12.32)
is light-like. So, it is sufficient to have a good estimate of its spacial part. Now,
if the source is distant from the point of observation, and if it is confined to a
relatively small ball in space,
When |x| is large, we need to approximate |x − z| not just to order |x|, but also
to order |x|0 = O(1), but we may neglect |x|−1
2 !
2 2 2 2 x·z `
|x − z| = |x| + |z| − 2x · z = |x| 1−2 2 +O (12.34)
|x| |x|
√
Approximating 1 + 2ε ≈ 1 + ε gives
x·z
|x − z| ≈ |x| 1 − (12.35)
|x|2
The implicit equation for the retardation y 0 for a distant source simplifies to:
This is still an implicit equation for y 0 , so we shall need to make some further
approximations if we want to find explicit approximate expression for y 0 .
ω δt 1 ⇐⇒ ω δy 0 c (12.38)
For example, in the case that the source is harmonic, this means that we need
to know the phase of the source y 0 ω/c to an accuracy that is much better than
2π. We increasingly better accuracy as ω gets large.
174 CHAPTER 12. RADIATION
y10 = x0 − |x|
y20 = x0 − |x| + x̂ · z(y10 ) = x0 − |x| + x̂ · z(x0 − |x|) (12.39)
...
Let us see how good the successive approximations are. We can estimate the
error in y10 by comparing with the defining equation for y 0 :
ct x
y20
x0 − |x|
x
−`/2 `/2
Figure 12.4: The blue line gives the orbit of the charge. The lowest oredr dipole
approximation is x0 −|x|. A better approximation is y20 which takes into account
the position of the charge at x0 − |x|. One could get better approximations by
also taking into account the velocity of the charge.
The most interesting thing to observe is that this outgoing spherical wave is a
consequence of retardation.
Exercise 12.3. Consider the radiation from two opposite charges ±e executing
harmonic motion with opposite amplitudes ±z0 . In this case one needs to con-
sider the retardation beyond the leading order in the dipole approximation, i.e.
one needs to keep the last term in Eq. 12.45.
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) = (12.49)
c2 |x|
,
12.7 Power
The power emitted by dipole can be computed from Poynting
E×B
P=c (12.50)
4π
Both E and B lie in the plane perpendicular to the line of sight so P is parallel
to r̂.
Linear Dipole: Suppose that a = aẑ. The magnitude of P in the direction
of the spherical angle θ relative to the z-axis is
E2 e2 a2 sin2 θ
P (θ) = c = (12.51)
4π 4πc3 r2
The power through a spherical shell of radius r is then
Z
PT = 2πr2 dθ sin θ P (θ) (12.52)
Evidently
Z Z
2
4
dθ sin θ sin θ = − d(cos θ) (1 − cos2 θ) = 2 1 − 31 = (12.53)
3
12.7. POWER 177
B
E
a
Figure 12.5: Dipole oriented along the z-axis and the associated fields. φ is the
angle between the z-axis and the blue arrrow.
Figure 12.6: Polar plot of the power radiated by a dipole antenna as function of
the angle, Eq. 12.51. The maximal power is radiated in the plane perpendicular
to the dipole.
The radiation from a dipole antenna is not isotropic: It does not radiate at
all in the directions of the dipole. The good news is that you can play with the
geometry of the antenna to make it directional so you do not waste power in
directions that you do not want.
Remark 12.4. You can not make an isotropic antenna. No matter how com-
plicated an antenna you make the Poynting vector must vanish it at least in two
directions. This is a consequence of topology: The vector B is tangent to the
178 CHAPTER 12. RADIATION
sphere. It is a fact that every vector field tangent to the sphere must vanish at
two points, at least. (Or vanish quadratically at one point.) This is sometimes
expressed as you can not comb a tennis ball. Hence P must vanish at two points
at least.
Ė 1 d(E −3 )
k=− = 3 (12.56)
E4 dt
Hence
E(t) = E0 (1 − γt)−1/3 , γ = −3E03 k > 0
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 2
2 e
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
12.8. CLASSICAL INSTABILITY OF ATOMS 179
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
that lead to quantum mechanics.
E
t
Figure 12.7: Blowup at finite time: The energy of a charged particle encircling
the nucleus goes to −∞ in finite time.
180 CHAPTER 12. RADIATION
Chapter 13
Radiation reaction
181
182 CHAPTER 13. RADIATION REACTION
of the strings.
You may wonder if the problem with infinities in not cured by quantum me-
chanics. After all, Heisenberg uncertainty relation does not allow to fully localize
a particle to a point. So, perhaps, a quantum theory of charges coupled with a
classical electromagnetic field is a consistent theory? The trouble is that it is not
possible to consistently couple a quantum theory for the charges with a classical
theory for the electromagnetic fields. This is most easily seen in the Heisen-
berg picture: In a quantum theory the observables are non-commuting matrices
while in a classical theory they are commuting functions. If the two theories are
coupled the classical theory will be “infected” by the non-commutativity of the
quantum theory.
What about QED: A relativistic quantum theory of both charges and fields?
It is now widely believed that this theory although practically useful, is in
principle, ill-defined. It is a “phenomenological” mirror of a more involved,
hopefully consistent, QCD.
13.1.2 Mathematics
Newton equations are non-linear ODE’s coupled to Maxwell equations which are
linear PDE’s. As the the charges are point-like the source terms are singular, and
so are the fields. The standard theorems in the theory of differential equations
do not cover singular PDE coupled to non-linear ODE. This is worse than the
case of Navier-Stokes equations, which is still open.
∇ · A = 0, ∆Φ = −4πρ (13.3)
A 7→ A + ∇ × C (13.5)
Maxwell’s equations for slow particles, to leading order in v/c, reduce to Pois-
son’s equation for Φ. The electromagnetic field is not a dynamical field anymore:
It is a slave of the motion of the charges ρ. Since we know how Poisson’s equa-
tion can be explicitly solved given the position of the cahrges, the dynamics is
only in the motion of the charges. This gives Eq. 13.2.
a2 = (a ·˙ v) − ȧ · v = −ȧ · v
2 e2 2 2 e2
P = 3
a = − 3 ȧ · v = −FAL · v
3c 3c
If the particle is moving at constant speed, an opposite (non-electromagnetic)
force to FAL must be applied to feed back the energy lost to radiation. This
184 CHAPTER 13. RADIATION REACTION
v ȧ
Exercise 13.1 (Covariant form of Abraham Lorentz force). Using the fact that
Newton law
maµ = fµ
is consistent with u · u = −c2 provided f · u = 0. Using a · a + u · ȧ = 0, we see
that the covariant form of Abaraham-Lorentz force is
2e2 a·a
maµ = 3 ȧµ − 2 uµ (13.8)
3c c
e2
mc2 = (13.10)
r0
Then τ is the time it takes light to cross the classical radius:
r0
τ= (13.11)
c
For the electron
τ = O(10−23 ) [s] (13.12)
The ratio in Eq. 13.9 is a ridiculously small number unless the particle has
violent jerk, with characteristic frequencies ω 1023 Hz.
13.3.2 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:
1 2 e2
a= f (x) + τ ȧ, τ = (13.13)
m 3 mc3
Viewing τ as a smallest time scale in the problem we treat it as a perturbation
and solve the equation of motion by iteration. To leading order
1
a1 = f (x1 ) (13.14)
m
To second order
1
a2 = f (x2 ) + τ ȧ1
m
1
= f (x2 ) + τ ḟ (x1 )
m
1
= (f (x2 ) + τ (v1 · ∇)f (x1 )) (13.15)
m
which we approximate by an equation with a weak friction term
z 2 ≈ 1 ± iτ ω (13.19)
Exercise 13.2. Write the equations of motion for the Kepler problem with
radiation reaction friction term.
To address this consider, following Amos Ori, a the forces that diffeent parts
of an extended body apply on each other. In the limit that the size of the object,
ε → 0, we shall recovers Abaraham Lorentz with the extra bonus that we get
an interpretation of the mass in terms of the energy of the field.
We shall derive the self-force force on a dumbbell made of two point charges
separated by a rod of length ε.
We first compute the forces that the dumbbell applies on itself when it moves
in a prescribed way. The world-lines of the dumbbell are shown in the figure
(red) and the light-cone is drawn blue. We shall see that as a consequence of
retardation, Newton third law is violated, and there is a net force acting on an
extended body.
13.3. RADIATION REACTION: THE ABRAHAM-LORENTZ FORCE 187
hhhh
hh
x
The dumbbell moves along the x-axis and is aligned with the y-axis. So the
world line of the two charges is
The two charges communicate when x± (t + τ ) − x∓ (t) is light like. The time
delay τ = O(ε) 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 forces at time τ and express them in terms of the
acceleration and jerk at the same time τ . For simplicity we take the jerk to be
constant. Then
a(τ ) = a = a(0) + ȧτ
From this it follows that
The retardationτ and the dumbbell size 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 . Hence, to leading order
cτ ≈ ε (13.23)
The retarded field at time τ is determined by the velocity and acceleration of
the particle at time 0 when the signal was generated:
R[µ aν] R[µ uν] 2 R[µ uν]
Fµν = −e − R·a−c (13.24)
(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 from Eq. 13.20
Since this term appears in the denominator of F, the limit ε → 0 is singular and
has to be taken carefully. In particular, we need to keep terms in the numerator
to O(ε3 ).
The electric field in the direction of motion on one of the charges due to the
other at time τ is
!
R0 a1 R1 u0 2 R1 u0
−Ex = F01 = −e + R · a +c
(c2 τ )2 (−c2 τ )3 (−c2 τ )3
| {z } | {z }
a0 =0 u1 =0
e2
2
2e
mb + a= ȧ + Fex
2τ c3 3c3
We interpret the brackets as the effective mass which gets a contribution from
the electromagnetic energy in the field.
This is quite tantalizing, for it offers a new interpretation of the mass as
being generated by the field. The interpretation is not completely satisfactory
for as we send τ → 0 we need to take the bare mass large and negative in order
to get a finite effective mass.
Remark 13.3. In this computation we neglected the self-force of the point
charge on itself. One can get the self radiation reaction using the following
trick Amos Ori taught me. Let f (e) denote this self force. It is a quadratic
function of e. To find it use the fact the in the limit τ → 0 the total force on
the dumbbell gives an equation for f (e):
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.
2e2
a = τ ȧ , τ=
3mc3
It admit the solution
a(0) t/τ
a(t) = a(0)et/τ =⇒ v(t) = v(0) + e (13.28)
τ
A particle, initially at rest, self accelerate to large velocities. This can only be
avoided if you tune a(0) = 0 precisely. In practice, we only tune x(0) and v(0),
so how come we do not see these self-accelerations? Classical Electrodynamics
for point particles is fundamentally flawed. You could use this to argue for going
quantum.