Sie sind auf Seite 1von 217

Kostas Kokkotas

Field Theory

– Handouts from the course –

July 9, 2010
Preface
These notes are from the transparencies written for the course “Field Theory”.
The parts related to Electrodynamics were based mainly on the book by
• J.D. Jackson Classical Electrodynamics while other books like Glassical
Electrodynamics by W. Greiner
while we have also used
• The Classical Theory of Fields by L.D. Landau & E.M. Lifshitz
• Eric Poisson’s notes
• Electrodynamics by Fulvio Melia
• Classical Electrodynamics by Walter Greiner
For the parts related to General Theory of Relativity we have used the
books
• General Relativity: An Introduction for Physicists by M.P.Hobson, G. Ef-
stathiou & A.N. Lasenby
• A first course in General Relativity by B.F. Schutz.
• Gravity: an Introduction to Einstein’s General Relativity by James B. Har-
tle
Contents

1 Electrostatics-Magnetostatics (****) . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Green Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Laplace Equation in Spherical Coordinates . . . . . . . . . . . . . . . . . 5
1.4 Legendre Equation and Legendre Polynomials . . . . . . . . . . . . . . . 6
1.4.1 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Legendre Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6 Associated Legendre Functions and Spherical Harmonics . . . . . 10
1.6.1 Spherical Harmonics Ylm (θ, φ) . . . . . . . . . . . . . . . . . . . . . . 11
1.6.2 Addition Theorem for Spherical Harmonics . . . . . . . . . . . 12
1.7 Multipole Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.7.1 Monopole moment l = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.7.2 Dipole moment l = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.7.3 Quadrupole moment l = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.7.4 Multipole Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.8 Energy of a Charge Distribution in an External Field . . . . . . . . 18
1.8.1 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.9 Magnetostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.10 Biot & Savart Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.10.1 Diff. Equations for Magnetostatics & Ampere’s Law . . . 21
1.11 Vector Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2 Classical Field Theory: Maxwell Equations . . . . . . . . . . . . . . . . 25


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 Faraday’s Law of Induction (****) . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.1 Faraday’s law for a moving circuit . . . . . . . . . . . . . . . . . . . 26
2.3 Energy in the Magnetic Field (***) . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 Maxwell Equations (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.4.1 Vector and Scalar Potentials . . . . . . . . . . . . . . . . . . . . . . . . 33
2.4.2 Gauge Transformations : Lorenz Gauge . . . . . . . . . . . . . . 34
2.4.3 Gauge Transformations : Coulomb Gauge . . . . . . . . . . . . 35
VIII Contents

2.5 Green Functions for the Wave Equations (***) . . . . . . . . . . . . . . 35


2.5.1 Poynting’s Theorem : Conservation of Energy (****) . . . 38
2.5.2 Poynting’s Theorem : Conservation of Momentum (***) 39
2.6 Maxwell Stress Tensor (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.7 Conservation of Angular Momentum (**) . . . . . . . . . . . . . . . . . . . 41

3 Electromagnetic Waves (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43


3.1 Maxwell Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2 Plane Electromagnetic Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Linear and Circular Polarization of EM Waves . . . . . . . . . . . . . . 46
3.3.1 Circular Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3.2 Elliptically Polarized EM Waves . . . . . . . . . . . . . . . . . . . . . 48
3.4 Stokes Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5 Reflection & Refraction of EM Waves . . . . . . . . . . . . . . . . . . . . . . 50
3.5.1 E : Perpendicular to the plane of incidence . . . . . . . . . . . 53
3.5.2 E : Parallel to the plane of incidence . . . . . . . . . . . . . . . . . 54
3.5.3 Normal incidence i = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4 Simple Radiating Systems (****) . . . . . . . . . . . . . . . . . . . . . . . . . . 57


4.1 Fields and Radiation of a Localized Oscillating Source . . . . . . . 57
4.2 The Near Zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2.1 The Far Zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2.2 The Intermediate Zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2.3 Electric Monopole Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.3 Electric Dipole Fields and Radiation . . . . . . . . . . . . . . . . . . . . . . . 60
4.3.1 Electric Dipole : Power Radiated . . . . . . . . . . . . . . . . . . . . 62
4.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4.1 Example I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4.2 Example II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.5 Magnetic Dipole & Electric Quadrupole Fields . . . . . . . . . . . . . . 65
4.5.1 Magnetic Dipole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.5.2 Electric Quadrupole Fields . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.5.3 Example : Electric Quadrupole Fields . . . . . . . . . . . . . . . . 69
4.5.4 Example : Pulsar spin-down . . . . . . . . . . . . . . . . . . . . . . . . 69

5 Radiation by Moving Charges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73


5.1 Liénard - Wiechert Potentials (**) . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.1.1 Special Note about the shrinkage factor (1 − n · β) . . . . 76
5.2 Liénard - Wiechert potentials : radiation fields (****) . . . . . . . . 76
5.2.1 Some observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.3 Power radiated by an accelerated charge (****) . . . . . . . . . . . . . 78
5.3.1 Larmor Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.3.2 Relativistic Extension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.4 Applications (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Contents IX

5.5 Angular Distribution of Radiation Emitted by an Accelerated


Charge (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.6 Angular distribution of radiation from a charge in circular
motion (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.7 Radiation from a charge in arbitrary motion (***) . . . . . . . . . . . 86
5.8 Distribution in Frequency and Angle of Energy Radiated ...
(****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.9 What Is Synchrotron Light? (***) . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.9.1 Synchrotrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.10 Synchrotron Radiation (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.11 Thomson Scattering of Radiation (***) . . . . . . . . . . . . . . . . . . . . . 99

6 Special Theory of Relativity (****) . . . . . . . . . . . . . . . . . . . . . . . . 103


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.2 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
6.3 Invariance of Electric Charge; Covariance in Electrodynamics . 107
6.4 Dual Field-Strength Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.5 Transformation of Electromagnetic Fields . . . . . . . . . . . . . . . . . . . 110
6.6 Transformation of Electromagnetic Fields: Example . . . . . . . . . . 111

7 Dynamics of Relativistic Particles and EM Fields (**) . . . . . 115


7.1 Lagrangian Hamiltonian for a Relativistic Charged Particle . . . 115
7.1.1 Relativistic Lagrangian (Elementary) . . . . . . . . . . . . . . . . 116
7.1.2 Relativistic Lagrangian (Covariant Treatment) . . . . . . . . 118
7.1.3 Relativistic Hamiltonian (Covariant Treatment) . . . . . . . 119
7.2 Motion in a Uniform, Static Magnetic Field (****) . . . . . . . . . . 121
7.3 Motion in Combined, Uniform, Static E- and B- Field (***) . . 122
7.4 Lowest Order Relativistic Corrections to the Lagrangian... . . . . 124
7.5 Lagrangian for the Electromagnetic Field . . . . . . . . . . . . . . . . . . . 125
7.6 Proca Lagrangian; Photon Mass Effect . . . . . . . . . . . . . . . . . . . . . 126
7.7 Conservation Laws : Canonical Stress Tensor . . . . . . . . . . . . . . . . 127
7.8 Conservation Laws : Symmetric Stress Tensor . . . . . . . . . . . . . . . 129
7.9 Conservation Laws for EM fields interacting with Charged
Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

8 A Short Introduction to Tensor Analysis . . . . . . . . . . . . . . . . . . 133


8.1 Scalars and Vectors (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
8.1.1 Vector Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
8.1.2 Contravariant and Covariant Vectors . . . . . . . . . . . . . . . . . 134
8.2 Tensors: at last (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.2.1 Tensor algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.2.2 Tensors: Differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.2.3 Tensors: Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.2.4 Covariant Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.2.5 Parallel Transport of a vector . . . . . . . . . . . . . . . . . . . . . . . 137
X Contents

8.2.6 Curvature Tensor (**) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138


8.3 Geodesics (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
8.4 Metric Tensor (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
8.4.1 The Determinant of gµν (**) . . . . . . . . . . . . . . . . . . . . . . . 141
8.4.2 Christoffel Symbols (****) . . . . . . . . . . . . . . . . . . . . . . . . . 142
8.5 Geodesics in a Riemann Space (***) . . . . . . . . . . . . . . . . . . . . . . 143
8.5.1 Euler-Lagrange Eqns vs Geodesic Eqns (***) . . . . . . . . . 145
8.5.2 Tensors: Geodesics (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8.5.3 Null Geodesics (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
8.5.4 Geodesic Eqns & Affine Parameter (**) . . . . . . . . . . . . . 146
8.6 Riemann Tensor (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
8.6.1 The Ricci and Einstein Tensors (***) . . . . . . . . . . . . . . . . 147
8.6.2 Flat & Empty Spacetimes . . . . . . . . . . . . . . . . . . . . . . . . . . 148
8.7 Weyl Tensor (-) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
8.8 Tensors : An example for parallel transport (****) . . . . . . . . . . . 149

9 Physics on Curved Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151


9.1 The electromagnetic (EM) force of a moving charge (***) . . . . . 151
9.2 The 4-current density (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
9.3 The EM field equations (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
9.4 Electromagnetism in the Lorenz gauge (***) . . . . . . . . . . . . . . . . 153
9.5 Electric and Magnetic Fields in inertial frames (***) . . . . . . . . . 154
9.6 Electromagnetism in arbitrary coordinates (***) . . . . . . . . . . . . . 155
9.7 Equations of motion for charged particles (***) . . . . . . . . . . . . . 156
9.8 Energy-Momentum Tensor(***) . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
9.8.1 The Energy-Momentum tensor of a perfect fluid(**) . . . 159
9.8.2 The Energy-Momentum tensor of a “real fluid” (*) . . . . 159
9.9 Conservation of Energy and Momentum for a perfect fluid (**) 160
9.10 Linear Field Equations for Gravitation (**) . . . . . . . . . . . . . . . . . 161
9.11 Interaction of Gravitation and Matter (*) . . . . . . . . . . . . . . . . . . . 163
9.12 Local Cartesian coordinates(-) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
9.12.1 Local geodesic coordinates and Cartesian coordinates(-) 167
9.13 Geodesic deviation (-) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
9.14 Tidal forces in a curved spacetime (-) . . . . . . . . . . . . . . . . . . . . . . 169

10 Einstein’s Theory of Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171


10.1 Newtonian Gravity (***) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
10.2 Equivalence Principle (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
10.2.1 Equivalence Principle : Dicke’s Experiment . . . . . . . . . . . 173
10.3 Einstein’s Equations (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
10.3.1 Newtonian Limit (**) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
10.4 Solutions of Einstein’s Equations (*) . . . . . . . . . . . . . . . . . . . . . . . 175
10.5 Schwarzschild Solution (****) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
10.5.1 Schwarzschild Solution: Geodesics . . . . . . . . . . . . . . . . . . . 178
10.5.2 Radial motion of massive particles (**) . . . . . . . . . . . . . . . 179
Contents XI

10.5.3 Circular motion of massive particles (**) . . . . . . . . . . . . . 181


10.5.4 Stability of massive particle orbits (-) . . . . . . . . . . . . . . . . 182
10.5.5 Trajectories of photons (-) . . . . . . . . . . . . . . . . . . . . . . . . . . 185
10.5.6 Radial motion of photons(-) . . . . . . . . . . . . . . . . . . . . . . . . 186
10.5.7 Circular motion of photons (-) . . . . . . . . . . . . . . . . . . . . . . 186
10.5.8 Stability of photon orbits (-) . . . . . . . . . . . . . . . . . . . . . . . 186
10.6 The slow-rotation limit : Dragging of inertial frames (-) . . . . . . 187
10.7 The Classical Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
10.7.1 The Classical Tests: Perihelion Advance (***) . . . . . . . . . 188
10.7.2 The Classical Tests : Deflection of Light Rays (***) . . . . 190
10.7.3 The Classical Tests : Gravitational Redshift (***) . . . . . 192
10.7.4 The Classical Tests : Radar Delay (***) . . . . . . . . . . . . . . 193

11 Solutions of Einstein’s Equations & Black Holes . . . . . . . . . . . 197


11.1 Schwarzschild Solution: Black Holes (***) . . . . . . . . . . . . . . . . . . 197
11.1.1 Eddington-Filkenstein coordinates . . . . . . . . . . . . . . . . . . 197
11.1.2 Kruskal - Szekeres Coordinates : Maximal Extension (*) 199
11.1.3 Wormholes (*) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
11.2 Reissner-Nordstrøm Solution (1916-18) (*) . . . . . . . . . . . . . . . . . . 200
11.3 Kerr Solution (1963) (**) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
11.3.1 Infinite redshift surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
11.3.2 Penrose Process (1969) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
11.3.3 Black-Hole: Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
11.4 The Tolman-Oppenheimer-Volkov (TOV) Solution (-) . . . . . . . . 204
11.4.1 TOV : A uniform density star . . . . . . . . . . . . . . . . . . . . . . . 205

A Useful constants in geometrical units . . . . . . . . . . . . . . . . . . . . . . 207


1
Electrostatics-Magnetostatics (****)

1.1 Electrostatics
The behavior of an electrostatic field can be described by two differential
equations:
∇ · E = 4πρ (1.1)
(Gauss law) and
∇×E=0 (1.2)
the latter equation being equivalent to the statement that E is the gradient
of a scalar function, the scalar potential Φ:

E = −∇Φ (1.3)

Eqns (1.1) and (1.3) can be combined into one differential equation for a single
scalar function Φ(x):
∇2 Φ = −4πρ (1.4)
This equation is called Poisson equation.
In the regions of space where there is no charge density, the scalar potential
satisfies the Laplace equation:

∇2 Φ = 0 (1.5)

For a general distribution ρ(x0 ), the potential is expected to be the sum


over all increments of charge d3 x0 ρ(x0 ), i.e.,

ρ(x0 ) 3 0
Z
Φ(x) = d x (1.6)
|x − x0 |

This potential should satisfy Poisson’s equation. But does it? If we


operate with ∇2 on both sides of (1.6) we get (on x not on x0 )
2 1 Electrostatics-Magnetostatics (****)
Z  
2 0 3 0 2 1
∇ Φ(x) = ρ(x )d x ∇ (1.7)
|x − x0 |

But ∇2 (1/|x − x0 |) = 0 as long as x 6= x0 ! (Why ?)


The singular nature of ∇2 (1/|x − x0 |) = ∇2 (1/r) can be best expressed in
terms of the Dirac δ-function.
Since ∇2 (1/r) = 0 for r 6= 0 and its volume integral is −4π (Why?) we
can write  
1
∇ 2
= −4πδ(|x − x0 |) (1.8)
|x − x0 |

By definition, if the integration volume contains the point x = x0


Z
δ 3 (x − x0 )d3 x = 1

otherwise is zero. This way we recover Poisson’ s equation

∇2 Φ(x) = −4πρ(x0 )x0 =x (1.9)

Thus, we have not only shown that the potential from Coulomb’s law satisfies
Poisson’s eqn, but we have established (through the solution of Poisson’s eqn)
the important result that :
the potential from a distributed source is the superposition of the individual
potentials from the constituent parcels of charge.
We may consider situations in which ρ is comprised of N discrete charges
qi , positioned at x0i so that
N
X
ρ(x0 ) = qi δ 3 (x0 − x0i ) (1.10)
i=1

In this case the solution for the potential is a combination of terms propor-
tional to 1/|x − x0 |.

1.2 Green Theorem

If in the electrostatic problem involved localized discrete or continuous distri-


butions of charge with no boundary surfaces, the general solution (1.6) would
be the most convenient and straight forward solution to any problem.
To handle the boundary conditions it is necessary to develop some new
mathematical tools, namely, the identities or theorems due to George Green
(1824). These follow as simple applications of the divergence theorem
Z I
∇ · A d3 x = A · n da (1.11)
V S
1.2 Green Theorem 3

which applies to any well-behaved vector field A defined in the volume V


bounded by the closed surface S.
Let A = Φ∇ψ, (Φ and ψ arbitrary scalar fields). Then

∇ · (Φ∇ψ) = Φ∇2 ψ + ∇Φ · ∇ψ (1.12)


∂ψ
Φ∇ψ · n = Φ (1.13)
∂n
where ∂/∂n is the normal derivative at the surface S.
When (1.12) and (1.13) substituted into the divergence theorem (1.8) pro-
duces the so-called Green’s 1st identity
Z I
∂ψ
Φ∇2 ψ + ∇Φ · ∇ψ d3 x =

Φ da . (1.14)
V S ∂n

If we rewrite (1.14) with Φ and ψ interchanged, and then subtract it from


(1.14) we obtain Green’s 2nd identity or Green’s Theorem:
Z I  
2 2
 3 ∂ψ ∂Φ
Φ∇ ψ − ψ∇ Φ d x = Φ −ψ da (1.15)
V S ∂n ∂n

Now we can apply Poisson’s equation (1.8) for discrete charge, substituting
for ψ = 1/|x − x0 |

4πρ(x0 ) 3
Z  
3 0 0
−4πδ (x − x )Φ(x ) + d x
V |x − x0 |
I    
∂ 1 1 ∂Φ
= Φ 0 − da0 (1.16)
S ∂n |x − x0 | |x − x0 | ∂n0

Integrating the Dirac delta function over all values of x0 within V and for
x within the volume V yields a nonzero result

ρ(x0 ) 3 0
Z I   
1 1 ∂Φ ∂ 1
Φ(x) = 0
d x + −Φ 0 da0
V |x − x | 4π S |x − x0 | ∂n0 ∂n |x − x0 |
(1.17)
• The (blue) correction term goes to zero as the surface S goes to infinity
(because S falls of faster than 1/|x − x0 |)
• If the integration volume is free of charges, then the first term of equation
(1.17) becomes zero, and the potential is determined only by the values of the
potential and the values of its derivatives at the boundary of the integration
region (the surface S).
• Physical experience leads us to believe that specification of the po-
tential on a closed surface defines a unique potential problem. This is called
Dirichlet problem or Dirichlet boundary conditions.
• Similarly it is plausible that specification of the electric field (normal
derivative of the potential) everywhere on the surface (corresponding to a
4 1 Electrostatics-Magnetostatics (****)

given surface-charge density) also defines a unique problem. The specification


of the normal derivative is known as the Newmann boundary condition.
• As it turns out either one of the two conditions results in unique
solution. It should be clear that a solution to the Poisson eqn with both
Φ and ∂Φ/∂n specified arbitrarily on a closed boundary (Cauchy boundary
conditions) does not exist since there are unique solutions for Dirichlet and
Newmann condition separately.
The solution of the Poisson or Laplace eqn in a finite volume V with
either Dirichlet or Neumann boundary conditions on the boundary surface S
can be obtained by means of Green’s theorem (1.15) and the so-called Green
functions.
In obtaining the result (1.17) we have chosen ψ = 1/|x − x0 | satisfying
 
1
∇2 = −4πδ(|x − x0 |) (1.18)
|x − x0 |
The function 1/|x − x0 | is only one of a class of functions depending on the
variables x and x0 and called Green functions.
In general
∇02 G (x, x0 ) = −4πδ(|x − x0 |) (1.19)
where
1
G (x, x0 ) = + F (x, x0 ) (1.20)
|x − x0 |
and F satisfying the Laplace equation inside the volume V

∇02 F (x, x0 ) = 0 (1.21)

If we substitute G(x, x0 ) in eqn (1.17) we get


Z I  
1 ∂Φ ∂
Φ(x) = ρ(x0 )G(x, x0 )d3 x0 + G(x, x0 ) 0 − Φ(x0 ) 0 G(x, x0 ) da0
V 4π S ∂n ∂n
(1.22)
The freedom in the definition of G means that we can make the surface integral
depend only on the chosen type of BC.
• For the Dirichlet BC we demand:

GD (x, x0 ) = 0 for x0 ∈ S (1.23)

Then the 1st term on the surface integral of (1.22) vanishes


Z I
1 ∂
Φ(x) = ρ(x0 )GD (x, x0 )d3 x0 − Φ(x0 ) 0 GD (x, x0 )da0 (1.24)
V 4π S ∂n
• For Neumann BC the simplest choice of BC on G is
∂GN
(x, x0 ) = 0 for x0 ∈ S (1.25)
∂n0
1.3 Laplace Equation in Spherical Coordinates 5

but application o the Gauss theorem on (1.19) shows that (how?)


I
∂GN 0
0
da = −4π 6= 0
S ∂n

which is incosistent with ∇02 G(x, x0 ) = −4πδ 3 (x − x0 ).


This will mean that the outflux of G cannot be zero when there is a source
enclosed by S. Then the simplest boundary condition that we can use is
∂GN 4π
0
(x, x0 ) = − for x0 ∈ S (1.26)
∂n S
S is the total area of the boundary surface. Then the solution will be:
Z I
0 0 3 0 1 ∂Φ
Φ(x) = ρ(x )GN (x, x )d x + hΦiS + GN (x, x0 )da0 (1.27)
V 4π S ∂n0

where hΦiS is the average value of the potential over the whole surface
I
1
hΦiS ≡ Φ(x0 )da0 (1.28)
S S

In most cases S is extremely large (or even infinite), in which case hΦiS → 0.
The physical meaning of F (x, x0 ) : it is a solution of the Laplace eqn
inside V and so represents the potential of charges external to the volume V
chosen as to satisfy the homogeneous BC of zero potential on the surface S.
• It is important to understand that no matter how the source is dis-
tributed, finding the Green function is completely independent or ρ(x0 ).
• G(x, x0 ) depends exclusively on the geometry of the problem, is a “tem-
plate”, potential and not the actual potential for a given physical problem.
• In other words G(x, x0 ) is the potential due to a unit charge, positioned
arbitrarily within the surface S consistent either with GD = 0 or ∂GN /∂n =
−4π/S on the surface.
• The “true potential” is a convolution of this template with the given
ρ(x0 ).

1.3 Laplace Equation in Spherical Coordinates


In spherical coordinates (r, θ, φ) the Laplace equation can be written in the
form
1 ∂2 ∂2Φ
 
1 ∂ ∂Φ 1
2 2
(rΦ) + 2
sin θ + 2 2 =0 (1.29)
r ∂r r sin θ ∂θ ∂θ r sin θ ∂φ2
If we assume
1
Φ= U (r)P (θ)Q(φ) (1.30)
r
6 1 Electrostatics-Magnetostatics (****)

Then by substituting in (1.29) and multiplying with r2 sin2 θ/U P Q we obtain

1 d2 U 1 d2 Q
  
2 2 1 d dP
r sin θ + sin θ + =0 (1.31)
U dr2 r2 sin θP dθ dθ Q dφ2

We see that the therms depending on φ have been isolated and we can set

1 d2 Q
= −m2 (1.32)
Q dφ2
with solution
Q = e±imφ (1.33)

Similarly the remaining terms can be separated as:

m2
   
1 d dP
sin θ + l(l + 1) − P =0 (1.34)
r2 sin θ dθ dθ sin2 θ
d2 U l(l + 1)
2
− U =0 (1.35)
dr r2
The radial equation will have a solution

U = Arl+1 + Br−l (1.36)

while l is still undetermined.

1.4 Legendre Equation and Legendre Polynomials


The equation (1.34) for P (θ) can be expressed in terms of x = cos θ

m2
   
d 2 dP
(1 − x ) + l(l + 1) − P =0 (1.37)
dx dx 1 − x2

This is the generalized Legendre equation and its solutions are the associated
Legendre functions.
We will consider the solution of (1.54) for m2 = 0
 
d dP
(1 − x2 ) + l(l + 1)P = 0 (1.38)
dx dx

The solution should be single valued, finite, and continuous on the interval
−1 ≤ x ≤ 1 in order that it represents a physical potential.
The solution can be found in the form of a power series

X
P (x) = xk aj xj (1.39)
j=0
1.4 Legendre Equation and Legendre Polynomials 7

where k is a parameter to be determined.


By substitution in (1.54) we get the recurrence relation (how?)

(k + j)(k + j + 1) − l(l + 1)
aj+2 = aj (1.40)
(k + j + 1)(k + j + 2)
while for j = 0, 1 we find that

if a0 6= 0 then k(k − 1) = 0 (1.41)


if a1 6= 0 then k(k + 1) = 0 (1.42)

These two relations are equivalent and it is sufficient to choose either a0 or a1


different from zero but not both (why?). We also see that the series expansion
is either only odd or only on even powers of x. By choosing either k = 0 or
k = 1 it is possible to prove the following properties:
• The series converges for x2 < 1, independent of the value of l
• The series diverges for x = ±1, unless it terminates.
Since we want solution that is finite at x = ±1, as well as for x2 < 1,
we demand that the series terminates.
Since k and j are positive integers or zero, the recurrence relation (1.40)
will terminate only if l is zero or a positive integer.
If l is even (odd), then only the k = 0 (k = 1) series terminates.
The polynomials in each case have xl as their highest power. By convention
these polynomials are normalized to have the value unity for x = +1 and are
called the Legendre polynomials of order l.
The first few are:

P0 (x) = 1
P1 (x) = x
1
P2 (x) = (3x2 − 1) (1.43)
2
1
P3 (x) = (5x3 − 3x)
2
1
P4 (x) = (35x4 − 30x2 + 3)
8
The Legendre polynomials can be taken from Rodrigue’s formula:

1 dl 2
Pl (x) = (x − 1)l (1.44)
2l l! dxl
The Legendre polynomials form a complete orthonormal set of functions
on the interval −1 ≤ x ≤ 1 (prove it)
Z 1
2
Pl0 (x)Pl (x) = δl 0 l (1.45)
−1 2l + 1
8 1 Electrostatics-Magnetostatics (****)

Since the Legendre polynomials form a complete set of orthonormal func-


tions, any function f (x) on the interval −1 ≤ x ≤ 1 can be expanded in terms
of them i.e.

X
f (x) = Al Pl (x) (1.46)
l=0

where (how?)
Z 1
2l + 1
Al = f (x)Pl (x)dx (1.47)
2 −1
Thus for problems with azimuthal symmetry i.e. m = 0 the general solution
is:
∞ h
X i
Φ(r, θ) = Al rl + Bl r−(l+1) Pl (cos θ) (1.48)
l=0

where the coefficients Al [it is not the same as in eqn (1.47)] and Bl can be
determined from the boundary conditions.

1.4.1 Example

Boundary Value Problems with Azimuthal Symmetry

Let’s specify as V (θ) the potential on the surface of a sphere of radius R,


and try to find the potential inside the sphere.
If there are no charges at the origin (r = 0) the potential must be finite
there. Consequently Bl = 0 for all l. Then on the surface of the sphere

X
V (r = R, θ) = Al Rl Pl (cos θ) (1.49)
l=0

and the coefficients Al will be taken via eqn (1.47)


2l + 1 π
Z
Al = V (θ)Pl (cos θ) sin θdθ (1.50)
2Rl 0

If, for example V (θ) = ±V on the two hemispheres then the coefficients
can be derived easily and the potential inside the sphere is (how?):
   
3 r 7  r 3 11  r 5
Φ(r, θ) = V P1 (cos θ) − P3 (cos θ) + P5 (cos θ) − . . .
2 R 8 R 16 R
(1.51)
To find the potential outside the sphere we merely replace (r/R)l by
(R/r)l+1 and the resulting potential will be (how?):
 2 "  2 #
3 R 7 R
Φ(r, θ) = V P1 (cos θ) − P3 (cos θ) + . . . (1.52)
2 r 12 r
1.5 Legendre Polynomials 9

Fig. 1.1.

1.5 Legendre Polynomials


An important expansion is that of the potential at x due to a unit point charge
at x0
∞ l
1 X r<
0
= l+1
Pl (cos γ) (1.53)
|x − x | r
l=0 >

where r< (r> ) is the smaller (larger) of |x| and |x0 | and γ is the angle between
|x| and |x0 |.

Fig. 1.2.

Show that the potential is :



1 X
l −(l+1)

= A l r + B l r Pl (cos γ) on the z-axis
|x − x0 |
l=0
10 1 Electrostatics-Magnetostatics (****)
∞  l
1 1 X r<
= on the x-axis
|x − x0 | r> r>
l=0

1.6 Associated Legendre Functions and Spherical


Harmonics
For problems without azimuthal (axial) symmetry, we need the generalization
of Pl (cos θ), namely, the solutions of

m2
   
d dP
(1 − x2 ) + l(l + 1) − P =0 (1.54)
dx dx 1 − x2

for arbitrary l and m.


It can be shown that in order to have finite solutions on the interval −1 ≤
x ≤ 1 the parameter l must be zero or a positive integer and that the integer
m can take on only the values −l, −(l − 1), ..., 0 , ... ,(l − 1), l (why?).
The solution having these properties is called an associated Legendre
function Plm (x) . For positive m it is defined as

dm
Plm (x) = (−1)m (1 − x2 )m/2 Pl (x) (1.55)
dxm
If Rodrigues’ formula is used an expression valid for both positive and negative
m is obtained:
(−1)m 2 m/2 d
m
Plm (x) = (1 − x ) Pl (x) (1.56)
2l l! dxm

There is a simple relation between Pl−m (x) and Plm (x) :

(l − m)! m
Pl−m (x) = (−1)m P (x) (1.57)
(l + m)! l

For fixed m the functions Plm (x) form an orthonormal set in the index l on
the interval −1 ≤ x ≤ 1. The orthogonality relation is
Z 1
2 (l + m)!
Plm m
0 (x)Pl (x)dx = δl 0 l (1.58)
−1 2l + 1 (l − m)!

We have found that Qm (φ) = eimφ , this function forms a complete set of
orthogonal functions in the index m on the interval 0 ≤ φ ≤ 2π.
The product Plm Qm forms also a complete orthonormal set on the surface
of the unit sphere in the two indices l, m.
From the normalization condition (1.58) we can conclude that the suitably
normalized functions, denoted by Ylm (θ, φ), are :
1.6 Associated Legendre Functions and Spherical Harmonics 11
s
2l + 1 (l − m)! m
Ylm (θ, φ) = P (cos θ)eimφ (1.59)
4π (l + m)! l
and also

Yl,−m (θ, φ) = (−1)m Ylm (θ, φ) (1.60)

The normalization and orthogonality conditions are:


Z 2π Z π
dφ sin θdθYl∗0 m0 (θ, φ)Ylm (θ, φ) = δl0 l δm0 m (1.61)
0 0

An arbitrary function g(θ, φ) can be expanded in spherical harmonics


∞ X
X l
g(θ, φ) = Alm Ylm (θ, φ) (1.62)
l=0 m=−l

where the coefficients are


Z

Alm = dΩYlm (θ, φ)g(θ, φ) . (1.63)

The general solution for a boundary-value problem in spherical coordinates


can be written in terms of spherical harmonics and powers of r in a general-
ization of (1.48) :
∞ X
X l h i
Φ(r, θ, φ) = Alm rl + Blm r−(l+1) Ylm (θ, φ) (1.64)
l=0 m=−l

If the potential is specified on a spherical surface, the coefficients can be


determined by evaluating (1.64) on the surface and using (1.63).

1.6.1 Spherical Harmonics Ylm (θ, φ)


r
1
l = 0 Y00 =

r
3
l = 1 Y11 = − sin θeiφ

r
3
Y10 = cos θ

r
1 15
l = 2 Y22 = sin2 θe2iφ
4 2π
r
15
Y21 = − sin θ cos θ eiφ

r  
5 3 1
Y20 = cos2 θ −
4π 2 2
12 1 Electrostatics-Magnetostatics (****)

r
1 35
l = 3 Y33 = − sin3 θ e3iφ
4 4π
r
1 105
Y32 = sin2 θ cos θe2iφ
4 2π
r
1 21
Y31 = − sin θ(5 cos2 θ − 1) eiφ
4 4π
r
1 7
sin θ 5 cos3 θ − 3 cos θ

Y30 =
2 4π
(1.65)

Fig. 1.3. Schematic representation of Ylm on the unit sphere. Ylm is equal to 0 along
m great circles passing through the poles, and along l − m circles of equal latitude.
The function changes sign each time it crosses one of these lines.

1.6.2 Addition Theorem for Spherical Harmonics

The spherical harmonics are related to Legendre polynomials Pl by a relation


known as the addition theorem.
The addition theorem expresses a Legendre polynomial of order l in the
angle γ in terms of products of the spherical harmonics of the angles θ, φ and
θ 0 , φ0 :
1.6 Associated Legendre Functions and Spherical Harmonics 13

Fig. 1.4.

l
4π X

Pl (cos γ) = Ylm (θ0 , φ0 )Ylm (θ, φ) (1.66)
l(l + 1)
m=−l

where γ is the angle between the vectors x and x0 , x · x0 = x · x0 cos γ and


cos γ = cos θ cos θ0 + sin θ sin θ0 cos(φ − φ0 ).
Pl (cos γ) is a function of the angles θ, φ with the angles θ0 , φ0 as parameters
and it maybe expanded in a series (1.63) :
0

X l
X
Pl (cos γ) = Al0 m0 (θ0 , φ0 )Yl0 m0 (θ, φ) (1.67)
l0 =0 m=−l0
14 1 Electrostatics-Magnetostatics (****)

The addition theorem offer the possibility to extend the expansion valid
for a point charge (axially symmetric distribution) to an arbitrary charge
distribution.
By substituting (1.66) into (1.53) we obtain
∞ X
l l
1 X 1 r<
= 4π Y ∗ (θ0 , φ0 )Ylm (θ, φ) (1.68)
0
|x − x | 2l + 1 r> lm
l+1
l=0 m=−l

This equation gives the potential in a completely factorized form in the coor-
dinates x and x0 . This is useful in any integration over charge densities, when
one is the variable of integration and the other the observation point.

1.7 Multipole Expansion


A localized distribution of charge is described by the charge density ρ(x0 ),
which is nonvanishing only inside a sphere a around some origin.
The potential outside the sphere can be written as an expansion in spher-
ical harmonics
∞ X
l
X 4π 1
Φ(x) = qlm l+1 Ylm (θ, φ) (1.69)
2l + 1 r
l=0 m=−l

Fig. 1.5.

This type of expansion is called multipole expansion;


The l = 0 term is called monopole term,
The l = 1 are called dipole terms etc.
1.7 Multipole Expansion 15

The problem to be solved is the determination of the constants qlm in


terms of the properties of the charge distribution ρ(x0 ).
The solution is very easily obtained from the integral for the potential

ρ(x0 ) 3 0
Z
Φ(x) = d x (1.70)
|x − x0 |

Using the expansion (1.68) for 1/|x − x0 | i.e.


∞ X
l Z
X 4π 1 ∗
Φ(x) = l+1
Ylm (θ, φ) Ylm (θ0 , φ0 )ρ(x0 )r0l d3 x0 (1.71)
2l + 1 r
l=0 m=−l

Consequently the coefficients in (1.69) are :


Z

qlm = Ylm (θ0 , φ0 )ρ(x0 )r0l d3 x0 (1.72)

and called the multipole moments of the charge distribution ρ(x0 )

1.7.1 Monopole moment l = 0

Here, the only component is


Z Z
0 00 ∗ 0 0 3 0 1 q
q00 = ρ(x )r Y00 (θ , φ )d x = √ ρ(x0 )d3 x0 = √ (1.73)
4π 4π
Observed from a large distance r, any charge distribution acts approximately
as if the total charge q (monopole moment) would be concentrated at one
point since the dominant term in (1.68)

∗ 1 q
Φ(x) = 4πq00 Y00 + · · · = + . . . (1.74)
r r

NOTE:
The moments with m ≥ 0 are connected (for real charge density) too the
moments with m < 0 through

ql,−m = (−1)m qlm (1.75)

1.7.2 Dipole moment l = 1

In Cartesian coordinates the dipole moment is given by


Z
p = x0 ρ(x0 )d3 x0 (1.76)
16 1 Electrostatics-Magnetostatics (****)

In Spherical representation one obtains


Z

q11 = ρ(x0 )r0 Y11

(θ0 , φ0 )d3 x0
r Z
3
=− ρ(x0 )r0 (sin θ0 cos φ0 − i sin θ0 sin φ0 ) d3 x0

in Cartesian represantation
r Z r
3 0 0 0 3 0 3
=− ρ(x )(x − iy )d x = − (px − ipy ) (1.77)
8π 8π
also
Z r Z
∗ 0 0 ∗ 3
q10 = ρ(x )r Y10 (θ0 , φ0 )d3 x0
= ρ(x0 )r0 cos θ0 d3 x0

r Z r
3 3
= z 0 ρ(x0 )d3 x = pz (1.78)
4π 4π

1.7.3 Quadrupole moment l = 2

Z

q22 = ρ(x)r02 Y22

(θ0 , φ0 )d3 x0
r Z
1 15 2
= ρ(x) [r0 sin θ0 (cos φ0 − i sin φ0 )] d3 x0
4 2π
r Z
1 15
= ρ(x)(x0 − iy 0 )2 d3 x0
4 2π
1  02
because (x0 − iy 0 )2 = (3x − r02 ) − 6ix0 y 0 − (3y 02 − r02 )

r 3
1 15
= (Q11 − 2iQ12 − Q22 ) (1.79)
12 2π
where Qij is the traceless (why?) quadrupole moment tensor:
Z
3x0i x0j − r2 δij ρ(x0 )d3 x0

Qij = (1.80)

Analogously
Z r Z
∗ 0 02 ∗ 15
q21 = ρ(x )r Y21 (θ0 , φ0 )d3 x0 =− ρ(x0 )z 0 (x0 − iy 0 )d3 x0

r
1 15
=− (Q13 − iQ23 ) (1.81)
3 8π
and
1.7 Multipole Expansion 17
Z r Z
∗ 1 5
q20 = ρ(x0 )r02 Y20

(θ0 , φ0 )d3 x0 = ρ(x0 )(3z 0 − r02 )d3 x0
2 4π
r
1 5
= Q33 (1.82)
2 4π
From eqn (1.60) we can get the moments with m < 0 through

ql,−m = (−1)m qlm (1.83)

1.7.4 Multipole Expansion

By direct Taylor expansion of 1/|x − x0 | the expansion of Φ(x) in rectangular


coordinates is:
q p·x 1X xi xj
Φ(x) = + 3 + Qij 5 + . . . (1.84)
r r 2 i,j r

The electric field components for a given multipole can be expressed most
easily in terms of spherical coordinates. From E = −∇Φ and (1.69) for fixed
(l, m) we get(how?):

4π(l + 1) 1
Er = qlm l+2 Ylm
2l + 1 r
4π 1 ∂
Eθ = − qlm l+2 Ylm (1.85)
2l + 1 r ∂θ
4π 1 1 ∂
Eφ = − qlm l+2 Ylm
2l + 1 r sin θ ∂φ
For a dipole p along the z-axis they reduce to:
2p cos θ p sin θ
Er = , Eθ = , Eφ = 0 (1.86)
r3 r3

The field intensity at a point x due to a dipole p at a point x0 (r = |x−x0 |)


is
3n (p · n) − p r
E(x) = with n= (1.87)
|x − x0 |3 r
because Φ(x) = p · x/r3 and
 
∇(p · x) 1 3n (p · n) − p
E(x) = −∇Φ = − − (p · x)∇ = (1.88)
r3 r3 |x − x0 |3
18 1 Electrostatics-Magnetostatics (****)

1.8 Energy of a Charge Distribution in an External Field


The multipole expansion of the potential of a charge distribution can also be
used to describe the interaction of the charge distribution with an external
field.
The electrostatic energy of the charge distribution ρ(x) placed in an external
field Φ(x) is given by Z
W = ρ(x)Φ(x)d3 x (1.89)
V
The external field (if its is slowly varying over the region where ρ(x) is non-

Fig. 1.6.

negligible) may be expanded in a Taylor series:


3 3
1 XX ∂2Φ
Φ(x) = Φ(0) + x · ∇Φ(0) + xi xj (0) + . . . (1.90)
2 i=1 j=1 ∂xi ∂xj

Since E = −∇Φ for the external field


3 3
1 XX ∂Ej
Φ(x) = Φ(0) − x · E(0) − xi xj (0) + . . . (1.91)
2 i=1 j=1 ∂xi

Since ∇ · E = 0 for the external field we can substract


1 2
r ∇ · E(0)
6
from the last term to obtain finally the expansion:
3 3
1 XX  ∂Ej
Φ(x) = Φ(0) − x · E(0) − 3xi xj − r2 δij (0) + . . . (1.92)
6 i=1 j=1 ∂xi
1.8 Energy of a Charge Distribution in an External Field 19

When this inserted into (1.89) the energy takes the form:
3 3
1 XX ∂Ej
W = qΦ(0) − p · E(0) − Qij (0) + . . . (1.93)
6 i=1 j=1 ∂xi

Notice, that:
• the total charge interacts with the potential,
• the dipole moment with the electric field,
• the quadrupole with the electric field gradient, etc

1.8.1 Examples
EXAMPLE 1 : Show that for a uniform charged sphere all multipole mo-
ments vanish except q00 .
If the sphere has a radius R0 and constant charge density ρ(x) = ρ then
Z Z R0
Rl+3
Z

qlm =ρ r0l Ylm

(θ0 , φ0 )r02 dr0 dΩ 0 = ρ 0 Y ∗ (θ0 , φ0 )dΩ 0
Ω0 0 l + 3 Ω 0 lm

but since Y00 = 1/ 4π we have from the orthogonality relation
Rl+3 √ Rl+3 √
Z
∗ ∗
qlm =ρ 0 4π Ylm Y00 dΩ 0 = ρ 0 4πδl0 δm0
l+3 Ω0 l+3
EXAMPLE 2: Perform multipole decomposition of a uniform charge dis-
tribution whose surface is a weakly deformed sphere:
2
!
X

R = R0 1 + a2m Y2m (θ, φ) , |a2m | << 1
m=−2

The multipole moments are

Z Z R(θ 0 ,φ0 )

qlm = ρr0l Ylm
∗ 02 0
r dr dΩ 0
Ω0 0
2
!l+3
Rl+3
Z X

=ρ 0 Ylm (θ0 , φ0 ) 1+ a∗2m Y2m

(θ0 , φ0 ) dΩ 0
l+3 Ω0 m=−2
20 1 Electrostatics-Magnetostatics (****)

2
!
Rl+3
Z X
∗ ∗
qlm ≈ρ 0 Ylm (θ0 , φ0 ) 1 + (l + 3) a∗2m Y2m

(θ0 , φ0 ) dΩ 0
l+3 Ω0 m=−2

Apart from the monopole moment q00 of the previous example, we find
2
R0l+3 X Z X

qlm =ρ (l + 3) a∗2µ ∗
Ylm Y2µ dΩ 0 = ρR0l+3 a∗2µ δmµ δl2 (l > 0)
l+3 µ=−2 Ω 0
µ


Thus the nonvanishing multipole moments are : q2m = ρR05 a∗2m .

1.9 Magnetostatics
• There is a radical difference between magnetostatics and electrostatics :
there are no free magnetic charges
• The basic entity in magnetic studies is the magnetic dipole.
• In the presence of magnetic materials the dipole tends to align itself in
a certain direction. That is by definition the direction of the magnetic flux
density , denoted by B (also called magnetic induction).
• The magnitude of the flux density can be defined by the magnetic
torque N exerted on the magnetic dipole:

N=µ×B (1.94)

where µ is the magnetic moment of the dipole.


• In electrostatics the conservation of charge demands that the charge
density at any point in space be related to the current density in that neigh-
borhood by the continuity equation

∂ρ/∂t + ∇ · J = 0 (1.95)

• Steady-state magnetic phenomena are characterized by no change in the


net charge density anywhere in space, consequently in magnetostatics

∇·J=0 (1.96)

1.10 Biot & Savart Law


If d` is an element of length (pointing in the direction of current flow) of a
wire which carries a current I and x is the coordinate vector from the element
1.10 Biot & Savart Law 21

of length to an observation point P , then the elementary flux density dB at


the point P is given in magnitude and direction by
d` × x
dB = kI (1.97)
|x|3

NOTE: (1.97) is an inverse square law, just as is Coulomb’s law of electro-

Fig. 1.7.

statics. However, the vector character is very different.


• If we consider that current is charge in motion and replace Id` by qv
where q is the charge and v the velocity. The flux density for such a charge
in motion would be
v×x
B = kq (1.98)
|x|3
This expression is time-dependent and valid only for charges with small ve-
locities compared to the speed of light.

1.10.1 Diff. Equations for Magnetostatics & Ampere’s Law

The basic law (1.97) for the magnetic induction can be written down in general
form for a current density J(x):

(x − x0 ) 3 0
Z
1
B(x) = J(x0 ) × d x (1.99)
c |x − x0 |3

This expression for B(x) is the magnetic analog of electric field in terms of
the charge density:

(x − x0 ) 3 0
Z
1
E(x) = ρ(x0 ) d x (1.100)
c |x − x0 |3

In order to obtain DE equivalent to (1.99) we use the relation


22 1 Electrostatics-Magnetostatics (****)

(x − x0 )
   
1 1 r
= −∇ as ∇ =− 3 (1.101)
|x − x0 |3 |x − x0 | r r

and (1.99) transforms into

J(x0 ) 3 0
Z
1
B(x) = ∇ × d x (1.102)
c |x − x0 |

From (1.102) follows immediately:

∇·B=0 (1.103)

• This is the 1st equation of magnetostatics and corresponds to ∇×E = 0


in electrostatics.
By analogy with electrostatics we now calculate the curl of B

J(x0 ) 3 0
Z
1
∇×B= ∇×∇× d x (1.104)
c |x − x0 |
1
which for steady-state phenomena (∇ · J = 0) reduces to (how?)

∇×B= J (1.105)
c
• This is the 2nd equation of magnetostatics and corresponds to ∇ · E =
4πρ in electrostatics.
The integral equivalent of (1.105) is called Ampere’s law It can be ob-
tained by applying Stokes’s theorem to the integral of the normal component
of (1.105) over the open surface S bounded by a closed curve C. Thus
Z Z

∇ × B · n da = J · nda (1.106)
S c S

is transformed into I Z

B · d` = J · nda (1.107)
C c S
Since the surface integral of the current density is the total current I passing
through the closed curve C, Ampére’s law can be written in the form:
I

B · d` = I (1.108)
C c

1
Remember : ∇ × (∇ × A) = ∇ (∇ · A) − ∇2 A
1.11 Vector Potential 23

Fig. 1.8.

1.11 Vector Potential


The basic differential laws in magnetostatics are

∇×B= J and ∇ · B = 0 (1.109)
c
The problems is how to solve them.
• If the current density is zero in the region of interest, ∇ × B = 0
permits the expression of the vector magnetic induction B as the gradient of
a magnetic scalar potential B = −∇ΦM , then (1.109) reduces to the Laplace
equation for ΦM .
• If ∇ · B = 0 everywhere, B must be the curl of some vector field A(x),
called the vector potential

B(x) = ∇ × A(x) (1.110)

and from (1.102) the general form of A is:


J(x0 ) 3 0
Z
1
A(x) = d x + ∇Ψ (x) (1.111)
c |x − x0 |

The added gradient of an arbitrary scalar function Ψ shows that, for a


given magnetic induction B, the vector potential can be freely transformed
according to
A → A + ∇Ψ (1.112)
This transformation is called a gauge transformation. Such transformations
are possible because (1.110) specifies only the curl of A.
• If (1.110) is substituted into the first equation in (1.109), we find
4π 4π
∇ × (∇ × A) = J or ∇ (∇ · A) − ∇2 A = J (1.113)
c c
If we exploit the freedom implied by (1.112) we can make the convenient
choice of gauge (Coulomb gauge) ∇ · A = 0. Then each component of the
vector potential satisfies the Poisson equation
24 1 Electrostatics-Magnetostatics (****)


∇2 A = − J (1.114)
c
The solution for A in unbounded space is (1.111) with Ψ =constant:

J(x0 ) 3 0
Z
1
A(x) = d x (1.115)
c |x − x0 |
2
Classical Field Theory: Maxwell Equations

2.1 Introduction
• Electrostatics and Magnetostatics deal with steady-state problems in
electricity and in magnetism.
• The almost independent nature of electric and magnetic fields phenomena
disappears when we consider time-dependent problems.
• Time varying magnetic fields give rise to electric fields and vice-versa. We
then must speak of electromagnetic fields rather than electric or magnetic
fields.
• The interconnection between electric and magnetic fields and their essen-
tial sameness becomes clear only within the framework of Special Theory
of Relativity

2.2 Faraday’s Law of Induction (****)


Faraday (1831), observed that a transient current is induced in a circuit if:

1. A steady current flowing in an adjacent circuit is turned on or off


2. The adjacent circuit with a steady current flowing is moved relative to
the first circuit
3. A permanent magnet is thrust into or out of the circuit
The changing magnetic flux induces an electric field around the circuit, the
line integral of which is called the electromotive force E and causes a current
flow according to Ohm’s law: J = σE ( σ is the conductivity).
The magnetic induction in the neighboring of the circuit is B and the
magnetic flux linking the circuit is defined by
Z
F = B · n da (2.1)
S
26 2 Classical Field Theory: Maxwell Equations

Fig. 2.1.

The electromotive force around the circuit is defined by


I
E= E0 · d` (2.2)
C

where E0 is the electric field at the element d` of the circuit C.


Thus Faraday’s observation is summed up in the mathematical law:
dF
E = −k (2.3)
dt
That is, the induced electromotive force around the circuit is proportional
to the time rate of change of magnetic flux linking the circuit.
The sign is specified by Lenz’s law, which states that the induced current
(and accompanying magnetic flux) is in such a direction to oppose the change
of flux through the circuit. For SI units k = 1, Gaussian units k = 1/c.

2.2.1 Faraday’s law for a moving circuit


I Z
d
E0 · d` = −k B · n da (2.4)
C dt S
This is eqn (2.3) in terms of integrals. We can observe that :
• The induced electromotive force is proportional to the total time
derivative of the flux.
• The flux can be changed by changing: the magnetic induction or the
shape or the orientation or the position of the circuit.
2.2 Faraday’s Law of Induction (****) 27

Fig. 2.2.

If the circuit C is moving with a velocity v in some direction the total time
derivative in eqn (2.4) must take into account the motion e.g. convective
derivative
d ∂
= + v · ∇ and
dt ∂t
dB ∂B ∂B
= + (v · ∇) B = + ∇ × (B × v) + v (∇ · B)
dt ∂t ∂t
The flux through the circuit may change because:
• the flux changes with time at a point
• the translation of the circuit changes the location of the boundary.
Eqn (2.4) can be written as (why?):
I Z
∂B
[E0 − k (v × B)] · d` = −k · n da . (2.5)
C S ∂t

In a comoving coordinate system


I Z
∂B
E · d` = −k · n da (2.6)
C S ∂t

where E is the electric field in the comoving frame. Thus

E0 = E + k (v × B) (2.7)

With the present choice of units for charge and current, Galilean covariance
requires that k = 1/c (why?).
Finally, the transformation of the electromotive force integral into a surface
integral leads to (how?)
Z  
1 ∂B
∇×E+ · n da = 0 (2.8)
S c ∂t
28 2 Classical Field Theory: Maxwell Equations

Since the circuit C and the bounding surface S are abritrary, the integrant
must vanish at all points in space. Thus the differential form of Faraday’s
law is:
1 ∂B
∇×E+ = 0. (2.9)
c ∂t
Note that for time-independent electrostatic fields : ∇ × E = 0.
KELVIN-STOKES THEOREM :
relates the surface integral of the curl of a vector field over a surface S in
Euclidean 3-space to the line integral of the vector field over its boundary
Z I
∇ × F · dS = F · d` (2.10)
S ∂S

z
n

da
S
C
0
x y

Fig. 2.3.

2.3 Energy in the Magnetic Field (***)


• Typically in magnetostatics, the creation of a steady-state configuration of
currents and associated magnetic fields involves an initial transient period
during which the currents and fields are brought from zero to the final
values.
2.3 Energy in the Magnetic Field (***) 29

• For such time-varying fields there are induced electromotive forces which
cause the sources of current to do work.
• Since the energy in the field is by definition the total work done to establish
it, we must consider these contributions.
• Suppose that we have a circuit with a constant current I flowing in it.
• If the flux through the circuit changes an electromotive force E is induced
around it.
• In order to keep the current constant, the sources of current must
do work.
• To determine the rate we note that the time rate of change of energy of
a particle with velocity v acted on by a force F is
dE
=v·F
dt
With a changing flux the added field E0 on each conducting electron of charge
q and drift velocity v gives rise to a change in energy per unit time of qv · E0
per electron.
Summing over all the electrons in the circuit, we find that the sources do
work to maintain the current at the rate
dW 1 dF
= −IE = I
dt c dt

Thus, if the flux change through a circuit carrying a current I is δF , the


work done by the source is :
1
δW = IδF
c

We consider the problem of the work done in establishing a general


steady-state distribution of currents and fields.
• We can imagine that the build-up process occurs at an infinitesimal rate
so that ∇ · J = 0 (J is the current).
• The current distribution can be broken up into a network of elementary
loops, the typical ones of which is an elemental tube of current of cross-
sectional area ∆σ following a close path C and spanned by a surface S
with normal n
We can express the increment of work done against the induced electro-
motive force in terms of the change of the magnetic induction through the
loop: Z
J∆σ
∆(δW ) = n · δBda (2.11)
c S
the extra ∆ is because we consider only one elementary circuit.
30 2 Classical Field Theory: Maxwell Equations

Fig. 2.4. Distribution of current density broken up into elementary current loops

• If we express B in terms of the vector potential A i.e. B = ∇ × A then


we have Z
J∆σ
∆(δW ) = (∇ × δA) · n da (2.12)
c S
With application of Stokes’s theorem this can be written
I
J∆σ
∆(δW ) = δA · d` (2.13)
c C

but J∆σd` ≡ Jd3 x since l is parallel to J. Evidently the sum over all such
elemental loops will be the volume integral.
• Hence the total increment of work done by an external source due to a
change δA(x) in the vector potential is
Z
1
δW = δA · J d3 x (2.14)
c

By using Ampère’s law



∇×H= J
c
we can get an expression in terms of the magnetic fields. Then
Z
1
δW = δA · (∇ × H) d3 x (2.15)

which transforms to (how?)


Z
1
δW = [H · (∇ × δA) + ∇ · (H × δA)] d3 x (2.16)

If the field distribution is assumed to be localized, the 2nd integrant van-
ishes (why?) and we get
Z
1
δW = H · δBd3 x (2.17)

2.4 Maxwell Equations (****) 31

which is the analog of the electrostatic equation for the energy change
Z
1
δW = E · δDd3 x (2.18)

where D = E + 4πP is the electric displacement and P the electric
polarization (dipole moment per unit volume).
If we bring the fields from zero to the final values the total magnetic
energy will be (why?) Z
1
W = H · B d3 x (2.19)

which is the magnetic analog of the total electrostatic energy
Z
1
W = E · D d3 x (2.20)

The energy of a system of charges in free space i.e. electrostatic energy
is: Z
1
W = ρ(x)Φ(x) d3 x (2.21)
2
The magnetic equivalent for this expression i.e. the magnetic energy is
Z
1
W = J · A d3 x (2.22)
2c

2.4 Maxwell Equations (****)


Maxwell’s equations are based on the following empirical facts:
1. The electric charges are sources of the vector field of the dielectric dis-
placement density D. Hence, for the flux of the dielectric displacement
through a surface enclosing the charge we have
I Z
1
D · nda = Q = ρd3 x (2.23)
4π S V

This relation can be derived from Coulomb’s force law.


2. Faraday’s induction law:
I Z
1 ∂
E= E · dl = − B · nda (2.24)
C c ∂t S
3. The fact that there are no isolated monopoles implies
I
B · nda = 0 (2.25)
S
32 2 Classical Field Theory: Maxwell Equations

4. Ampère’s law: I Z

H · dl = J · nda (2.26)
C c S
The basic laws of electricity and magnetism can be summarized in differ-
ential form:
Coulomb’s law ∇ · D = 4πρ (2.27)
4π 1 ∂D
Ampère’s law (∇ · J = 0) ∇ × H = J+ (2.28)
c c ∂t
1 ∂B
Faraday’s law ∇ × E + =0 (2.29)
c ∂t
no free magnetic poles ∇ · B = 0 (2.30)
where E and B are the averaged E and B of the microscopic or vacuum
Maxwell equations. The two extra field quantities D and H usually called the
electric displacement and magnetic field (B is then called the magnetic
induction and M is the macroscopic magnetization)
!
X ∂Q0
ab
Da = Ea + 4π Pa − + . . . → D = E + 4πP + . . . (2.31)
∂xb
b
Ha = Ba − 4π(Ma + . . . ) → H = B − 4πM + . . .(2.32)

• The quantities P, M, Qab represent the macroscopically averaged electric


dipole, magnetic dipole and electric quadrupole moment densities
of the material medium in the presence of of applied fields.
• Similarly, the charge and current densities ρ and J are macroscopic aver-
ages of the free charge and current densities in the medium.
• The macroscopic Maxwell equations are a set of 8 eqns involving the com-
ponents of 4 fields E, B, D and H.
• The 4 homogeneous eqns can be solved formally by expressing E and B
in terms of the scalar potential Φ and the vector potential A
• The inhomogeneous eqns cannot be solved until the derived fields D and
H are known in terms of E and B. These connections which are implicit
in (2.32) are known as constitute relations, e.g. D = E and H = E/µ
(: electric permittivity & µ magnetic permeability).
• All but Faraday’s law were derived from steady-state observations and
there is no a priori reason to expect that the static equations hold un-
changed for time-dependent fields.
The above equations without the red term in Ampere’s law are inconsis-
tent. While Ampére’s law ( ∇ · J = 0) is valid for steady - state problems, the
complete relation is given by the continuity equation for charge and current
∂ρ
∇·J+ = 0. (2.33)
∂t
2.4 Maxwell Equations (****) 33

Maxwell replaced J in Ampère’s law by its generalization


1 ∂D
J→J+ (2.34)
4π ∂t
Maxwell called the added term displacement current, without it there
would be no electromagnetic radiation (Can you repeat his steps?).
Maxwell’s equations, form the basis of all classical electromag-
netic phenomena. When combined with the Lorentz force equation
 v 
F=q E+ ×B (2.35)
c
and Newton’s 2nd law of motion, provide a complete description of the clas-
sical dynamics of interacting charged particles and EM fields.

2.4.1 Vector and Scalar Potentials

In electrostatics and magnetostatics we have used the scalar potential Φ and


the vector potential A to simplify certain equations.
Since ∇ · B = 0, we can define B in terms of a vector potential A :

B=∇×A (2.36)
1 ∂B
Then Faraday’s law ∇ × E + = 0 can be written
c ∂t
 
1 ∂A
∇× E+ =0 (2.37)
c ∂t
Thus the quantity with vanishing curl can be written as the gradient of a
scalar potential Φ:
1 ∂A 1 ∂A
E+ = −∇Φ or E = −∇Φ − (2.38)
c ∂t c ∂t
The definition of B and E in terms of the potentials A and Φ satisfies inden-
tically the 2 homogeneous Maxwell equations. While A and Φ are determined
by the 2 inhomogeneous Maxwell equations.
If we restrict our considerations to the vacuum form of the Maxwell equa-
tions the inhomogeneous form of Maxwell equations can be write in terms of
the potentials as:
1 ∂
∇2 Φ + (∇ · A) = −4πρ (2.39)
c ∂t
1 ∂2A
 
1 ∂Φ 4π
∇2 A − 2 2 − ∇ ∇ · A + =− J (2.40)
c ∂t c ∂t c
These eqns are equivalent to Maxwell eqns but they are still coupled. Thanks
to the abritrariness in the definition of the potentials we can choose transfor-
mations of the form
34 2 Classical Field Theory: Maxwell Equations

A → A0 = A + ∇Λ (2.41)
1 ∂Λ
Φ → Φ0 = Φ − (2.42)
c ∂t
Thus we can choose a set of potentials (A, Φ) such that

1 ∂Φ
∇·A+ =0 (2.43)
c ∂t

This way we uncouple the pair of equations (2.39) and (2.40) and leave
two inhomogeneous wave equations, one for Φ and one for A

1 ∂2Φ
∇2 Φ − = −4πρ (2.44)
c2 ∂t2
1 ∂2A 4π
∇2 A − 2 2 = − J (2.45)
c ∂t c
This set of equations is equivalent in all respects to the Maxwell
equations.

2.4.2 Gauge Transformations : Lorenz Gauge

The transformations

A → A0 = A + ∇Λ (2.46)
1 ∂Λ
Φ → Φ0 = Φ − (2.47)
c ∂t
are called gauge transformation and the invariance of the fields under these
transformations is called gauge invariance.
The relation between A and Φ :
1 ∂Φ
∇·A+ =0 (2.48)
c ∂t
is called Lorenz condition 1 . [Prove that there will always exist potentials
satisfying the Lorentz condition].
The Lorenz gauge is commonly used because:
• It leads to the wave equations (2.44) and (2.45) which treat Φ
and A on equal footings
• It is a coordinate independent concept and fits naturally into the
considerations of special relativity.
1
The condition is from the Danish mathematician and physicist Ludvig Valentin Lorenz (1829-1891) and not
from the Dutch physicist Hendrik Lorentz (1853-1928)
2.5 Green Functions for the Wave Equations (***) 35

2.4.3 Gauge Transformations : Coulomb Gauge

In this gauge
∇·A=0 (2.49)
From eqn (2.39) we see that the scalar potential satisfies the Poisson eqn :

∇2 Φ = −4πρ (2.50)

with solution
ρ(x0 , t) 3 0
Z
Φ(x, t) = d x (2.51)
|x − x0 |
The scalar potential is just the instantaneous Coulomb potential due to the
charge density ρ(x, t). This is the origin of the name Coulomb gauge.
From eqn (2.40) we find that the vector potential satisfies:

1 ∂2A 4π 1 ∂Φ
∇2 A − 2 2
=− J+ ∇ (2.52)
c ∂t c c ∂t

Finally, if we define the transverse (or solenoidal) current:


Z
1 J
Jt = ∇×∇× d3 x0 (2.53)
4π |x − x0 |

and longitudinal (or irrotational) current Jl for which ∇ × Jl = 0 which may


cancel the contribution of the term with the potential Φ.
Then, the wave equation for A can be expressed entirely in terms of the
transverse current Jt [Can you explain it?]

1 ∂2A 4π
∇2 A − = − Jt (2.54)
c2 ∂t2 c
The Coulomb or transverse gauge is often used when no sources are
present. Then Φ = 0, and A satisfies the homogeneous wave equation. The
fields are given by
1 ∂A
E=− , and B = ∇ × A (2.55)
c ∂t

2.5 Green Functions for the Wave Equations (***)


The wave eqns (2.44), (2.45) and (2.54) all have the same structure,

1 ∂2Ψ
∇2 Ψ − = −4πf (x, t) (2.56)
c2 ∂t2
36 2 Classical Field Theory: Maxwell Equations

• To solve (2.56) it is useful to find a Green function (as in electrostatics)


• In order to remove the time dependence we introduce a Fourier transform
with respect to frequency
• We suppose that both Ψ (x, t) and f (x, t) have a Fourier integral repre-
sentation
Z ∞ Z ∞
1 −iωt 1
Ψ (x, t) = Ψ (x, ω)e dω , f (x, t) = f (x, ω)e−iωt dω (2.57)
2π −∞ 2π −∞

with the inverse transformations,


Z ∞ Z ∞
Ψ (x, ω) = Ψ (x, t)eiωt dt , f (x, ω) = f (x, t)eiωt dt (2.58)
−∞ −∞

When the representation (2.57) are inserted into (2.56) it is found that the
Fourier transform Ψ (x, ω) satisfies the inhomogeneous Helmholtz wave
equation for each value of ω and k = ω/c

∇2 + k 2 Ψ (x, ω) = −4πf (x, ω)



(2.59)

Equation (2.59) is an elliptic PDE similar to Poisson eqn to which it reduces


for k = 0.
The Green function G(x, x0 ) appropriate to (2.59) satisfies the inhomoge-
neous equation
∇2 + k 2 Gk (x, x0 ) = −4πδ(x − x0 )

(2.60)
If there are no boundary surfaces, the Green function can only depend on
R = x − x0 , and must be spherically symmetric, that is depend only on
R = |R|. This means that in spherical coordinates Gk (R) satisfies

1 d2
(RGk ) + k 2 Gk = 4πδ(R) (2.61)
R dR2

In other words, everywhere except R = 0 the quantity RGk (R) satisfies


the homogeneous equation

d2
(RGk ) + k 2 (RGk ) = 0
dR2
with solution :
RGk (R) = AeikR + Be−ikR
The general solution for the Green function is :
(+) (−)
Gk (R) = AGk (R) + BGk (R) (2.62)

where
(±) e±ikR
Gk (R) = (2.63)
R
2.5 Green Functions for the Wave Equations (***) 37

with A + B = 1 and the correct normalization condition at R → 0


1
lim Gk (R) = (2.64)
kR→0 R
The first term of (2.62) represents a diverging spherical wave propagating
from the origin, while the second term represents a converging spherical
wave.
(+) (−)
To understand the different time behaviors associated with Gk and Gk
we need to construct the corresponding time-dependent Green functions that
satisfy
1 ∂2
 
(±)
∇ − 2 2 Gk (x, t; x0 , t0 ) = −4πδ(x − x0 )δ(t − t0 )
2
(2.65)
c ∂t
0
note that the source term for (2.59) is −4πδ(x − x0 )eiωt
Using the Fourier transforms (2.57) the time-dependent Green functions
become (how?)
Z ∞ ±ikR
1 e
G(±) (R, τ ) = e−iωτ dω (2.66)
2π −∞ R
where τ = t − t0 is the relative time.
The integral actually is a δ function and the Green function becomes:
 
1 R
G(±) (R, τ ) = δ τ ∓ (2.67)
R c
or
|x − x0 |
  
(±) 0 0 1 0
G (x, t; x , t ) = δ t − t∓ (2.68)
|x − x0 | c
? The infinite space Green function is thus a function only of the relative
distance R and the relative time τ between the source and the observation
point.
The Green function G(+) is called the retarded Green function and G(−)
is called the advanced Green function
Particular integrals of the inhomogeneous wave equation (2.56) are
Z Z
Ψ (±) (x, t) = G(±) (x, t; x0 , t0 )f (x0 , t0 )d3 x0 dt0

to either of these maybe added solutions of the homogeneous equation in order


to specify a definite physical problem.
EXAMPLES
|x − x0 |
Z Z   
1 0
Φ(x, t) = δ t − t ∓ ρ(x0 , t0 )d3 x0 dt0
|x − x0 | c
|x − x0 | J(x0 , t0 ) 3 0 0
Z Z   
1 0
A(x, t) = 0
δ t − t ∓ d x dt
|x − x | c c
38 2 Classical Field Theory: Maxwell Equations

2.5.1 Poynting’s Theorem : Conservation of Energy (****)

For a continuous distribution of charge and current, the total rate of doing
work by the fields in a finite volume V is
Z
J · Ed3 x (2.69)
V

This represents a conversion of EM energy into mechanical or thermal energy


and it must be balanced by a corresponding rate of decrease of energy in the
EM field within the volume V .
In order to derive this conservation law explicitly we use the Maxwell eqns
to express (2.69) in other terms. We use Ampère law to eliminate J
Z Z  
3 1 ∂D 3
J · Ed x = cE · (∇ × H) − E · d x (2.70)
V 4π V ∂t

If we use the vector identity,

∇ · (E × H) = H · (∇ × E) − E · (∇ × H)

together with the Faraday’s law, we get


Z  
−1
Z
3 ∂D ∂H 3
J · Ed x = c∇ · (E × H) + E · +B· d x (2.71)
V 4π V ∂t ∂t

The terms with the time derivatives can be interpreted as the time derivatives
of the electrostatic and magnetic energy densities.
If we also remember that the sum/integrals
Z Z
1 1
WE = E · Dd3 x and WB = H · Bd3 x (2.72)
8π 8π
represents the total EM energy (even for time varying fields). Then the total
energy density is denoted by
 
1 1 2 1 2
u= (E · D + B · H) ≡ 0 E + B (2.73)
8π 2 µ0

Then equation (2.71) can be written (how?):


Z Z  
∂u c
− J · Ed3 x = + ∇ · (E × H) d3 x (2.74)
V V ∂t 4π

Since the volume is abritrary, this can be cast into the form of a differ-
ential continuity equation or conservation law
∂u
+ ∇ · S = −J · E (2.75)
∂t
2.5 Green Functions for the Wave Equations (***) 39

The vector S represents the energy flow and is called Poynting vector
c
S= (E × H) ≡ (E × H) (2.76)

• Poynting’s theorem (Conservation of energy) : The physical meaning of
the above relations is that the time rate of change of EM energy within a
certain volume, plus the energy flowing out through the boundary surfaces of
the volume per unit time, is equal to the negative of the total work done by
the fields on the sources within the volume.
• In other words Poynting’s theorem for microscopic field (E, B) is
a statement of conservation of energy of the combined system of
particles and fields.
• If we denote the total energy of the particles within the volume V
as Emech and assume that no particles move out of the volume we have
Z
dEmech
= J · Ed3 x (2.77)
dt V

and the total field energy within V as


Z Z
3 1
E2 + B2 d3 x

Efield = ud x = (2.78)
V 8π V

Poynting’s theorem expresses the conservation of energy for the


combined system as:
I
dE d
= (Emech + Efield ) = − n · Sda (2.79)
dt dt S

2.5.2 Poynting’s Theorem : Conservation of Momentum (***)

The conservation of linear momentum can be similarly considered. The


total EM force on a charged particle is
 v 
F=q E+ ×B (2.80)
c
If we denote as Pmech the sum of all the momenta of all the particles in a
volume V we can write the Newton’s 2nd law (how?),
Z  
dPmech 1
Ftot ≡ = ρE + J × B d3 x (2.81)
dt V c

where (J ≡ ρv) we converted the sum over particles to an integral over charge
& current densities.
We can use Maxwell equations to eliminate ρ and J from (2.81) by using
40 2 Classical Field Theory: Maxwell Equations
 
1 c 1 ∂E
ρ= ∇ · E, J= ∇×B− (2.82)
4π 4π c ∂t

After some manipulations we can show (how?) that the rate of change
of mechanical momentum of eqn (2.81) can be written
Z
dPmech d 1
+ (E × B) d3 x (2.83)
dt dt V 4πc
Z
1
= [E(∇ · E) − E × (∇ × E) + B(∇ · B) − B × (∇ × B)] d3 x
4π V
We can identify the volume integral on the left as the total EM momen-
tum Pfield in the volume V
Z
1
Pfield = (E × B) d3 x (2.84)
4πc V
The integrant can be interpreted as the density of EM momentum
1
g= (E × B) (2.85)
4πc
The EM momentum density g is proportional to the energy-flux density S
with proportionality constant c−2 .

2.6 Maxwell Stress Tensor (***)


In order to establish that (2.84) is a conservation law for momentum, we must
convert the volume integral on the right into surface integral of something that
will be identified as momentum flow.
By defining the Maxwell stress tensor Tab as
 
1 1
Tab = Ea Eb + Ba Bb − δab (E · E + B · B) (2.86)
4π 2
where the a-th component can be written as
X ∂  1

[E(∇ · E) − E × (∇ × E)]a = Ea Eb − δab E · E (2.87)
∂xb 2
b

also
X ∂
(∇ · T)a = Tab (2.88)
∂xb
b

Then (2.84) can be written as:


Z
d X ∂
(Pmech + Pfield )a = Tab d3 x (2.89)
dt V ∂xb
b
2.7 Conservation of Angular Momentum (**) 41

Then via the divergence theorem we get


I X
d
(Pmech + Pfield )a = Tab nb da (2.90)
dt S b

If (2.90) represents a statement of conservation of momentum,


X
Tab nb
b

is the a-th component of the flow per unit area of momentum across the
surface S into the volume V .
In other words it is the force per unit area transmitted across the surface
S and acting on the combined system of particles and fields inside V .

2.7 Conservation of Angular Momentum (**)

The derivation of the EM angular momentum shares the same tactical ap-
proach as that of the linear momentum
Let us define the mechanical angular momentum of the system as

Lmech = r × pmech (2.91)

where pmech is the mechanical momentum density. Then


 
dLmech 1
= r × ρE + J × B , (2.92)
dt c

and substitution of ρ and J from Maxwell’s eqns leads to


 
d 1
Lmech + r × (E × B) (2.93)
dt 4πc
1
= r × [E(∇ · E) − E × (∇ × E) + B(∇ · B) − B × (∇ × B)]

By using the definition of the Maxwell stress tensor (2.86), we can simplify
eqn (2.94) considerably

d
(Lmech + Lfield ) = r × ∇ · T (2.94)
dt
where
Lfield = r × g (2.95)
has the interpretation of being the EM field angular momentum density
In integral form since
42 2 Classical Field Theory: Maxwell Equations

r × ∇ · T = ∇ · (r × T) and ∇ × r = 0 (2.96)

we get  Z  Z
d
Lmech + Lfield d3 x = (r × T) · nda (2.97)
dt V S

The right-hand side of this equation represents the integrated torque density
due to the fields over the boundary surface S.
3
Electromagnetic Waves (****)

3.1 Maxwell Equations


? A basic feature of Maxwell equations for the EM field is the existence of
travelling wave solutions which represent the transport of energy from one
point to another.
? The simplest and most fundamental EM waves are transverse, plane
waves.
In a region of space where there are no free sources (ρ = 0, J = 0),
Maxwell’s equations reduce to a simple form given
1 ∂B
∇·E=0 , ∇×E+ =0
c ∂t
µ ∂E
∇·B=0 , ∇×B− =0 (3.1)
c ∂t
where D and H are given by relations
1
D = E and H = B (3.2)
µ
where  is the electric permittivity and µ the magnetic permeability
which assumed to be independent of the frequency.

3.2 Plane Electromagnetic Waves


Maxwell’s equations can be written as

µ ∂ 2 B µ ∂ 2 E
∇2 B − =0 and ∇2 E − =0 (3.3)
c2 ∂t2 c2 ∂t2
In other words each component of B and E obeys a wave equation of the
form:
44 3 Electromagnetic Waves (****)

1 ∂2u c
∇2 u − =0 where v = √ (3.4)
v 2 ∂t2 µ
is a constant with dimensions of velocity characteristic of the medium.
The wave equation admits admits plane-wave solutions:

u = eik·x−iωt (3.5)
ikn·x−iωt ikn·x−iωt
E(x, t) = Ee and B(x, t) = Be (3.6)

where the relation between the frequency ω and the wave vector k is
ω √ ω  ω 2
k = = µ or k · k = (3.7)
v c v
also the vectors n, E and B are constant in time and space.
If we consider waves propagating in one direction, say x-direction then the
fundamental solution is:

u(x, t) = Aeik(x−vt) + Be−ik(x+vt) (3.8)

which represents waves traveling to the right and to the left with propagation
velocities v which is called phase velocity of the wave.
? From the divergence relations of (3.1) by applying (3.6) we get

n·E =0 and n·B =0 (3.9)

This means that E (or E) and B (or B) are both perpendicular to the
direction of propagation n. Such a wave is called transverse wave.
? The curl equations provide a further restriction
√ 1
B= µ n × E and E = − √ n × B (3.10)
µ

The combination of equations (3.9) and (3.10) suggests that the vectors n, E
and B form an orthonormal set.
Also, if n is real, then (3.10) implies that that E and B have the same
phase.
It is then useful to introduce a set of real mutually orthogonal unit vectors
(1 , 2 , n).
In terms of these unit vectors the field strengths E and B are

E = 1 E0 , B = 2 µE0 (3.11)

or

E = 2 E00 , B = −1 µE00 (3.12)
E0 and E00 are constants, possibly complex.
In other words the most general way to write the electric/magnetic field
vector is:
3.2 Plane Electromagnetic Waves 45

Fig. 3.1.

E= (E0 1 + E00 2 )eikn·x−iωt (3.13)



B= µ(E0 2 − E00 1 )eikn·x−iωt (3.14)

Thus the wave described by (3.6) and (3.11) or (3.12) is a transverse wave
propagating in the direction n.

Or that E and B are oscillating in a plane perpendicular to the wave vector


k, determining the direction of propagation of the wave.

Fig. 3.2.

The energy flux of EM waves is described by the real part of the


complex Poynting vector
1 c 1 c
S= E × H∗ = [ER × HR + EI × HI + i(EI × HR − ER × HI )]
2 4π 2 4π
46 3 Electromagnetic Waves (****)

where E and H are the measured fields at the point where S is evaluated.1
The time averaged flux of energy is:
r
c 
S= |E0 |2 n (3.15)
8π µ

The total time averaged density (and not just the energy density associ-
ated with the electric field component) is:
 
1 1 
u= E · E∗ + B · B∗ = |E0 |2 (3.16)
16π µ 8π

The ratio of the magnitude of (3.15) to (3.16) is the speed of energy flow i.e.

v = c/ µ. 2 (Prove the above relations)

Project:
What will happen if n is not real?
What type of waves you will get?
What will be the form of E?

3.3 Linear and Circular Polarization of EM Waves


The plane wave (3.6) and (3.11) is a wave with its electric field vector always
in the direction 1 . Such a wave is said to be linearly polarized with po-
larization vector 1 . The wave described by (3.12) is linearly polarized with
polarization vector 2 and is linearly independent of the first. The two waves
:

E1 = 1 E1 eik·x−iωt , E2 = 2 E2 eik·x−iωt
with (3.17)
√ k × Ei
Bi = µ , i = 1, 2
k

Can be combined to give the most general homogeneous plane waves prop-
agating in the direction k = kn,

E(x, t) = (1 E1 + 2 E2 ) eik·x−iωt (3.18)


h i
E(x, t) = 1 |E1 | + 2 |E2 |ei(φ2 −φ1 ) eik·x−iωt+iφ1 (3.19)
1
Note : we use the magnetic induction H because although B is the applied
induction, the actual field that carries the energy and momentum in media is H.
2
Note: To prove the above relations use hcos2 xi = 1/2 and since ER = (E+E∗ )/2
we get hE2R i = E · E∗ /2.
3.3 Linear and Circular Polarization of EM Waves 47

Fig. 3.3.

The amplitudes E1 = |E1 |eiφ1 and E2 = |E2 |eiφ2 are complex numbers in
order to allow the possibility of a phase difference between waves of different
polarization.
LINEARLY POLARIZED
If the amplitudes E1 = |E1 |eiφ1 and E2 = |E2 |eiφ2 have the same phase
(3.18) represents a linearly polarized wave with the polarization vector
making an angle θ = tan−1 (<(E2 )/=(E1 )) (which remains constant
p as the
field evolves in space and time) with 1 and magnitude E = E12 + E22 .

ELLIPTICALLY POLARIZED
If E1 and E2 have the different phase the wave (3.18) is elliptically po-
larized and the electric vector rotates around k.

3.3.1 Circular Polarization

• E1 = E2 = E0
• φ1 − φ2 = ±π/2 and the wave becomes

E(x, t) = E0 (1 ± i2 ) eik·x−iωt (3.20)

At a fixed point in space, the fields are such that the electric vector is constant
in magnitude, but sweeps around in a circle at a frequency ω.

Fig. 3.4.
48 3 Electromagnetic Waves (****)

The components of the electric field, obtained by taking the real part of
(3.20)

Ex (x, t) = E0 cos(kz − ωt) , Ey (x, t) = ∓E0 cos(kz − ωt) (3.21)

For the upper sign (1 + i2 ) the rotation is counter-clockwise when the
observer is facing into the oncoming wave. The wave is called left circularly
polarized in optics while in modern physics such a wave is said to have
positive helicity.
For the lower sign (1 − i2 ) the wave is right circularly polarized or it
has negative helicity.

3.3.2 Elliptically Polarized EM Waves

An alternative general expression for E can be given in terms of the complex


orthogonal vectors
1
± = √ (1 ± i2 ) (3.22)
2
with properties

∗± · ∓ = 0 , ∗± · 3 = 0 , ∗± · ± = 1 . (3.23)

Then the general representation of the electric vector

E(x, t) = (E+ + + E− − ) eik·x−iωt (3.24)

where E− and E+ are complex amplitudes


If E− and E+ have different amplitudes but the same phase eqn (3.24)
represents an elliptically polarized wave with principle axes of the ellipse
in the directions of 1 and 2 .
The ratio of the semimajor to semiminor axis is |(1 + r)/(1 − r)|, where
E− /E+ = r.
The ratio of the semimajor to semiminor axis is |(1 + r)/(1 − r)|, where
E− /E+ = r.
If the amplitudes have a phase difference between them E− /E+ = reiα ,
then the ellipse traced out by the E vector has its axes rotated by an angle
α/2.
Note : For r = ±1 we get back to a linearly polarized wave.

3.4 Stokes Parameters


The polarization content of an EM wave is known if it can be written in the
form of either (3.18) or (3.24) with known coefficients (E1 , E2 ) or (E− , E+ ) .
In practice, the converse problem arises i.e. given a wave of the form (3.6),
how can we determine from observations on the beam the state of polarization?
3.4 Stokes Parameters 49

Fig. 3.5. The figure shows the general case of elliptical polarization and the ellipses
traced out by both E and B at a given point in space.

Fig. 3.6. The figure shows the linear, circular and elliptical polarization

A useful tool for this are the four Stokes parameters. These are
quadratic in the field strength and can be determined through intensity mea-
surements only. Their measurements determines completely the state of po-
larization of the wave.
For a wave propagating in the z-direction the scalar products

1 · E , 2 · E , ∗+ · E , ∗− · E (3.25)

are the amplitudes of radiation respectively, with linear polarization in the


x-direction, linear polarization in the y-direction, positive helicity and
negative helicity.
50 3 Electromagnetic Waves (****)

The squares of these amplitudes give a measure of the intensity of each


type of polarization.
The phase information can be taken by using cross products
In terms of the linear polarization bases (1 , 2 ), the Stokes parameters
are:

s0 = |1 · E|2 + |2 · E|2 = a21 + a22


s1 = |1 · E|2 − |2 · E|2 = a21 − a22
s2 = 2< [(1 · E)∗ (1 · E)] = 2a1 a2 cos(δ1 − δ2 ) (3.26)

s3 = 2= [(1 · E) (1 · E)] = 2a1 a2 sin(δ1 − δ2 )

where we defined the coefficients of (3.18) or (3.24) as magnitude times a


phase factor:

E1 = a1 eiδ1 , E2 = a2 eiδ2 , E+ = a+ eiδ+ , E− = a− eiδ− (3.27)

Here s0 and s1 contain information regarding the amplitudes of linear


polarization, whereas s2 and s3 say something about the phases. Knowing
these parameters (e.g by passing a wave through perpendicular polarization
filters) is sufficient for us to determine the amplitudes and relative phases of
the field components.
In terms of the linear polarization bases (+ , − ), the Stokes parameters
are:

s0 = |∗+ · E|2 + |∗− · E|2 = a2+ + a2−


s1 = 2< (∗+ · E)∗ (∗− · E) = 2a+ a− cos(δ− − δ+ )
 
(3.28)
s2 = 2= (∗+ · E)∗ (∗− · E) = 2a+ a− sin(δ− − δ+ )
 

s3 = |∗+ · E|2 − |∗− · E|2 = a2+ − a2−

Notice an interesting rearrangement of roles of the Stokes parameters with


respect to the two bases.
The four Stokes parameters are not independent since they depend on only
3 quantities a1 , a2 and δ1 − δ2 . They satisfy the relation

s20 = s21 + s22 + s23 . (3.29)

3.5 Reflection & Refraction of EM Waves


The reflection and refraction of light at a plane surface between two media of
different dielectric properties are familiar phenomena.
The various aspects of the phenomena divide themselves into two classes
• Kinematic properties:
3.5 Reflection & Refraction of EM Waves 51

– Angle of reflection = angle of incidence


sin i n0
– Snell’s law: sin r = n where i, r are the angles of incidence and
refraction, while n, n0 are the corresponding indices of refraction.
• Dynamic properties:
– Intensities of reflected and refracted radiation
– Phase changes and polarization
? The kinematic properties follow from the wave nature of the phenom-
ena and the need to satisfy certain boundary conditions (BC). But not on the
detailed nature of the waves or the boundary conditions.
? The dynamic properties depend entirely on the specific nature of
the EM fields and their boundary conditions.

Fig. 3.7. Incident wave k strikes plane interface between different media, giving
rise to a reflected wave k00 and a refracted wave k0 . The media below and above the
plane z = 0 have permeabilities and dielectric constants
√ µ,  and µ’, ’ respectively.

The indices of refraction are n = µ and n0 = µ0 0 .

According to eqn (3.18) the 3 waves are:


INCIDENT
√ k×E
E = E0 eik·x−iωt , B= µ (3.30)
k
REFRACTED
0 k 0 × E0
E0 = E00 eik ·x−iωt , B0 =
p
µ0 0 (3.31)
k0
REFLECTED
00 k00 × E00
E00 = E000 eik ·x−iωt
B0 =
p
, µ0 0 (3.32)
k 00
52 3 Electromagnetic Waves (****)

The wave numbers have magnitudes:


ω√ ωp 0 0
|k| = |k00 | = k = µ , |k0 | = k 0 = µ (3.33)
c c

AT the boundary z = 0 the BC must be satisfied at all points on the


plane at all times, i.e. the spatial & time variation of all fields must be the
same at z = 0.
Thus the phase factors must be equal at z = 0

(k · x)z=0 = (k00 · x)z=0 = (k0 · x)z=0 (3.34)

independent of the nature of the boundary conditions.


? Eqn (3.34) contains the kinematic aspects of reflection and refraction.
Note that all 3 wave vectors must lie in a plane. From the previous figure
we get
k sin i = k 00 sin r0 = k 0 sin r (3.35)
Since k = k 00 , we find that i = r0 ; the angle of incidence equals the angle of
reflection.
Snell’s law is: s
sin i k0 µ0 0 n0
= = = (3.36)
sin r k µ n

The dynamic properties are contained in the boundary conditions :


• normal components of D = E and B are continuous
• tangential components of E and H = [c/(ωµ)]k × E are continuous
In terms of fields (3.30)-(3.32) these boundary conditions at z = 0 are:

[ (E0 + E000 ) − 0 E00 ] · n = 0


[k × E0 + k00 × E000 − k0 × E00 ] · n = 0
(E0 + E000 − E00 ) × n = 0 (3.37)
 
1 00 00 1 0 0
(k × E0 + k × E0 ) − 0 (k × E0 ) × n = 0
µ µ

Two separate situations, the incident plane wave is linearly polarized :


• The polarization vector is perpendicular to the plane of incidence (the
plane defined by k and n ).
• The polarization vector is parallel to the plane of incidence.
• The case of arbitrary elliptic polarization can be obtained by appropriate
linear combinations of the two results.
3.5 Reflection & Refraction of EM Waves 53

3.5.1 E : Perpendicular to the plane of incidence

• Since the E-fields are parallel to the surface the 1st BC of (3.38) yields
nothing
• The 3rd and 4th of of (3.38) give (how?):

E0 + E000 − E00 = 0
s
0 0
r

(E0 − E000 ) cos i − E cos r = 0 (3.38)
µ µ0 0

• The 2nd, using Snell’s law, duplicates the 3rd.


(prove all the above statements)

Fig. 3.8. Reflection and refraction with polarization perpendicular to the plane of
incidence. All the E-fields shown directed away from the viewer.

The relative amplitudes of the refracted and reflected waves can be found
from (3.38)

E00
 
2n cos i 2 2 sin r cos i
= = =
1 + µµ0 tan i
p
E0 n cos i + µµ0 n02 − n2 sin2 i tan r
sin(i + r) µ=µ0
p
E000 n cos i − µµ0 n02 − n2 sin2 i 1 − µµ0 tan i  
tan r sin(r − i)
= = = (3.39)
1 + µµ0 tan i
p
E0 n cos i + µ0 n02 − n2 sin2 i tan r
sin(i + r) µ=µ0
µ
p
Note that n02 − n2 sin2 i = n0 cos r but Snell’s law has been used to express
it in terms of the angle of incidence.
For optical frequencies it is usually permitted to put µ = µ0 .
54 3 Electromagnetic Waves (****)

3.5.2 E : Parallel to the plane of incidence

Boundary conditions involved:


• normal E : 1st eqn in (3.38)
• tangential E : 3rd eqn in (3.38)
• tangential B : 4th eqn in (3.38)

The last two demand that


(E0 − E000 ) cos i − E00 cos r = 0
s
0 0
r
 00
(E0 + E0 ) − E =0 (3.40)
µ µ0 0

Fig. 3.9. Reflection and refraction with polarization parallel to the plane of inci-
dence.

The condition that normal E is continuous, plus Snell’s law, merely dubli-
cates the 2nd of the previous equations.
The relative amplitudes of refracted and reflected fields are therefore
(how?)
E00 2nn0 cos i
= p
E0 µ 02
0 n cos i + n n02 − n2 sin2 i
µ
0
2 nn0
 
2 sin r cos i
= =
1 + 0 tan i
tan r
sin(i + r) cos(i − r) µ=µ0
p
µ 02 02 2 2
E000 µ0 n cos i − n n − n sin i
= p
E0 µ 02 02 2 2
µ0 n cos i + n n − n sin i
1 − 0 tan i  
tan r tan(i − r)
= = (3.41)
1 + 0 tan i
tan r
tan(i + r) µ=µ0
3.5 Reflection & Refraction of EM Waves 55

3.5.3 Normal incidence i = 0

For normal incidence i = 0 both (3.39) and (3.41) reduce to (how?)

E00 2 2n
= q 0 → 0
E0 1 + µ n +n
µ0 
q 0
E000 −1 + µ µ0  n0 − n
= q 0 → 0 (3.42)
E0 1 + µ n +n
µ0 

EXERCISES:
What are the conditions for:
• Total reflection
• Total transmision

Fig. 3.10. The conversion of an incident circularly polarized wave into an elliptically
polarized wave. The ratio of semimajor to semiminor transmitted amplitude for this
example in 1.3
4
Simple Radiating Systems (****)

4.1 Fields and Radiation of a Localized Oscillating


Source
For a system of charges and currents varying in time we can make a Fourier
analysis of the time dependence and handle each Fourier component sep-
arately. The potentials, and fields from a localized system of charges and
currents which vary sinusoidally in time become:
ρ(x, t) = ρ(x)eiωt , J(x, t) = J(x)eiωt (4.1)
The physical quantities will be taken from the real part of such expressions
while the EM potentials and fields assumed to have the same time dependence.
In the Lorentz gauge the solution (provided no boundary surfaces are present)
for the vector potential A(x, t) is
J(x0 , t0 ) |x − x0 |
Z Z  
1 3 0 0
A(x, t) = d x δ t + − t dt0 (4.2)
c |x − x0 | c
Using the sinusoidal dependence (4.1) we get (k = ω/c)
J(x0 ) ik|x−x0 |
Z
1
A(x) = d3 x0 e (4.3)
c |x − x0 |
Given a current distribution J(x0 ) the fields can be determined by calcu-
lating the integral (4.3). The magnetic induction will be given by
B=∇×A (4.4)
while outside the source the electric field is
i
E= ∇×B (4.5)
k
If the sources dimensions are of order d and the wavelength is λ = 2πc/ω
and if d  λ then there are 3 spatial regions of inderest
58 4 Simple Radiating Systems (****)

• The near (static) zone : drλ


• The intermediate (induction) zone : dr∼λ
• The far (radiation) zone : dλr
The fields have very different properties in the different zones.
• In the near zone they behave as static and depend on the properties of
the source.
• In the far zone the fields are transverse to the radius vector and fall as
1/r (typical for radiation fields).

4.2 The Near Zone


The general solution is:
J(x0 ) ik|x−x0 |
Z
1
A(x) = d3 x0 e (4.6)
c |x − x0 |

For the near zone where r  λ (or kr  1) the exponential in (4.6) can
be replaced by unity.
Then by using:
∞ X
l `
1 X 1 r<
= 4π Y ∗ (θ0 , φ0 )Y`m (θ, φ)
`+1 `m
(4.7)
|x − x0 | 2` + 1 r>
`=0 m=−`

The vector potential gets the form


Z
1 X 4π Y`m (θ, φ)
lim A(x) = J(x0 )r0` Y`m

(θ0 , φ0 )d3 x0 (4.8)
kr→0 c 2` + 1 r`+1
`,m

This shows that the near fields are quasi-stationary, oscillating harmonically
as e−iωt , but otherwise static in character.

4.2.1 The Far Zone

In the far zone (kr  1) the exponential in (4.3) oscillates rapidly and deter-
mines the behavior of the vector potential.
In this region is sufficient to approximate

|x − x0 | ≈ r − n · x0 (4.9)

where n is the unit vector in the direction of x.


If only the leading term in kr is desired, the inverse distance can be re-
placed by r and the vector potential is

eikr
Z
0
lim A(x) = J(x0 )e−ikn·x d3 x0 (4.10)
kr→∞ cr
4.2 The Near Zone 59

I.e. in the far zone the vector potential behaves as an outgoing spherical wave.
It is easy to show that the fields calculated from (4.4) and (4.5) are trans-
verse to the radius vector (how?) and fall like 1/r. Thus they correspond to
radiation fields.
If the source size is small compared to a wavelength one can expand in
powers of k:

eikr X (−ik)n
Z
lim A(x) = J(x0 )(n · x0 )n d3 x0 (4.11)
kr→∞ cr n n!

The magnitude of the n-th term is given by


Z
1
J(x0 )(kn · x0 )n d3 x0 (4.12)
n!

Since |x0 | ≈ d and kd  1 by assumption, the successive terms in the expan-


sion of A fall off rapidly with n.
Consequently the radiation emitted from the source will come mainly from
the first nonvanishing term in the expansion (4.11).

4.2.2 The Intermediate Zone

In the intermediate zone neither kr  1 or kr  1 can be used; all power of


kr must be retained. For points outside the source, eqn (4.3) becomes
Z
4πik X (1)
A(x) = h` (kr)Ym` (θ, φ) J(x0 )j` (kr0 )Y`m

(θ0 , φ0 )d3 x0 (4.13)
c
`,m

If the source dimensions are small compared to a wavelength, j` (kr0 ) can be


approximate as follows: 1

xl x3
 
j` (x) → 1− + ... (4.14)
(2` + 1)!! 2(2` + 3)

Then the vector potential is of the form (4.8) but with the replacement

1 eikr 
1 + a1 (ikr) + a2 (ikr)2 + · · · al (ikr)l

→ (4.15)
r`+1 r `+1

The coefficients ai come from explicit expressions for the spherical Hankel
functions.
1
[(2` + 1)!! = (2` + 1)(2` − 1)(2` − 3) · · · 5 · 3 · 1]
60 4 Simple Radiating Systems (****)

4.2.3 Electric Monopole Fields

For a source varying with the time the analogue of (4.2) for a scalar potential
is
0 0
|x − x0 |
Z Z  
3 0 0 ρ(x , t ) 0
Φ(x, t) = d x dt δ t + −t (4.16)
|x − x0 | c
The electric monopole contribution is obtained by replacing in the integral
|x − x0 | → |x| ≡ r. The result is

q(t0 = t − r/c)
Φmonopole (x, t) =
r
where q(t) is the total charge of the source. Since charge is conserved and a
localized source has not charge flowing into or away from it, the total charge
q is independent of time.
? Thus the electric monopole part of the potential (and fields) of
localized source is of necessity static.
? The fields with harmonic time dependence e−iωt with ω 6= 0 have no
monopole terms.

4.3 Electric Dipole Fields and Radiation


If only the first term in (4.11) is kept, the potential is

eikr
Z
A(x) = J(x0 )d3 x0 (4.17)
cr

Examination of (4.13) and (4.15) shows that (4.17) is the ` = 0 part of the
series and that it is valid everywhere outside the source, not just in the far
zone. Integrating by part we get
Z Z Z
Jd x = − x (∇ · J)d x = −iω x0 ρ(x0 )d3 x0
3 0 0 0 3 0
(4.18)

since from the continuity equation

∇ · J = iωρ (4.19)

Thus the vector potenial is:

eikr
A(x) = −ik p (4.20)
r
Z
p(x) = x0 ρ(x0 )d3 x0 (4.21)

is the electric dipole moment, as defined in electrostatics


4.3 Electric Dipole Fields and Radiation 61

The electric dipole fields from (4.4) and (4.5) are (how?)
  ikr
2 1 e
B = k (n × p) 1 − (4.22)
ikr r
ikr
 
2 e 1 ik
E = k (n × p) × n + [3n(n · p) − p] − 2 eikr (4.23)
r r3 r

Note that the magnetic induction is transverse to the radius vector at all
distances, but the electric field has components parallel and perpendicular to
n.
In the radiation zone (kr  1) the fields take the limiting form

eikr
B = k 2 (n × p) , E=B×n (4.24)
r
In the near zone (kr  1) the fields approach
1 1
B = ik(n × p) , E = [3n(n · p) − p] (4.25)
r2 r3

? The electric field apart from its oscillation in time is just the static
electric dipole field seen earlier
? The magnetic induction is smaller than the electric field by a factor
(kr) in the region where kr  1.
? Thus the fields in the near zone are dominantly electric in
nature.
? The magnetic induction vanishes, in the static limit k → 0.
? The following identities might be useful for future derivations

[p − (p · n)n] · (n × p) = p · (n × p) − (p · (n))n · (n × p) = 0 (4.26)


(n × p) × n = (n · n) p − (n · p) n = p − (p · n) n (4.27)

? If we let p point in the z-direction, then p = pez also

n = sin θ cos φex + sin θ sin φey + cos θez


p · n = p cos θ
p − (p · n) n = −p sin θ (cos θ cos φex + cos θ sin φey − sin θz) = −p sin θeθ
n × p = −p sin θ (− sin φez + cos φey ) = −p sin θeφ

and we get (prove it)

eikr eikr
B = −k 2 p sin θ eφ and E = −k 2 p sin θ eθ (4.28)
r r
62 4 Simple Radiating Systems (****)

4.3.1 Electric Dipole : Power Radiated

The Poynting vector is


c c c c
S= E×B= (B × n) × B = [(B · B) n − (B · n) B] = |B|2 n
8π 8π 8π 8π
(4.29)
the blue term is vanishing because B · n = 0.
The time-averaged power radiated per unit solid angle by the oscillating dipole
moment p is:
dP c  2
< r n · (E × B∗ )

= (4.30)
dΩ 8π
where E and B are given by (4.24). Thus we find

dP c 4 2
= k |(n × p) × n| (4.31)
dΩ 8π
If all the components of p have the same phase, the angular distribution is a
typical dipole pattern (where θ is measured from the direction of p )

dP c 4 2 2
= k |p| sin θ (4.32)
dΩ 8π

The total power radiated, independent of the relative phases of the com-
ponents of p, is
1 2
P = ck 4 |p| (4.33)
3

4.4 Examples
4.4.1 Example I

For p = pez we can substitute (4.28) into (4.29) and we get (how?)

c 4 2 sin2 θ 2ikr
S= k p e n (4.34)
8π 2 r2
Since the energy flux oscillates in time after averaging over a complete wave
cycle we get (hcos2 (kr)i = 1/2)

ck 4 p2 sin2 θ
hSi = n (4.35)
16π 2 r2
From the angular dependence of this expression we see that:
• most of the energy is emitted near the equatorial plane (θ = π/2), and
• none of the energy propagates along the z-axis (θ = 0 or θ = π).
4.4 Examples 63

Fig. 4.1. Power distribution.

The total power radiated is obtained by integrating S over any closed


surface S. The averaged power is then
I
hP i = hSi · da (4.36)
S

we can choose S to be a sphere of constant r then da = r2 sin θdθdφn


Z π
ck 4 p2 c 4 2
hP i = 2π sin3 θdθ = k p (4.37)
16π 0 6π

4.4.2 Example II

A simple example of an electric dipole radiator is a centerfed linear antenna


whose length d is small compared to a wavelength. The antenna is assumed to
be oriented along the z-axis, with the narrow gab at the center for purposes
of excitation. The current is in the same direction in each half of the antenna,
with a value I0 at the gap and falling linearly to zero at the ends:
 
2|z| −iωt
I(z)e−iωt = I0 1 − e (4.38)
d
From the continuity eqn (4.19) the linear charge density ρ0 (charge per unit
length) along each arm is constant with value

2iI0
ρ0 (z) = ± (4.39)
ωd
64 4 Simple Radiating Systems (****)

Fig. 4.2. Electric field lines of an oscillating dipole p(t) = p0 cos(ωt)ez at selected
moments of time for ω = 1. Witness the transition between near-zone behavior and
wave-zone behavior which occurs near r = λ = 2πc/ω ≡ 1.

Fig. 4.3. Electric field lines of an rotating dipole p(t) = p0 [cos(ωt)ez + sin(ωt)ey ]
at selected moments of time, for ω = 1. Witness the transition between near-zone
behavior and wave-zone behavior which occurs near r = λ = 2πc/ω ≡ 1.
4.5 Magnetic Dipole & Electric Quadrupole Fields 65

Fig. 4.4.

The dipole moment (4.21) is parallel to the z-axis with magnitude


Z d/2
iI0 d
p= zρ0 (z)dz = (4.40)
−d/2 2ω

The angular distribution of the power radiated is


dP I2
= 0 (kd)2 sin2 θ (4.41)
dΩ 32πc
while the total power radiated is
I02
P = (kd)2 (4.42)
12πc
We see that for a fixed input current the power radiated increases as the
square of the frequency, at least in the long-wavelength domain (kd  1).
The coefficient of I02 /2 has dimensions of a resistance and is called the
radiation resistance Rrad of the antenna.

4.5 Magnetic Dipole & Electric Quadrupole Fields


Electric-dipole radiation corresponds to the leading-order approximation of
the EM field in an expansion in powers of k or v/c, where v is the typical
internal velocity of the field. In some case, however, the dipole moment p
either vanishes or does not depend on time, and the leading term is actually
zero. In such cases or when higher accuracy is required, we need to compute
the next term in the expansion.
We will see that at the next-to-leading-order, the wave-zone fields depend
on the dipole moment vector p, the magnetic moment vector term M
and the electric quadrupole moment tensor Qab .
66 4 Simple Radiating Systems (****)

The next term in the expansion (4.11) and (4.15) leads to a vector potential

eikr 1
 Z
A(x) = − ik J(x0 ) (n · x0 ) d3 x0 (4.43)
cr r

This vector potential can be written as the sum of two terms, one of which
gives a transverse magnetic induction and the other who gives a transverse
electric field.
This can be achieved by writing the integrand as the sum of a part sym-
metric in J and x0 and a part that is antisymmetric.
1 1 1 0
(n · x0 )J = [(n · x0 )J + (n · J)x0 ] + (x × J) × n (4.44)
c 2c 2c
The 1st symmetric term will be shown to be related to the electric
quadrupole moment density.
The 2nd antisymmetric part is the so called magnetization due to
the current J:
1
M= (x × J) (4.45)
2c

4.5.1 Magnetic Dipole

Considering only the magnetization term, we have the vector potential

eikr
 
1
A(x) = ik(n × m) 1− (4.46)
r ikr

where m is the magnetic dipole moment


Z Z
1
m = Md3 x = (x × J) d3 x (4.47)
2c
The fields can be determined by noting that the vector potential (4.46) is pro-
portional to the magnetic induction (4.23) for an electric dipole. This means
that the magnetic induction will be equal to the electric field for the electric
dipole, with substitution p → m

eikr
 
2 1 ik
B = k (n × m) × n + [3n(n · m] − m] − 2 eikr (4.48)
r r3 r

Similarly, the electric field for a magnetic dipole source is the negative of
the magnetic field for an electric dipole:

eikr
 
1
E = −k 2 (n × m) 1− (4.49)
r ikr
4.5 Magnetic Dipole & Electric Quadrupole Fields 67

• All the arguments concerning the behavior of the fields in the near and
far zones are the same as for the electric dipole source, with the interchange
E → B, B → −E, p → m.

• Similarly the radiation pattern and total power radiated are the same
for the two kinds of dipole. The only difference in the radiation fields is the
polarization.

• For an electric dipole the electric vector lies in the plane defined by n
and p, while for a magnetic dipole it is perpendicular to the plane defined by
n and m.

• The total power radiated, can be estimated (how?) by substituting


p → m in (4.33)
c 2
PM = k 4 |m| (4.50)
3
Thus  2  
PM m0 vc 2
= ≈ (4.51)
PE p0 c c
2
Thus the magnetic dipole radiation power is by a factor vcc  1 smaller
compared with electric dipole radiation power.
Note that: p ∼ (ρr)(r3 ) ∼ ρr4 while m ∼ (Jr)(r3 ) ∼ Jr4 ∼ ρvc r4 .

4.5.2 Electric Quadrupole Fields

The integral of the symmetric term in (4.44) can be transformed by an inte-


gration by parts and some rearrangement (how?):
Z Z
1 ik
[(n · x0 ) J + (n · J) x0 ] d3 x0 = − x0 (n · x0 ) ρ(x0 )d3 x0 (4.52)
2c 2
where the continuity eqn (4.19) has been used to replace ∇·J by iωρ. Since the
integral involves second moments of the charge density, this symmetry
part corresponds to an electric quadrupole source. The vector potential is

k 2 eikr
 Z
1
A(x) = − 1− x0 (n · x0 )ρ(x0 )d3 x0 (4.53)
2 r ikr
Since the complete fields are complicated to write down, we will study fields
in the radiation zone. Then it is easy to see that

B = ik(n × A) and E = ik(n × A) × n (4.54)

Consequently the magnetic induction is

ik 3 eikr
Z
B=− (n × x0 )(n · x0 )ρ(x0 )d3 x0 (4.55)
2 r
68 4 Simple Radiating Systems (****)

Using the definition (??) for the quadrupole moment tensor


Z
Qab = (3xa xb − r2 δab )ρ(x)d3 x (4.56)

the integral (4.55) can be written (how?)


Z
1
n × x0 (n · x0 )ρ(x0 )d3 x0 = n × Q(n) (4.57)
3
The vector Q(n) is defined as having components
X
Qa = Qab nb (4.58)
b

The magnetic induction will be written as:

ik 3 eikr
B=− n × Q(n) (4.59)
6 r
The time-averaged power radiated per unit solid angle
dP c 2
= k 6 |[n × Q(n)] × n| (4.60)
dΩ 288π
and the direction of the radiated electric field is given by the vector inside the
absolute value.
The general angular distribution is complicated. But the total power ra-
diated can be calculated in a straightforward way. We can write the angular
dependence as
2
|[n × Q(n)] × n| = Q∗ · Q − |n · Q|2
X X
= A∗ab Qac nb nc − Q∗ab Qcd na nb nc nd (4.61)
a,b,c a,b,c,d

The necessary angular integrals over products of the rectangular components


of n are:
Z Z

nb nc dΩ = δbc , na nb nc dΩ = 0 (4.62)
3
Z

na nb nc nd dΩ = (δab δcd + δac δbd + δad δbc )
15
Then
 
Z
2 4π X 4π X X X
|[n × Q(n)] × n| dΩ = |Qab |2 −  Q∗aa Qcc + 2 |Qab |2 
3 13 a c
a,b a,b
(4.63)
4.5 Magnetic Dipole & Electric Quadrupole Fields 69

As we have mentioned earlier the trace of Qab is zero and thus the total
power radiated by a quadrupole source is
c 6X
P = k |Qab |2 (4.64)
360
a,b

Notice that the radiated power varies as the 6th power of the frequency for
fixed quadrupole moments, compared to the 4th power for dipole radiation.

In orders of magnitude Qab ∼ k 6 (ρr2 )(r3 ) ∼ k 6 ρr5 then


PE−Q  v 2
∼ (4.65)
PE−D c
For slowly-moving distributions, the power emitted in electric-quadrupole ra-
diation is smaller than the power emitted in electric-dipole radiation by a
factor of order (v/c)2  1.

4.5.3 Example : Electric Quadrupole Fields

An oscillating spheroidal distribution of charge is a simple example of a radi-


ating quadrupole source.
The off-diagonal elements of Qab vanish (why?).
The diagonal terms may written
1 1
Q11 = Q22 = − Q33 = − Q0 (4.66)
2 2
Then the angular distribution of radiated power is

dP ck 6
= Q0 sin2 θ cos2 θ (4.67)
dΩ 128π
The total power radiated by this quadrupole is:
c
P = k 6 Q0 (4.68)
128π

4.5.4 Example : Pulsar spin-down

As an application of the magnetic-dipole radiation we consider the oblique-


rotator model of a pulsar, a rotating neutron star that emits pulses of EM
radiation at regular intervals. The pulsars (e.g. Crab) is observed to spin down
and if we associate the spin frequency with the pulse frequency then we can
interpret it in terms of pulsar losing rotational energy
1 2
Erot = IΩ (4.69)
2
70 4 Simple Radiating Systems (****)

Fig. 4.5. Quadrupole radiation pattern

where I is the star’s moment of inertia and Ω its angular velocity. If P


denotes the pulses period, then Ω = 2π/P . If we model the neutron star as a
solid sphere then I = 25 M R2 . The loss of rotational energy translates into a
decrease of Ω

Ėrot = IΩ Ω̇ = −(2π)2 I 3 (4.70)
P
For the Crab pulsar the observed values are Ṗ ≈ 4 × 10−13 s/s and P ≈ 0.03s.
If we assume M = 1.4M and R = 12km then the rate of loss of rotational
energy is
Ėrot ≈ −7 × 1031 J/s (4.71)
This is comparable to the energy required to power the Crab nebula. The en-
ergetics of the Crab can therefore be explained by the pulsar losing rotational
energy.
4.5 Magnetic Dipole & Electric Quadrupole Fields 71

The energy carried away by the radiation will then come to the expense of
the star’s rotational energy. For this we need the field’s orientation to differ
from star’s rotational axis (oblique rotator model).
If the NS maintains a magnetic dipole moment
m(t) = m0 (sin α cos Ωtex + sin α sin Ωtey + cos αez ) (4.72)
Because the NS is located within the near zone
1
B(t) = 3 [3(m · n)n − m] (4.73)
R

The field is maximum at the magnetic pole, where n is aligned with m


2m0
Bmax = 3 (4.74)
R
The time-changing magnetic moment produces a magnetic-dipole radiation
that takes energy away from the star at a rate given by eqn (4.50)
1
Ėrad = |m̈|2 (4.75)
3c3
and by substitution of (4.72) we get
1 1
Ėrad = (m0 Ω 2 sin α)2 = ... = (Bmax Ω 2 R3 sin α)2 (4.76)
3c3 12c3
and if we set
Ėrad = Ėrot
we get
Bmax sin α ≈ 5 × 1012 Gauss
which is large but not unreasonable. A main sequence star typically supports
a magnetic field of 103 G if the field is frozen in the star during the collapse
the magnetic flux 4πR2 B is conserved. Thus because R decrease by a factor
105 then B increases by a factor 1010 .
5
Radiation by Moving Charges

5.1 Liénard - Wiechert Potentials (**)


• The Liénard-Wiechert potential describes the electromagnetic
effect of a moving charge.
• Built directly from Maxwell’s equations, this potential describes the com-
plete, relativistically correct, time-varying electromagnetic field for a point-
charge in arbitrary motion.
• These classical equations harmonize with the 20th century development of
special relativity, but are not corrected for quantum-mechanical effects.
• Electromagnetic radiation in the form of waves are a natural result of the
solutions to these equations.
• These equations were developed in part by Emil Wiechert around 1898
and continued into the early 1900s.
We will study potentials and fields produced by a point charge, for which
a trajectory x0 (t0 ) has been defined a priori.
It is obvious that when a charge q is radiating is giving away momentum
and energy, and possibly angular momentum and this emission affects the
trajectory. This will be studied later. For the moment, we assume that the
particle is moving with a velocity much smaller than c.
The density of the moving charge is given by

ρ(x0 , t0 ) = qδ(x0 − x0 [t0 ]) (5.1)

and since in general the current density J is ρv, we also have


dx0
J(x0 , t0 ) = qvδ(x0 − x0 [t0 ]) , where v(t0 ) = (5.2)
dt0
In the Lorentz gauge ( ∇ · A + (1/c)∂Φ/∂t = 0) the potential satisfy the wave
equations (2.44) and (2.45) whose solutions are the retarded functions
74 5 Radiation by Moving Charges

ρ(x0 , t − |x − x0 |/c) 3 0
Z
Φ(x, t) = d x (5.3)
|x − x0 |

J(x0 , t − |x − x0 |/c) 3 0
Z
1
A(x, t) = d x (5.4)
c |x − x0 |
It is not difficult to see that these retarded potentials take into account the
finite propagation speed of the EM disturbances since an effect measured at
x and t was produced at the position of the source at time

|x − x0 (t̃)|
t̃ = t − (5.5)
c
Thus, using our expressions for ρ and J from eqns (5.1) and (5.2) and putting
β ≡ v/c,
δ(x0 − x0 [t − |x − x0 |/c]) 3 0
Z
Φ(x, t) = q d x (5.6)
|x − x0 |

β(t − |x − x0 |/c)δ(x0 − x0 [t − |x − x0 |/c]) 3 0


Z
A(x, t) = q d x (5.7)
|x − x0 |

Note that for a given space-time point (x, t), there exists only one point
on the whole trajectory, the retarded coordinate x̃ coresponding to the re-
tarded time t̃ defined in (5.5) which produces a contribution

x̃ = x0 (t̃) = x0 (t − |x − x0 |/c) (5.8)


Let us also define the vector

R(t0 ) = x − x0 (t0 ) (5.9)

in the direction n ≡ R/R. Then

δ(x0 − x0 [t − R(t0 )/c]) 3 0


Z
Φ(x, t) = q d x (5.10)
R(t0 )
β(t − R(t0 )/c)δ(x0 − x0 [t − R(t0 )/c]) 3 0
Z
A(x, t) = q d x (5.11)
R(t0 )

Because the integration variable x0 appears in R(t0 ) we transform it by intro-


ducing a new parameter r∗ , where

x∗ = x0 − x0 [t − R(t0 )/c] (5.12)

The volume elements d3 x∗ and d3 x0 are related by the Jacobian transfor-


mation
d3 x∗ = Jd3 x0 , where J ≡ [1 − n(t0 ) · β(t0 )] (5.13)
5.1 Liénard - Wiechert Potentials (**) 75

Fig. 5.1.

is the Jacobian (how?).


With the new integration variable, the integrals for the potential transform
to
δ(x∗ ) d3 x∗
Z
Φ(x, t) = q ∗
(5.14)
|x − x − x0 (t̃)|(1 − n · β)
and
β(t̃) δ(x∗ ) d3 x∗
Z
A(x, t) = q (5.15)
|x − x∗ − x0 (t̃)|(1 − n · β)
which can be evaluated trivially, since the argument of the Dirac delta function
restricts x∗ to a single value
   
q q
Φ(x, t) = = (5.16)
(1 − n · β)|x − x̃| t̃ (1 − n · β)R t̃
   
qβ qβ
A(x, t) = = (5.17)
(1 − n · β)|x − x̃| t̃ (1 − n · β)R t̃
   
q q
Φ(x, t) = = (5.18)
(1 − n · β)|x − x̃| t̃ (1 − n · β)R t̃
   
qβ qβ
A(x, t) = = (5.19)
(1 − n · β)|x − x̃| t̃ (1 − n · β)R t̃
These are the Liénard - Wiechert potentials.
It is worth noticing the presence of the term (1−n·β), which clearly arises
from the fact that the velocity of the EM waves is finite, so the retardation
effects must be taken into account in determining the fields.
76 5 Radiation by Moving Charges

5.1.1 Special Note about the shrinkage factor (1 − n · β)

Consider a thin cylinder moving along the x-axis with velocity v.


To calculate the field at x when the ends of the cylinder are at (x1 , x2 ),
we need to know the location of the retarded points x̃1 and x̃2

Fig. 5.2.

x1 − x̃1 v x2 − x̃2 v
= and = (5.20)
x − x̃1 c x − x̃2 c

by setting L̃ ≡ x̃2 − x̃1 and L ≡ x2 − x1 and subtracting we get


v L
L̃ − L = L̃ → L̃ = (5.21)
c 1 − v/c

That is, the effective length L̃ and the natural length L differ by the
factor (1 − x · β)−1 = (1 − v/c)−1 because the source is moving relative to
the observer and its velocity must be taken into account when calculating the
retardation effects.

5.2 Liénard - Wiechert potentials : radiation fields (****)


The next step after calculating the potentials is to calculate the fields via the
relations
1 ∂A
B = ∇ × A and E = − − ∇Φ (5.22)
c ∂t
and we write the Liénard - Wiechert potentials in the equivalent form
δ(t0 − t − R(t0 )/c) 0
Z
Φ(x, t) = q dt (5.23)
R(t0 )
β(t0 )δ(t0 − t + R(t0 )/c]) 0
Z
A(x, t) = q dt (5.24)
R(t0 )

where R(t0 ) ≡ |x−x0 (t0 )|. This can be verified by using the following property
of the Dirac delta function (how?)
5.2 Liénard - Wiechert potentials : radiation fields (****) 77
Z X g(x)

g(x)δ[f (x)]dx = (5.25)
i
|df /dx| f (xi )=0

which holds for regular functions g(x) and f (x) of the integration variable x
where xi are the zeros of f (x).
• The advantage in pursuing this path is that the derivatives in eqn (5.22)
can be carried out before the integration over the delta function.
This procedure simplifies the evaluation of the fields considerably since,
we do not need to keep track of the retarded time until the last step.
We get for the electric field
 0
δ(t − t + R(t0 )/c)
Z 
E(x, t) = −q ∇ dt0
R(t0 )
β(t0 )δ(t0 − t + R(t0 )/c)
Z
q ∂
− dt (5.26)
c ∂t R(t0 )
Thus, differentiating the integrand in the first term, we get (HOW?)
R(t0 ) R(t0 )
Z     
n 0 n 0 0
E(x, t) = q δ t −t+ − δ t −t+ dt0
R2 c cR c
β(t0 )δ(t0 − t + R(t0 )/c)
Z
q ∂
− (5.27)
c ∂t R(t0 )
But (HOW?)
R(t0 ) R(t0 )
   

δ 0 t0 − t + = − δ t0 − t + (5.28)
c ∂t c
R(t0 ) R(t0 )
   
(n − β)
Z Z
n 0 0 q ∂ 0
E(x, t) = q δ t − t + dt + δ t − t + dt0
R2 c c ∂t cR(t0 ) c
(5.29)

We evaluate the integrals using the Dirac delta function expressed in equa-
tion (5.25). But we need to know the derivatives of the delta function’s argu-
ments with respect to t0 . Using the chain rule of differentiation
R(t0 )
 
d 0
t − t + = (1 − n · β)t̃ (5.30)
dt0 c
with which we get the result (HOW?):
   
n q ∂ n−β
E(r, t) = q + (5.31)
(1 − n · β)R2 t̃ c ∂t (1 − n · β)R t̃
Since
∂t0 ∂t0 ∂t0
      
∂R ∂R ∂ t̃ 1
= = −n · v =c 1− ⇒ =
∂t ∂t0 ∂t ∂t ∂t ∂t (1 − n · β)
(5.32)
78 5 Radiation by Moving Charges

Thus
   
∂ n−β 1 ∂ n−β
= (5.33)
∂t (1 − n · β)R t̃ (1 − n · β) ∂ t̃ (1 − n · β)R2 t̃

By using the additional pieces

Ṙ|t̃ = −c (n · β)t̃ (5.34)


c
ṅ|t̃ = [n(n · β) − β]t̃ (5.35)
R
d  
(1 − n · β)t̃ = − n · β̇ + β · ṅ (5.36)
dt̃
and we finally get
 h i
 (n − β)(1 − β 2 ) n × (n − β) × β̇ 
E(r, t) = q + (5.37)
 (1 − n · β)3 R2 c(1 − n · β)3 R 

A similar procedure for B shows that

B(r, t) = ∇ × A = n(t̃) × E (5.38)

5.2.1 Some observations

• When the particle is at rest and unaccelerated with respect to us, the
field reduces simply to Coulomb’s law qn/R2 . whatever corrections are
introduced the do not alter the empirical law.
• We also see a clear separation into the near field (which falls off as 1/R2 )
and the radiation field (which falls off as 1/R)
• Unless the particle is accelerated (β̇ 6= 0), the field falls off rapidly at large
distances. But when the radiation field is present, it dominate over the
near field far from the source.
• As β → 1 with β̇ = 0 the field displays a “bunching” effect. This “bunch-
ing” is understood as being a retardation effect, resulting from the finite
velocity of EM waves.

5.3 Power radiated by an accelerated charge (****)


If the velocity of an accelerated charge is small compared to the speed of light
(β → 0) then from eqn (5.37) we get
" #
q n × (n × β̇)
E= (5.39)
c R
ret
5.3 Power radiated by an accelerated charge (****) 79

Fig. 5.3.

Fig. 5.4.

The instantaneous energy flux is given by the Poynting vector


c c
S= E×B= |E|2 n (5.40)
4π 4π
The power radiated per unit solid angle is

dP c 2 2 q2
= R |E| = |n × (n × β̇)|2 (5.41)
dΩ 4π 4πc
and if Θ is the angle between the acceleration v̇ and n then the power radiated
can be written as
80 5 Radiation by Moving Charges

dP q2
= |v̇|2 sin2 Θ (5.42)
dΩ 4πc3

5.3.1 Larmor Formula

The total instantaneous power radiated is obtained by integration over the


solid angle. Thus
Z π
q2 2 3 2 q2 2
P = |v̇| 2π sin ΘdΘ = |v̇| (5.43)
4πc3 0 3 c3
This expression is known as the Larmor formula for a nonrelativistic accel-
erated charge.

Fig. 5.5.

NOTE : From equation (5.39) is obvious that the radiation is polarized


in the plane containing v̇ and n.

5.3.2 Relativistic Extension

Larmor’s formula (5.43) has an “easy” relativistic extension so that can be


applied to charges with arbitrary velocities
2 e2
P = |ṗ|2 (5.44)
3 m 2 c3
where m is the mass of the charged particle and p its momentum.
The Lorentz invariant generalization is
2 e2 dpµ dpµ
 
P =− (5.45)
3 m2 c3 dτ dτ
5.4 Applications (***) 81

where dτ = dt/γ is the proper time and pµ is the charged particle’s energy-
momentum 4-vector. Obviously for small β it reduces to (5.44)
 2 2  2  2
dpµ dpµ

dp 1 dE dp 2 dp
− = − 2 = −β (5.46)
dτ dτ dτ c dτ dτ dτ
If (5.45) is expressed in terms of the velocity & acceleration (E = γmc2 &
p = γmv with γ = 1/(1 − β 2 )1/2 ), we obtain the Liénard result (HOW?)
2 e2 6  2 
 2 
P = γ β̇ − β × β̇ (5.47)
3 c

5.4 Applications (***)


• In the charged-particle accelerators radiation losses are sometimes the
limiting factor in the maximum practical energy attainable.
• For a given applied force the radiated power (5.45) depends inversely
on the square of the mass of the particle involved. Thus these radiative effects
are largest for electrons.
• In a linear accelerator the motion is 1-D. From (5.46) we can find
that the radiated power is
 2
2 e2 dp
P = (5.48)
3 m2 c3 dt
The rate of change of momentum is equal to the rate of change of the
energy of the particle per unit distance. Thus
2
2 e2

dE
P = (5.49)
3 m2 c3 dx
showing that for linear motion the power radiated depends only on the exter-
nal forces which determine the rate of change of particle energy with distance,
not on the actual energy or momentum of the particle.
The ratio of power radiated to power supplied by external sources is
P 2 e2 1 dE 2 (e2 /mc2 ) dE
= → (5.50)
dE/dt 3 m2 c3 v dx 3 mc2 dx
Which shows that the radiation loss in an electron linear accelerator will be
unimportant unless the gain in energy is of the order of mc2 = 0.5MeV in a
distance of e2 /mc2 = 2.8 × 10−13 cm, or of the order of 2 × 1014 MeV/meter.
Typically radiation losses are completely negligible in linear accelerators since
the gains are less than 50MeV/meter.

? Can you find out what will happen in circular accelerators


like synchrotron or betatron?
82 5 Radiation by Moving Charges

In circular accelerators like synchrotron or betatron can change drastically.


In this case the momentum p changes rapidly in direction as the particle
rotates, but the change in energy per revolution is small. This means that:

dp
= γω|p|  1 dE (5.51)
dτ c dτ

Then the radiated power, eqn (5.45), can be written approximately

2 e2 2 2 2 2 e2 c 4 4
P = 2 3
γ ω |p| = β γ (5.52)
3m c 3 ρ2

where ω = (cβ/ρ), ρ being the orbit radius.


The radiative loss per revolution is:

2πρ 4π e2 3 4
δE = P = β γ (5.53)
cβ 3 ρ

For high-energy electrons (β ≈ 1) this gets the numerical value

[E(GeV)]4
δE(MeV) = 8.85 × 10−2 (5.54)
ρ(meters)

In a 10GeV electron sychrotron (Cornell with ρ ∼ 100m) the loss per revo-
lution is 8.85MeV. In LEP (CERN) with beams at 60 GeV (ρ ∼ 4300m) the
losses per orbit are about 300 MeV.

5.5 Angular Distribution of Radiation Emitted by an


Accelerated Charge (****)
The energy per unit area per unit time measured at an observation point at
time t of radiation emitter by charge at time t0 = t − R(t0 )/c is:
 2 
2 
e 1 n × [(n − β) × β̇] 

[S · n]ret = (5.55)
4πc  R2 (1 − β · n)3


ret

The energy radiated during a finite period of acceleration, say from t0 = T1


to t0 = T2 is
Z T2 +R(T2 )/c Z t0 =T2
dt 0
E= [S · n]ret dt = (S · n) dt (5.56)
T1 +R(T1 )/c t0 =T 1 dt0

Note that the useful quantity is (S · n)(dt/dt0 ) i.e. the power radiated per unit
area in terms of the charge’s own time. Thus we define the power radiated
per unit solid angle to be
5.5 Angular Distribution of Radiation Emitted by an Accelerated Charge (****) 83

dP (t0 ) dt
= R2 (S · n) 0 = R2 (S · n) (1 − β · n) (5.57)
dΩ dt
If β and β̇ are nearly constant (e.g. if the particle is accelerated for short
time) then (5.57) is proportional to the angular distribution of the energy
radiated.
For the Poynting vector (5.55) the angular distribution is

dP (t0 ) e2 |n × {(n − β) × β̇}|2


= (5.58)
dΩ 4πc (1 − n · β)5

The simplest example is linear motion in which β and β̇ are parallel i.e.
β × β̇ = 0 and (HOW?)

dP (t0 ) e2 v̇ 2 sin2 θ
= (5.59)
dΩ 4πc3 (1 − β cos θ)5
For β  1, this is the Larmor result (5.42). But as β → 1, the angular
distribution is tipped forward and increases in magnitude.

Fig. 5.6. radiation pattern for charge accelerated in its direction of motion. the two
patterns are not to scale, the relativistic one (appropriate for γ ∼ 2) having been
reduced by a factor ∼ 102 for the same acceleration.

• The angle θmax for which the intensity is maximum is:


1 p  1
cos θmax = 1 + 15β 2 − 1 for β ≈ 1 → θmax ≈ (5.60)
3β 2γ
For relativistic particles, θmax is very small, thus the angular distribution is
confined to a very narrow cone in the direction of motion.
• For small angles the angular distribution (5.59) can be written

dP (t0 ) 8 e2 v̇ 2 8 (γθ)2
≈ γ (5.61)
dΩ π c3 (1 + γ 2 θ2 )5
84 5 Radiation by Moving Charges

Fig. 5.7. Angular distribution of radiation for relativistic particles

The peak occures at γθ = 1/2, and the half-power points at γθ = 0.23 and
γθ = 0.91.
• The root mean square angle of emission of radiation in the relativistic
limit is
1 mc2
hθ2 i1/2 = = (5.62)
γ E
The total power can be obtained by integrating (5.59) over all angles

2 e2 2 6
P (t0 ) = v̇ γ (5.63)
3 c3
in agreement with (5.47) and (5.48). In other words this is a generalization of
Larmor’s formula.
• It is instructive to express this is terms of the force acting on the particle.
This force is F = dp/dt where p = mγv is the particle’s relativistic
momentum. For linear motion in the x-direction we have px = mvγ and
dpx
= mv̇γ + mv̇β 2 γ 3 = mv̇γ 3
dt
and Larmor’s formula can be written as
2 e2 |F|2
P = (5.64)
3 c3 m2
This is the total charge radiated by a charge in instantaneous linear motion.

5.6 Angular distribution of radiation from a charge in


circular motion (****)
The angular distribution of radiation for a charge in instantaneous circular
motion with acceleration β̇ perpendicular to its velocity β is another example.
5.6 Angular distribution of radiation from a charge in circular motion (****) 85

Fig. 5.8.

We choose a coordinate system such as β is in the z-direction and β̇ in


the x-direction then the general formula (5.58) reduces to (HOW?)

dP (t0 ) e2 |v̇|2 sin2 θ cos2 φ


 
= 1 − (5.65)
dΩ 4πc3 (1 − β cos θ)3 γ 2 (1 − β cos θ)2
Although, the detailed angular distribution is different from the linear accel-
eration case the characteristic peaking at forward angles is present. In the
relativistic limit (γ  1) the angular distribution can be written

dP (t0 ) 2e2 |v̇|2 4γ 2 θ2 cos2 φ


 
≈ 3 γ6 1 − (5.66)
dΩ πc (1 + γ 2 θ2 )3 γ 2 (1 + γ 2 θ2 )2

The root mean square angle of emission in this approximation is similar


to (5.62) just as in the 1-dimensional motion. (SHOW IT?)
The total power radiated can be found by integrating (5.65) over all
angles or from (5.47)
2 e2 |v̇|2 4
P (t0 ) = γ (5.67)
3 c3
Since, for circular motion, the magnitude of the rate of momentum is equal
to the force i.e. γmv̇ we can rewrite (5.67) as
2
2 e2 2

dp
Pcircular (t0 ) = γ (5.68)
3 m2 c3 dt

If we compare with the corresponding result (5.48) for rectilinear


motion, we find that the radiation emitted with a transverse accel-
eration is a factor γ 2 larger than with a parallel acceleration.
86 5 Radiation by Moving Charges

5.7 Radiation from a charge in arbitrary motion (***)


• For a charged particle in arbitrary & extremely relativistic motion the
radiation emitted is a superposition of contributions coming from accelerations
parallel to and perpendicular to the velocity.
• But the radiation from the parallel component is negligible by a factor
1/γ 2 , compare (5.48) and (5.68). Thus we will keep only the perpendicular
component alone.
• In other words the radiation emitted by a particle in arbitrary motion
is the same emitted by a particle in instantaneous circular motion, with a
radius of curvature ρ
v2 c2
ρ= ≈ (5.69)
v̇⊥ v̇⊥
where v̇⊥ is the perpendicular component of the acceleration.
• The angular distribution of radiation given by (5.65) and (5.66) corre-
sponds to a narrow cone of radiation directed along the instantaneous velocity
vector of the charge.
• The radiation will be visible only when the particle’s velocity is directed
toward the observer.
Since the angular width of the beam is 1/γ the particle will travel a dis-
tance of the order of
ρ
d=
γ
in a time
ρ
∆t =
γv
while illuminating the observer
If we consider that during the illumination the pulse is rectangular , then
in the time ∆t the front edge of the pulse travels a distance
ρ
D = c ∆t =
γβ
Since the particle is moving in the same direction with speed v and moves a
distance d in time ∆t the rear edge of the pulse will be a distance
 
1 ρ ρ
L=D−d= −1 ≈ 3 (5.70)
β γ 2γ
behind the front edge as the pulse moves off.
The Fourier decomposition of a finite wave train, we can find that the
spectrum of the radiation will contain appreciable frequency components up
to a critical frequency,  
c c 3
ωc ∼ ∼ γ (5.71)
L ρ
• For circular motion the term c/ρ is the angular frequency of rotation ω0
and even for arbitrary motion plays the role of the fundamental frequency.
5.8 Distribution in Frequency and Angle of Energy Radiated ... (****) 87

Fig. 5.9. A relativistic particle in periodic motion emits a spiral radiation pattern
that an observer at the point A detects as short bursts of radiation of time duration
T = L/c, occurring at regular intervals T0 = L0 /c. The pulse length is given by
(5.70), while the interval T0 = 2πρ/v ≈ 2πρ/c.

• This shows that a relativistic particle emits a broad spectrum of fre-


quencies up to γ 3 times the fundamental frequency.
• EXAMPLE : In a 200MeV sychrotron, γmax ≈ 400, while ω0 ≈
3 × 108 s−1 . The frequency spectrum of emitted radiation extends up to ω ≈
2 × 1016 s−1 .

5.8 Distribution in Frequency and Angle of Energy


Radiated ... (****)
• The previous qualitative arguments show that for relativistic motion the
radiated energy is spread over a wide range of frequencies.
• The estimation can be made precise and quantidative by use of Par-
seval’s theorem of Fourier analysis.
88 5 Radiation by Moving Charges

The general form of the power radiated per unit solid angle is

dP (t)
= |A(t)|2 (5.72)
dΩ
where  c 2
A(t) = [R E]ret (5.73)

and E is the electric field defined in (5.37).
• Notice that here we will use the observer’s time instead of the retarded
time since we study the observed spectrum.
The total energy radiated per unit solid angle is the time integral of (5.72):
Z ∞
dW
= |A(t)|2 dt (5.74)
dΩ −∞

This can be expressed via the Fourier transform’s as an integral over the
frequency.
The Fourier transform is:
Z ∞
1
A(ω) = √ A(t)eiωt dt (5.75)
2π −∞
and its inverse: Z ∞
1
A(t) = √ A(ω)e−iωt dω (5.76)
2π −∞

Then eqn (5.74) can be written


Z ∞ Z ∞ Z ∞
dW 1 0
= dt dω dω 0 A∗ (ω 0 ) · A(ω)ei(ω −ω)t (5.77)
dΩ 2π −∞ −∞ −∞

If we interchange the order of integration between t and ω we see that the


time integral is the Fourier represendation of the delta function δ(ω 0 − ω).
Thus the energy radiated per unit solid angle becomes
Z ∞
dW
= |A(ω)|2 dω (5.78)
dΩ −∞

The equality of equations (5.74) and (5.78) is a special case of Parseval’s


theorem.
NOTE: It is customary to integrate only over positive frequencies, since
the sign of the frequency has no physical meaning.
The energy radiated per unit solid angle per unit frequency interval is
Z ∞ 2
dW d I(ω, n)
= dω (5.79)
dΩ 0 dωdΩ
where
5.8 Distribution in Frequency and Angle of Energy Radiated ... (****) 89

d2 I
= |A(ω)|2 + |A(−ω)|2 (5.80)
dωdΩ
If A(t) is real, form (5.75) - (5.76) it is evident that A(−ω) = A∗ (ω). Then

d2 I
= 2|A(ω)|2 (5.81)
dωdΩ
which relates the power radiated as a funtion of time to the frequency spec-
trum of the energy radiated.
NOTE : We rewrite eqn (5.37) for future use
 h i
 n−β n × (n − β) × β̇ 
E(r, t) = e + (5.82)
 γ 2 (1 − n · β)3 R2 c(1 − n · β)3 R 
ret

By using (5.82) we will try to derive a general expression for the energy
radiated per unit solid angle per unit frequency interval in terms of an integral
over the trajectory of the particle.
We must calculate the Fourier transform of (5.73) by using (5.82)
 2 1/2 Z ∞ " #
e n × [(n − β) × β̇]
A(ω) = eiωt dt (5.83)
8π 2 c −∞ (1 − β · n)3
ret

where ret means evaluated at t = t0 + R(t0 )/c. By changing the integration


variable from t to t0 we get
1/2 Z ∞
e2

0 0 n × [(n − β) × β̇] 0
A(ω) = eiω(t +[R(t )/c]) dt (5.84)
8π 2 c −∞ (1 − β · n)2

since the observation point is assumed to be far away the unit vector n
can be assumed constant in time, while we can use the approximation

R(t0 ) ≈ x − n · r(t0 ) (5.85)

where x is the distance from the origin O to the observation point P , and
r(t0 ) is the position of the particle relative to O.
Then (5.84) becomes:
1/2 Z ∞
e2

n × [(n − β) × β̇]
A(ω) = eiω(t−n·r(t)/c) dt (5.86)
8π 2 c −∞ (1 − β · n)2

and the energy radiated per unit solid angle per unit frequency interval (5.81)
is Z 2
d2 I e2 ω 2 ∞ iω(t−n·r(t)/c) n × [(n − β) × β̇]
= e dt (5.87)
dωdΩ 4π 2 c −∞ (1 − β · n)2


90 5 Radiation by Moving Charges

Fig. 5.10.

For a specified motion r(t) is known, β(t) and β̇(t) can be computed, and the
integral can be evaluated as a function of ω and the direction of n.
If we study more than one accelerated charged particles, a coherent sum of
amplitudes Aj (ω) (one for each particle) must replace must replace the single
amplitude in (5.87).
If one notices that, the integrand in (5.86) is a perfect differential (exclud-
ing the exponential)
 
n × [(n − β) × β̇] d n × (n × β)
= (5.88)
(1 − β · n)3 dt 1−β·n
then by integration by parts we get to the following relation for the intensity
distribution:
2
d2 I e2 ω 2 ∞
Z
iω(t−n·r(t)/c)

= 2
n × (n × β)e dt (5.89)
dωdΩ 4π c −∞

For a number of charges ej in accelerated motion the integrand in (5.89)


becomes
N
X
−i(ω/c)n·r(t)
eβe → ej β j e−i(ω/c)n·rj (t) (5.90)
j=1

In the limit of a continuous distribution of charge in motion the sum over j


becomes an integral over the current density J(x, t) :
Z
1
eβe−i(ω/c)n·r(t) → d3 xJ(x, t)e−i(ω/c)n·x (5.91)
c
Then the intensity distribution becomes:
2
d2 I ω 2
Z Z
3 iω(t−n·x/c)

= 2 3
dt d x e n × [n × J(x, t)] (5.92)
dωdΩ 4π c

a result that can be obtained from the direct solution of the inhomogeneous
wave equation for the vector potential.
5.9 What Is Synchrotron Light? (***) 91

5.9 What Is Synchrotron Light? (***)


When charged particles are accelerated, they radiate. If electrons are con-
strained to move in a curved path they will be accelerating toward the inside
of the curve and will also radiate what we call synchrotron radiation.
• Synchrotron radiation of this type occurs naturally in the distant reaches
of outer space.
• Accelerator-based synchrotron light was seen for the first time at the GE
Research Lab (USA) in 1947 in a type of accelerator known as a synchrotron.
• First considered a nuisance because it caused the particles to lose energy,
it recognized in the 1960s as light with exceptional properties.
• The light produced at today’s light sources is very bright. In other words,
the beam of x-rays or other wavelengths is thin and very intense. Just as laser
light is much more intense and concentrated than the beam of light generated
by a flashlight.

Fig. 5.11.

5.9.1 Synchrotrons

Synchrotrons are particle accelerators - massive (roughly circular) machines


built to accelerate sub-atomic particles to almost the speed of light.
• The accelerator components include an electron gun, one or more in-
jector accelerators (usually a linear accelerator and a synchrotron but some-
times just a large linear accelerator) to increase the energy of the electrons,
and a storage ring where the electrons circulate for many hours.
• In the storage ring, magnets force the electrons into circular paths.
• As the electron path bends, light is emitted tangentially to the curved
path and streams down pipes called beamlines to the instruments where sci-
entists conduct their experiments.
92 5 Radiation by Moving Charges

Fig. 5.12. Components of a synchrotron light source typically include (1) an elec-
tron gun, (2) a linear accelerator, (3) a booster synchrotron, (4) a storage ring, (5)
beamlines, and (6) experiment stations.

• The storage ring is specifically designed to include special magnetic struc-


tures known as insertion devices (undulators and wigglers).
• Insertion devices generate specially shaped magnetic fields that drive
electrons into an oscillating trajectory for linearly polarized light or sometimes
a spiral trajectory for circularly polarized light.
• Each bend acts like a source radiating along the axis of the insertion
device, hence the light is very intense and in some cases takes on near-laser-
like brightness.

Fig. 5.13. The largest light source facilities are campuses onto themselves with
administrative, office, and laboratory buildings in addition to the light source itself.

They produce synchrotron radiation - an amazing form of light that re-


searchers are shining on molecules, atoms, crystals and innovative new mate-
rials in order to understand their structure and behaviour. It gives researchers
unparalleled power and precision in probing the fundamental nature of matter.
5.10 Synchrotron Radiation (****) 93

5.10 Synchrotron Radiation (****)


• To find the distribution of energy in frequency and in angle it is necessary
to calculate the integral (5.89)
• Because the duration of the pulse is very short, it is necessary to know
the velocity β and the position r(t) over only a small arc of the trajectory.
• The origin of time is chosen so that at t = 0 the particle is at the origin
of coordinates.
• Notice tht only for very small angles θ there will be appreciable radiation
intensity.
The vector part of the integrand in eqn (5.89) can be written
 
n × (n × β) = β −k sin(vt/ρ) + ⊥ cos(vt/ρ) sin θ (5.93)
• k = 2 is a unit vector in the y-direction, corresponding to the polariza-
tion in the plane of the orbit.
• ⊥ = n×2 is the orthogonal polarization vector corresponding approx.
to polarization perpendicular to the orbit plane (for small θ).
• The argument of the exponential is
     
n · r(t) ρ vt
ω t− = ω t − sin cos θ (5.94)
c c ρ
• Since we are dealing with small angle θ and very short time intervals we
can make an expansion to both trigonometric functions to obtain
94 5 Radiation by Moving Charges

Fig. 5.14. The trajectory lies on the plane x − y with instantaneous radius of
curvature ρ. The unit vector n can be chosen to lie in the x − z plane, and θ is the
angle with the x-axis.

c2 3
    
n · r(t) ω 1 2
ω t− ≈ + θ t + 2t (5.95)
c 2 γ2 3ρ

where β was set to unity wherever possible.


CHECK THE ABOVE RELATIONS
Thus the radiated energy distribution (5.89) can be written

d2 I e2 ω 2 2
= 2
−k Ak (ω) + ⊥ A⊥ (ω) (5.96)
dωdΩ 4π c
where the two amplitudes are (How?)

c ∞ c2 t 3
Z    
ω 1 2
Ak (ω) ≈ t exp i + θ t + dt (5.97)
ρ −∞ 2 γ2 3ρ2
Z ∞
c2 t 3
   
ω 1 2
A⊥ (ω) ≈ θ exp i + θ t + dt (5.98)
−∞ 2 γ2 3ρ2

by changing the integration variable


ct
x=
ρ(1/γ 2 + θ2 )1/2

and introducing the parameter ξ


 3/2
ωρ 1
ξ= + θ2 (5.99)
3c γ2

allows us to transform the integrals into the form


5.10 Synchrotron Radiation (****) 95
 Z∞   
ρ 1 2 3 1 3
Ak (ω) ≈ +θ x exp i ξ x + x dx (5.100)
c γ2 −∞ 2 3
 1/2 Z ∞   
ρ 1 2 3 1 3
A⊥ (ω) ≈ θ + θ exp i ξ x + x dx (5.101)
c γ2 −∞ 2 3

These integrals are identifiable as Airy integrals or as modified Bessel functions


(FIND OUT MORE)
Z ∞   
3 1 1
x sin i ξ x + x3 dx = √ K2/3 (ξ) (5.102)
0 2 3 3
Z ∞   
3 1 1
cos i ξ x + x3 dx = √ K1/3 (ξ) (5.103)
0 2 3 3
The energy radiated per unit frequency interval per unit solid angle is:
2 
d2 I e2  ωρ 2 1 θ2
 
2 2 2
= +θ K2/3 (ξ) + K (ξ) (5.104)
dωdΩ 3π 2 c c γ2 1/γ 2 + θ2 1/3

• The 1st term corresponds to radiation polarized in the orbital plane.


• The 2nd term term to radiation polarized perpendicular to that plane.
By integration over all frequencies we find the distribution of energy in
angle (Can you prove it?)
Z ∞ 2
7 e2 θ2
 
dI d I 1 5
= dω = 1+ (5.105)
dΩ 0 dωdΩ 16 ρ (1/γ 2 + θ2 )5/2 7 (1/γ 2 ) + θ2

This shows the characteristic behavior seen in the circular motion case e.g. in
equation (5.66).
This result can be obtained directly, by integrating a slight generalization
of the power formula for circular motion, eqn (5.65), over all times. Again:
• The 1st term corresponds polarization parallel to the orbital plane.
• The 2nd term term to perpendicular polarization.
Integration over all angles shows that seven (7) times as much energy is
radiated with parallel polarization as with perpendicular polarization. In other
words:
The radiation from a relativistically moving charge is very strongly, but
not completely, polarized in the plane of motion.
• The radiation is largely confined to the plane containing the motion,
being more confined the higher the frequency relative to c/ρ.
• If ω gets too large, then ξ will be large at all angles, and then there
will be negligible power emitted at those high frequencies.
• The critical frequency beyond which there will be negligible total energy
emitted at any angle can be defined by ξ = 1/2 and θ = 0 (WHY?).
Then we find    3
3 c 3 E c
ωc = γ 3 = (5.106)
2 ρ 2 mc2 ρ
96 5 Radiation by Moving Charges

this critical frequency agrees with the qualitative estimate (5.71).


• If the motion is circular, then c/ρ is the fundamental frequency of
rotation, ω0 .
• The critical frequency is given by
 3
3 E
ωc = nc ω0 with harmonic number nc = (5.107)
2 mc2

For γ  1 the radiation is predominantly on the orbital plane and we can


evaluate via eqn (5.104) the angular distribution for θ = 0.
Thus for ω  ωc we find
2  1/3  
d2 I e2 Γ (2/3)

3 ωρ 2/3
|θ=0 ≈ (5.108)
dωdΩ c π 4 c

For ω  ωc
d2 I 3 e2 2 ω −ω/ωc
|θ=0 ≈ γ e (5.109)
dωdΩ 4π c ωc
These limiting cases show that the spectrum at θ = 0 increases with frequency
roughly as ω 2/3 well bellow the critical frequency, reaches a maximum in the
neighborhood of ωc , and then drops exponentially to 0 above that frequency.
• The spread in angle at a fixed frequency can be estimated by deter-
mining the angle θc at which ξ(θc ) ≈ ξ(0) + 1.
• In the low frequency range (ω  ωc ), ξ(0) ≈ 0 so ξ(θc ) ≈ 1 which
gives
 1/3  1/3
3c 1 2ωc
θc ≈ = (5.110)
ωρ γ ω
We note that the low frequency components are emitted at much wider
angles than the average, hθ2 i1/2 ∼ γ −1 .
• In the high frequency limit (ω > ωc ), ξ(0)  1 and the intensity
falls off in angle as:

d2 I d2 I 2 2
≈ |θ=0 · e−3ωγ θ /2ω0 (5.111)
dωdΩ dωdΩ
Thus the critical angle defined by the 1/e point is
 1/2
1 2ωc
θc ≈ (5.112)
γ 3ω

This shows that the high-frequency components are confined to an angular


range much smaller than average.
The frequency distribution of the total energy emitted as the particle
passes by can be found by integrating (5.104) over angles
5.10 Synchrotron Radiation (****) 97

Fig. 5.15. Differential frequency spectrum as a function of angle. For frequencies


comparable to the critical frequency ωc , the radiation is confined to angles of order
1/γ For much smaller (larger) frequencies the angular spread is larger (smaller).

π/2 ∞
d2 I d2 I
Z Z
dI
= 2π cos θdθ ≈ 2π dθ (5.113)
dω −π/2 dωdΩ −∞ dωdΩ
• For the low-frequency range we can use (5.95) at θ = 0 and (5.108) at
θc , to get
dI d2 I e2  ωρ 1/3
∼ 2πθc |θ=0 ∼ (5.114)
dω dωdΩ c c
showing the the spectrum increases as ω 1/3 for ω  ωc . This gives a very
broad flat spectrum at frequencies below ωc .
• For the high-frequency limit ω  ωc we can integrate (5.111) over
angles to get: r  1/2
dI 3π e2 ω
≈ γ e−ω/ωc (5.115)
dω 2 c ωc

A proper integration of over angles yields the expression,


dI √ e2 ω Z ω
≈ 3 γ K5/3 (x)dx (5.116)
dω c ωc ω/ωc
In the limit ω  ωc , this reduces to the form (5.114) with numerical coefficient
13/4, while for ω  ωc it is equal to (5.115).
Bellow the behavior of dI/dω as function of the frequency. The peak
intensity is of the order of e2 γ/c and the total energy is of the order of
e2 γωc /c = 3e2 γ 4 /ρ. This is in agreement with the value 4πe2 γ 4 /3ρ for the
radiative loss per revolution (5.53) in circular accelerators.
The radiation represented by (5.104) and (5.116) is called synchrotron
radiation because it was first observed in electron synchrotrons (1948).
• For periodic circular motion the spectrum is actually discrete, be-
ing composed of frequencies that are integral multipoles of the fundamental
frequency ω0 = c/ρ.
98 5 Radiation by Moving Charges

Fig. 5.16. Normalized synchrotron radiation spectrum


√ Z
1 dI 9 3
= y K5/3 (x)dx where y = ω/ωc , I = 4πe2 γ 4 /3ρ
I dy 8π

• Thus we should talk about the angular distribution of power radiated


in the nth multiple of ω0 instead of the energy radiated per unit frequency
interval per passage of the particle. Thus we can write (WHY?)
 2
dPn 1 c d2 I
= |ω=nω0 (5.117)
dΩ 2π ρ dωdΩ
 2
1 c dI
Pn = |ω=nω0 (5.118)
2π ρ dω

• These results have been compared with experiment at various energy syn-
chrotrons. The angular, polarization and frequency distributions are all in
good agreement with theory.
• Because of the broad frequency distribution shown in previous Figure,
covering the visible, ultraviolet and x-ray regions, synchrotron radiation is a
useful tool for studies in condensed matter and biology.
5.11 Thomson Scattering of Radiation (***) 99

Fig. 5.17.

Fourier transform o the electric field produced by a charged particle in


circular motion. The plots reveal that the number of relevant harmonics of
the fundamental frequency ω0 increases with γ. And the dominant harmonic is
shifted to higher frequencies. (A. γ = 1, β = 0, ωc = 1, B. γ = 1.2, β = 0.55,
ωc = 1.7, C. γ = 1.4, β = 0.7, ωc = 2.7, D. γ = 1.6, β = 0.78, ωc = 4.1,

5.11 Thomson Scattering of Radiation (***)


When a plane wave of monochromatic EM radiation hits a free particle of
charge e and mass m the particle will be accelerated and so emit radiation.
The radiation will be emitted in directions other than the propagation
direction of the incident wave, but (for non-relativistic motion of the particle)
it will have the same frequency as the incident radiation.
According to eqn (5.41) the instantaneous power radiated into polarization
state  by a particle is (How?)

dP e2 2
= |∗ · v̇| (5.119)
dΩ 4πc3
If the propagation vector k0 and its the polarization vector 0 can be written
100 5 Radiation by Moving Charges

E(x, t) = 0 E0 eik0 ·x−iωt

Fig. 5.18.

Then from the force eqn (F = qE) the acceleration will be


e
v̇(t) = 0 E0 eik0 ·x−iωt (5.120)
m
If we assume that the charge moves a negligible part of a wavelegth during
one cycle of oscillation, the time average of |v̇|2 is 12 <(v̇ · v̇ ∗ )
Then the averaged power per unit solid angle can be expressed as
2
e2

dP c
h i= |E0 |2 |∗ · 0 |2 (5.121)
dΩ 8π mc2

And since the phenomenon is practically scattering then it is convenient to


used the scattering cross section as

dσ Energy radiated/unit time/unit solid angle


= (5.122)
dΩ Incident energy flux in energy/unit area/unit time

The incident energy flux is the time averaging Poynting vector for the plane
wave i.e. c|E0 |2 /8π. Thus from eqn (5.121) we get the differential scattering
cross section  2 2
dσ e
= |∗ · 0 |2 (5.123)
dΩ mc2

The scattering geometry with a choice of polarization vectors for the out-
going wave is shown in the Figure.
5.11 Thomson Scattering of Radiation (***) 101

The polarization vector 1 is in the plane containing n and k0 ; 2 is


perpendicular to it.
In terms of unit vectors parallel to the coordinate axes, 1 and 2 are:

1 = cos θ (x cos φ + y sin φ) − z sin θ


2 = −x sin φ + y cos φ

For an incident linearly polarized wave with polarization parallel to the


x-axis, the angular distribution is (cos2 θ cos2 φ + sin2 φ).
For polarization parallel to the y-axis it is (cos2 θ sin2 φ + cos2 φ).
For unpolarized incident radiation the scattering cross section is
 2 2
dσ e 1
1 + cos2 θ

= 2
(5.124)
dΩ mc 2
This is called the Thomson formula for scattering of radiation by a
free charge, and is appropriate for the scattering of x-rays by electrons or
gamma rays by protons.
The total scattering cross section called the Thomson cross section
 2 2
8π e
σT = (5.125)
3 mc2

The Thomson cross section for electrons is 0.665 × 10−24 cm2 . The unit of
length e2 /mc2 = 2.82 × 10−13 cm is called classical electron radius.
• This classical Thomson formula is valid only for low frequencies where
the momentum of the incident photon can be ignored.
• When the photon momentum ~ω/c becomes comparable to or larger
than mc modifications occur.
• The most important is that the energy or momentum of the scattered
photon is less than the incident energy because the charged particle recoils
during the collision.
102 5 Radiation by Moving Charges

The outgoing to the incident wave number is given by Compton formula


 −1
0 ~ω
k /k = 1 + (1 − cos θ) (5.126)
mc2

In quantum mechanics the scattering of photons by spinless point particles of


charge e and mass m yields the cross section:
2  2
e2 k0


= |∗ · 0 |2 (5.127)
dΩ mc2 k
6
Special Theory of Relativity (****)

6.1 Introduction
Einstein’s theory of special relativity is based on the assumption (which might
be a deep-rooted superstition in physics) that all physical laws should be
invariant under transformation between inertial systems.
The demand that Maxwell’s equations should be invariant under trans-
formations, and the failure of Galilean transformations to do it led to the
Lorentz transformations (β = v/c, γ = (1 − β 2 )−1/2 )

x0 = γ(x00 − βx01 )
x1 = γ(x01 − βx00 ) (6.1)
x2 = x02
x3 = x03

x0 = γ(x00 + βx01 )
x1 = γ(x01 + βx00 ) (6.2)
x2 = x02
x3 = x03

under which for example the equations of a spherical wave

c2 t2 − x2 + y 2 + z 2 = 0

(6.3)

propagating with fixed velocity c are invariant.


Lorentz transformations in general demand that the norm

s2 = x20 − x21 + x22 + x23



(6.4)

is invariant.
104 6 Special Theory of Relativity (****)

1st Postulate : The laws of nature and the results of all experiments
performed in a given frame of reference are independent of the translational
motion of the system as a whole
2nd Postulate : The speed of light is finite and independent of the motion
of the source
From the 1st postulate it follows that the mathematical equations express-
ing the laws of nature must be covariant, that is, invariant in form, under
the Lorentz transformations.
These demands call for rules on the ways that the scalars, 4-vectors and
4-tensors will transform in a spacetime whose norm is defined by (6.4).
SPACETIME
The space-time continuum is defined in terms of a 4-dimensional space with
coordinates x0 , x1 , x2 , x3 .

6.2 Tensors

If we assume that there is a well defined transformation that yields from the
coordinates x0 , x1 , x2 , x3 a new set of coordinates x00 , x01 , x02 , x03 according
to the rule
x0α = x0α (x0 , x1 , x2 , x3 ) (α = 0, 1, 2, 3) (6.5)
Here we will defined the tensors under their transformation prop-
erties.
A scalar (tensor of rank 0) is a single quantity whose value is not changed
under the transformation. for example the interval s2 in (6.4) is a scalar.
Vectors are tensors of rank 1, and we distinguish two kinds.
The contravariant vector Aα whose components transformed according
to the rule
3

X ∂x0α β ∂x0α β
A = A ≡ A (6.6)
∂xβ ∂xβ
β=0

where the partial derivatives are calculated from (6.5). Explicitly we have 4
equations of the form:

∂x0α 0 ∂x0α 1 ∂x0α 2 ∂x0α 3


A0α = A + A + A + A (6.7)
∂x0 ∂x1 ∂x2 ∂x3

The covariant vector Bα is defined by the rule


3
X ∂xβ ∂xβ
Bα0 = B β ≡ Bβ (6.8)
∂x0α ∂x0α
β=0

where the partial derivatives are calculated from the inverse of (6.5).
6.2 Tensors 105

The contravariant tensor of rank 2 F αβ consists of 16 quantities (com-


ponents) that transform according to

∂x0α ∂x0β γδ
F 0αβ = F (6.9)
∂xγ ∂xδ
A covariant tensor of rank 2 Gαβ transforms as

∂xγ ∂xδ
G0αβ = Gγδ (6.10)
∂x0α ∂x0β
The mixed tensor of rank 2 H α β transforms as

∂x0α ∂xδ γ
H 0α β = H δ (6.11)
∂xγ ∂x0β
The generalization to arbitrary rank tensors is quite obvious extension of the
above relations.
The inner or scalar product of two vectors is defined as the product of
the components of a covariant and a contravariant vector

B · A ≡ B α Aα (6.12)

with this definition the scalar product is an invariant or scalar under the
transfomation (6.5):

∂xβ ∂x0α γ ∂xβ


B0 · A0 = Bα0 A0α = B β A = B β Aγ
∂x0α ∂xγ ∂xγ
= δ β γ B β Aγ = B γ Aγ = B · A (6.13)

The geometry of the space-time of STR is defined by the invariant interval s2


defined in (6.4), which in differential form can be written as

(ds)2 = (dx0 )2 − (dx1 )2 − (dx2 )2 − (dx3 )2 (6.14)

This norm or metric is a special case of the general differential length element

ds2 = gαβ dxα dxβ (6.15)

where gαβ = gβα is called the metric tensor.


For the flat space-time of STR the metric tensor is diagonal with elements

g00 = 1 , g11 = g22 = g33 = −1 (6.16)

The contravariant tensor g αβ is defined as the normalized cofactor of gαβ . For


the flat spacetime of STR they are the same

g αβ = gαβ (6.17)
106 6 Special Theory of Relativity (****)

The contraction of the covariant and contravariant metric tensors defines


the Kronecker delta in 4-dimensions
gαγ g γβ = δα β (6.18)
where δαβ = 0 if α 6= β and δαα = 1.
From the definition of the scalar product (6.12) and (6.15) we can easily
conclude that
xα = gαβ xβ (6.19)
and its inverse
xα = g αβ xβ (6.20)
This is a more general procedure for lowering and raising indeces
...α...
F... = g αβ F...β...
...
and G... ...β...
...α... = gαβ G... (6.21)
From the definition of the flat spacetime metric tensor we can easily prove
that:
Aα = (A0 , A) , Aα = (A0 , −A) (6.22)
The scalar product (6.12) of two vectors is
B · A ≡ B α Aα = B 0 A0 − B · A
From the transformation property
∂ ∂xβ ∂
=
∂x0α ∂x0α ∂xβ
we conclude that the differentiation with respect to a contravariant component
of the coordinate vector transforms as the component of a covariant vector.
Thus we employ the notation
   
α ∂ ∂ ∂ ∂
∂ ≡ = , −∇ , ∂α ≡ = ,∇ (6.23)
∂xα ∂x0 ∂xα ∂x0
The 4-divergence of a 4-vector A is the invariant
∂A0
∂ α Aα = ∂α Aα = +∇·A (6.24)
∂x0
an equation familiar in form from continuity of charge and current density.
The 4-dimensional Laplacian operator is defined to be the invariant con-
traction
∂2
 ≡ ∂α ∂ α = − ∇2 (6.25)
∂x0 2
which is of course the operator of the wave equation in vacuum.

The previous examples show how the covariance of a physical law emerges
provided suitable Lorentz transformation properties are attributed to the
quantities entering the equation.
6.3 Invariance of Electric Charge; Covariance in Electrodynamics 107

6.3 Invariance of Electric Charge; Covariance in


Electrodynamics

• The invariance of the equations of electrodynamics under Lorentz transforms


was shown by Lorentz and Poincaré before the formulation of the STR.
• The invariance in form or covariance of the Maxwell and Lorentz force
equations implies that the various quantities ρ, J, E, B that enter into the
equations transform in a well defined way under Lorentz transformations.
Consider first the Lorentz force equation for a charged particle
dp  v 
=q E+ ×B (6.26)
dt c
we know that p transforms as the space part of energy and momentum

pα = (p0 , p) = m (U0 , U)

where p0 = E/c and U a is the 4-velocity

dx0 dx0 dt dx dx dt
U0 ≡ = = γc , U ≡ = = γu (6.27)
dτ dt dτ dτ dt dτ

If we use the proper time of the particle which is a Lorentz invariant


quantity defined as
1 p 1
dτ = ds = dt 1 − β 2 = dt (6.28)
c γ

for the differentiation of (6.26) we can write

dp q
= (U0 E + U × B) (6.29)
dτ c
the left hand side is the space part of a 4-vector. The corresponding time
component equation is the rate of change of the energy of the particle
Z
dp0 q dEmech
= U·E ⇐ = J · Ed3 x (6.30)
dt c dt V

The right-hand sides of the previous two equations involve three factors, the
charge q, the 4-velocity and the electromagnetic fields.
If the transformation properties of two of the three factors are known and
Lorentz covariance is demanded, then the transformation properties of the
3rd factor can be established.
The experimental invariance of electric charge and the requirement of
Lorentz covariance of the Lorentz force eqn (6.29) and (6.30) determines the
Lorentz transformation properties of the EM field.
108 6 Special Theory of Relativity (****)

For example, the requirement from (6.30) that U·E be the time component
of a 4-vector establishes that the components of E are the time-space parts
of a 2nd rank tensor F αβ such that

U · E = F 0β Uβ

We will consider Maxwell equations and we begin with the charge density
ρ(x, t) and current density J(x, t) and the continuity equation

∂ρ
+∇·J=0 (6.31)
∂t
It is natural to postulate that ρ and J together form a 4-vector J α :

J α = (cρ, J) (6.32)

and the continuity equation takes the covariant form:

∂α J α = 0 (6.33)

where the covariant differential operator ∂α is given by (6.23).


If we consider the Lorentz gauge
1 ∂Φ
+∇·A=0 (6.34)
c ∂t
then the wave equations for the vector and scalar potential are

1 ∂2A 4π
2 2
− ∇2 A = J
c ∂t c
(6.35)
1 ∂2Φ
− ∇2 Φ = 4πρ
c2 ∂t2
Notice that the differential operator in (6.35) is the invariant 4-D Laplacian
(6.25) while the right hand side are the components of the 4-vector (6.32).
Obviously, Lorentz covariance requires that the potentials Φ and A form
a 4-vector potential
Aα = (Φ, A) (6.36)
Then the wave equation (6.35) and the Lorentz condition (6.34) take the
covariant forms
4π α
Aα = J , ∂α Aα = 0 (6.37)
c

The fields E and B are expressed in terms of the potentials as


1 ∂A
E=− − ∇Φ , B=∇×A (6.38)
c ∂t
6.3 Invariance of Electric Charge; Covariance in Electrodynamics 109

where, for example, the x-component of E and B are explicitly


1 ∂Ax ∂Φ
= − ∂ 0 A1 − ∂ 1 A0

Ex = − −
c ∂t ∂x
(6.39)
∂Az ∂Ay
= − ∂ 2 A3 − ∂ 3 A2

Bx = −
∂y ∂z
These equations imply that the 6 in total components of the electric and
magnetic fields are the elements of a 2nd-rank, antisymmetric field-strength
tensor
F αβ = ∂ α Aβ − ∂ β Aα (6.40)
explicitly in matrix form
 
0 −Ex −Ey −Ez
 Ex 0 −Bz By 
F αβ =  (6.41)
 E y Bz 0 −Bx 
Ez −By Bx 0

In the covariant form is:


 
0 Ex Ey Ez
 −Ex 0 −Bz By 
Fαβ = gαγ gδβ F γδ =
 −Ey Bz
 (6.42)
0 −Bx 
−Ez −By Bx 0

The elements of Fαβ are obtained from F αβ by putting E → −E.


• Notice that
 
1
Fµν F µν = 2 B 2 − 2 E 2 = invariant (6.43)
c
and the Lorentz force equation becomes

dpα dpα dxβ


= qFαβ uβ or = qFαβ (6.44)
dτ dt dt

The inhomogeneous Maxwell equations are


1 ∂E 4π
∇ · E = 4πρ , ∇×B− = J
c ∂t c
in terms of F αβ and J α they take the covariant form (HOW?)
4π β
∂α F αβ = J (6.45)
c
Similarly the homogeneous Maxwell equations are
110 6 Special Theory of Relativity (****)

1 ∂B
∇ · B = 0, ∇×E+ =0
c ∂t
take the form (HOW?)

∂α Fβγ + ∂β Fγα + ∂γ Fαβ = 0 (6.46)

With the above definitions of the various quantities and the reformulation
of the wave and Maxwell equations the covariance of the equations of EM is
established.
Finally, the Lorentz force (6.29) and rate of change of energy (6.30) can
be set in manifestly covariant form
dpα dU α q
=m = F αβ Uβ (6.47)
dτ dτ c

6.4 Dual Field-Strength Tensor


 
0 −Bx −By −Bz
1 αβγδ  Bx 0 Ez −Ey 
F αβ =  Fγδ =
 By −Ez 0 Ex 
 (6.48)
2
Bz Ey −Ex 0
where 

 +1 for α = 0, β = 1, γ = 2, δ = 3

 and for any even permutation
αβγδ = (6.49)


 −1 for any odd permutation
0 if any two indices are equal

The elements of F αβ are obtained from F αβ by putting E → B and B → −E.


The homogeneous Maxwell equations can be written in terms of the dual field-
strength tensor (prove it) as
∂α F αβ = 0 (6.50)

6.5 Transformation of Electromagnetic Fields


Since both E and B are the elements of a 2nd-rank tensor F αβ , their values in
one inertial frame can be expressed in terms of the values in another inertial
frame, according to
∂x0α ∂x0β γδ
F 0αβ = F (6.51)
∂xγ ∂xδ
If the one system travels along the direction of x1 with speed cβ the explicit
transformations are (HOW?)
6.6 Transformation of Electromagnetic Fields: Example 111

E10 = E1 B10 = B1
E20 = γ(E2 − βB3 ) B20 = γ(B2 + βE3 ) (6.52)
E30 = γ(E3 + βB2 ) B30 = γ(B3 − βE2 )

This suggest that for a general Lorentz transformation between two systems
moving with a speed v relative to each other the transformation of the fields
can be written (HOW):

γ2
E0 = γ (E + β × B) − β (β · E)
γ+1
(6.53)
γ2
B0 = γ (B − β × E) − β (β · B)
γ+1
• These transformations show that E and B have no independent existence.
• A purely electric or magnetic field in one coordinate system will appear
as a mixture of electric and magnetic fields in another coordinate frame.
• Thus one should properly speak of the electromagnetic field F αβ rather
than E and B separately.
Finally, if no magnetic field exists in a frame K 0 the inverse of (6.53) shows
that in the frame K the magnetic field B and the electric field E are linked
by the simple relation
B=β×E (6.54)
note that E is the transformed field from K 0 to K.

6.6 Transformation of Electromagnetic Fields: Example


We will study the fields seen by an observer in the system K when a point
charge q moves in a straight line with velocity v.
The charge is at rest in the system K 0 and the transformation of the fields
is given by the inverse of (6.53) or (6.53) The observer is at the point P . In
the frame K 0 the observer’s point P , where the fields are to be evaluated,
p has
coordinates x01 = −vt0 , x02 = b, x03 = 0 and is at a distance r0 = b2 + (vt)2 .
In the rest frame K 0 of the charge the electric and magnetic fields at the
observation point are (WHY?)
0
E10 = − qvt
r 03 E20 = qb
r 03 E30 = 0
B10 = 0 B20 = 0 B30 = 0

In terms of the coordinates of K the nonzero field components are


qγvt qb
E10 = − , E20 = (6.55)
(b2 + γ 2 v 2 t2 )3/2 (b2 + γ 2 v 2 t2 )3/2
112 6 Special Theory of Relativity (****)

Then using the inverse of (6.53) we find the transformed fields in the
system K:
qγvt
E1 = E10 = −
(b2 + γ 2 v 2 t2 )3/2
γqb
E2 = γE20 = 2 (6.56)
(b + γ 2 v 2 t2 )3/2
B3 = γβE20 = βE2 (6.57)

with all the other components vanishing.


• Notice the magnetic induction in the direction x3 .
• The magnetic field becomes nearly equal to the transverse electric field
E2 as β → 1.
• At low velocities (γ ≈ 1) the magnetic induction is
qv×r
B≈
c r3
which is the approximate Ampére-Biot-Savart expression for the magnetic
field of a moving charge.
• At high velocities (γ  1) we see that the transverse electric field E2
becomes equal to γ times its non-relativistic value.
• At high velocities (γ  1) the duration of appreciable field strengths
at point P is decreased.
6.6 Transformation of Electromagnetic Fields: Example 113

Fig. 6.1. Fields of a uniformly moving charged oarticle (a) Fields at the observation
point P as function of time. (b) Lines of electric force for a particle at rest and in
motion (γ = 3).
7
Dynamics of Relativistic Particles and EM
Fields (**)

7.1 Lagrangian Hamiltonian for a Relativistic Charged


Particle
The equations of motion
dp h u i
=e E+ ×B (7.1)
dt c
dE
= eu · E (7.2)
dt
for a particle with charge e in external fields E and B can be written in the
covariant form (6.47)
dU α e αβ
= F Uβ (7.3)
dτ mc
where m is the mass, τ is the proper time, and U α = (γc, γu) = pα /m is the
4-velocity of the particle.
• The Lagrangian treatment of mechanics is based on the principle of of
least action or Hamilton’s principle.
• The system is described by generalized coordinates qi (t) & velocities
q̇i (t).
• The Lagrangian L is a functional of qi (t) and q̇i (t) and perhaps the
time.
• The action A is the time integral of L along a possible path of the
system.
• The principle of least action states that the motion of a mechanical
system is such that in going from a configuration a at time t1 to a configuration
b in time t2 the action
Z t2
A= L [qi (t), q̇i (t), t] dt (7.4)
t1

is an extremum.
116 7 Dynamics of Relativistic Particles and EM Fields (**)

By considering small variations of coordinates and velocities away from the


actual path and requiring δA = 0, one obtains the Euler-Lagrange equations
of motion1  
d ∂L ∂L
− =0 (7.5)
dt ∂ q̇i ∂qi

7.1.1 Relativistic Lagrangian (Elementary)

From the 1st postulate of STR the action integral must be a Lorentz
scalar, because the equations of motion described by the extremum condition
δA = 0. Then if we set in (7.4) dt = γdτ we get
Z τ2
A= γL dτ (7.6)
τ1

since the proper time is invariant the condition that A is invariant requires
that γL is also Lorentz invariant.
The Lagrangian for a free particle can be a function of the velocity (the
only invariant function of the velocity is ηαβ U α U β = Uα U α = c2 ) and the
mass of the particle but not its position.
r
2 u2
Lfree = −mc 1 − 2 (7.7)
c
and through (7.5) the free-particle equation of motion (remember that p =
γmu)
d
(γmu) = 0 (7.8)
dt

Since the non-relativistic Lagrangian is T − V and the potential energy for


the interaction is V = eΦ, the interaction part of the relativistic Lagrangian
must reduce in the non-relativistic limit to

Lint → LN R
int = −eΦ (7.9)

A wise guess is:


e
Lint = − Uα Aα (7.10)
γc
which may comes from the demand that γLint is :
• linear in the charge of the particle
• linear in the EM potentials
• translationally invariant
• a function no higher than the 1st time derivative of the particle coordinates.
Another writing is:
1
see e.g. Goldstein “Glassical Mechanics”
7.1 Lagrangian Hamiltonian for a Relativistic Charged Particle 117
e
Lint = −eΦ + u · A (7.11)
c
and the combination of (7.7) and (7.11) yields the complete Lagrangian
r
2 u2 e
L = −mc 1 − 2 + u · A − eΦ (7.12)
c c
(Verify that leads to the Lorentz force equation)
? The canonical momentum P conjugate to the position coordinate
x is obtained by the definition
∂L e
Pi ≡ = γmui + Ai (7.13)
∂ui c
Thus the conjugate momentum is
e
P=p+ A (7.14)
c
where p = γmu is the ordinary kinetic momentum.
? The Hamiltonian H is a function of the coordinate x and its conjugate
momentum P and is a constant of motion if the Lagrangian is not an
explicit function of time, in terms of the Lagrangian is :
H =P·u−L (7.15)
by eliminating u in favor of P and x we find (HOW?) that
cP − eA
u= q 2 (7.16)
P − eA
c + m2 c4

This equation together with (7.12)


q
2
H = (cP − eA) + m2 c4 + eΦ (7.17)
(Verify that from this Lagrangian you can get the Lorentz equation)
Equation (7.17) is an expression for the total energy W of the particle.
Actually, it differs by the potential energy term eΦ and by the replacement
p → [P − (e/c)A].
These two modifications are actually a conseqency of considering 4-vectors.
Notice that 2
2 2
(W − eΦ) − (cP − eA) = mc2 (7.18)
is just the 4-vector scalar product
2
pα pα = (mc) (7.19)
where    
E 1 e
pα ≡ ,p = (W − eΦ) , P − A (7.20)
c c c
Thus the total energy W/c acts as the time component of a canonically con-
jugate 4-momentum P α of which P is given by (7.14).
118 7 Dynamics of Relativistic Particles and EM Fields (**)

7.1.2 Relativistic Lagrangian (Covariant Treatment)

If ones wants to use of proper covariant description, has to abandon the vec-
torial writing x, u and to replace them with the 4-vectors xα , U α . Then the
free particle Lagrangian (7.7) will be written as
mc p
Lfree = − Uα U α (7.21)
γ

remember that Uα U α = c2 . The action integral will be:


Z τ2 p
A = −mc Uα U α dτ (7.22)
τ1

This invariant form can be the starting point for a variational calculation
leading to the equation of motion dU α /dτ = 0. One can further make use of
the constraint
Uα U α = c2 (7.23)
or the equivalent one:
dU α
Uα =0 (7.24)

The integrant in (7.22) is:
r
dxα dxα
p q
Uα U α dτ = dτ = g αβ dxα dxβ
dτ dτ
i.e. the infinitesimal length element in 4-space. Thus the action integral can
be written as Z s2 r
dxα dxβ
A = −mc g αβ ds (7.25)
s1 ds ds
where the 4-vector coordinate of the particle is xα (s), where s is a parameter
monotonically increasing with τ , but otherwise arbitrary.
• The action integral is an integral along the world line of the particle
• The principle of least action is a statement that the actual path is the
longest path, the geodesic. We should keep in mind that
r
dxα dxβ
g αβ ds = cdτ (7.26)
ds ds
and then a straightforward variational calculation with (7.25) leads to
 
d  dxα /ds 
mc =0 (7.27)
ds  dxβ dxβ 1/2

ds ds

or
7.1 Lagrangian Hamiltonian for a Relativistic Charged Particle 119

d2 xα
m =0 (7.28)
dτ 2
as expected for a free particle motion.
For a charged particle in an external field the form of the Lagrangian (7.11)
suggests that the manifesltly covariant form of the action integral is
Z s2 " r #
dx α dx β e dxα
A=− mc g αβ + Aα (x) ds (7.29)
s1 ds ds c ds

Hamilton’s principle yields the Euler-Lagrange equations:


" #
d ∂ L̃
− ∂ α L̃ = 0 (7.30)
ds ∂(dxα /ds)

where the Lagrangian is:


" r #
dxα dxβ e dxα α
L̃ = − mc g αβ + A (x) (7.31)
ds ds c ds

Then (7.30) becomes

d2 xα e dAα (x) e dxβ α β


m + − ∂ A (x) = 0
dτ 2 c dτ c dτ

Since dAα /dτ = (dxβ /dτ )∂ β Aα this equation can be written as

d2 xα e α β  dxβ e dxβ
m 2
= ∂ A − ∂ β Aα = F αβ (7.32)
dτ c dτ c dτ
which is the covariant equation of motion (7.3).

7.1.3 Relativistic Hamiltonian (Covariant Treatment)

The transition to conjugate momenta and a Hamiltonian is simple enough.


The conjugate 4-momentum is defined by

∂ L̃ e
Pα = − = mU α + Aα (7.33)
∂(dxα /ds) c

which is in agreement with (7.4). A Hamiltonian can be determined

H̃ = Pα U α + L̃ (7.34)

the by eliminating U α by means of (7.33) leads to the expression


120 7 Dynamics of Relativistic Particles and EM Fields (**)
s
α
eAα
    
1 eAα α eA eAα α
H̃ = Pα − P − −c Pα − P −
m c c c c
(7.35)
Then by using the constraint

eAα
  
α eAα
P − Pα − = m2 c2
c c

we get Hamilton’s equations:

dxα eAα
 
∂ H̃ 1
= = Pα −
dτ ∂Pα m c
and (7.36)
α
 
dP ∂ H̃ e eAβ
=− = Pβ − ∂ α Aβ
dτ ∂xα mc c

These two equation can be shown to be equivalent to the Euler-Lagrange


equation (7.32).
7.2 Motion in a Uniform, Static Magnetic Field (****) 121

7.2 Motion in a Uniform, Static Magnetic Field (****)


We consider the motion of charged particles in a uniform and static magnetic
field. The equations (7.1) and (7.2) are:
dp e dE
= v × B, =0 (7.37)
dt c dt
where v is the particle’s velocity.
Since the energy is constant in time, the magnitude of the velocity is
constant and so is γ.
Then the first equation can be written:
dv
= v × ωB (7.38)
dt
where
eB ecB
ωB = = (7.39)
γmc E
is the gyration or precession frequency.
The motion is a circular motion perpendicular to B and a uniform trans-
lation parallel to B.
The solution for the velocity is (HOW?)

v(t) = vk 3 + ωB a(1 − i2 )e−iωB t (7.40)

3 : is a unit vector parallel to the field


1 and 2 : are the other two orthogonal unit vectors
vk : is the velocity component along the field, and
a : the gyration radius
One can see that (7.40) represents a counterclockwise rotation for positive
charge e
Further integration leads to the displacement of the particles

x(t) = X0 + vk t 3 + ia(1 − i2 )e−iωB t (7.41)

The path is a helix of radius a and pitch angle α = tan−1 (vk /ωB a).
The magnitude of the gyration radius a depends on the magnetic induction
B and the transverse momentum p⊥ of the particle

cp⊥ = eBa

This relation allows for the determination of particle momenta. For particle
with charge equal to electron charge the momentum can be written numeri-
cally as

p⊥ (MeV/c) = 3 × 10−4 Ba(gauss-cm) = 300Ba(tesla-m) (7.42)


122 7 Dynamics of Relativistic Particles and EM Fields (**)

Fig. 7.1. This three basic motions of charged particles in a magnetic field: gyro,
bounce between mirror points, and drift. The pitch angle α between the directions
of the magnetic field B and the electron velocity v.

The angle between the direction of the magnetic field and a particle’s
spiral trajectory is referred to as the ”pitch angle”, which in a non-uniform
magnetic field changes as the ratio between the perpendicular and parallel
components of the particle velocity changes. Pitch angle is important because
it is a key factor in determining whether a charged particle will be lost to the
Earth’s atmosphere or not.

7.3 Motion in Combined, Uniform, Static E- and B-


Field (***)
We will consider a charged particle moving in a combination of electric and
magnetic fields E and B, both uniform and static, and for this study they will
be considered perpendicular.
From the energy equation (7.2) we notice that the particle’s energy is
not constant in time. Consequently we can obtain a simple equation for the
velocity, as was done for a static magnetic field.
An appropriate Lorentz transformation can simplify the equations of mo-
tion, here we consider a coordinate frame K 0 moving with velocity u with
respect to original frame K. Then the Lorentz force equation for a particle in
K 0 is:
dp0 v0 × B0
 
0
=e E +
dt0 c
The fields E0 and B0 can be estimated from relations of the previous chapter,
Eq (6.53).
7.3 Motion in Combined, Uniform, Static E- and B- Field (***) 123

CASE: |E| < |B|


If we chose u to be perpendicular to the orthogonal vectors E0 and B0 i.e.
E×B
u=c (7.43)
B2
we find that the fields in K 0 (HOW?)
 u 
E0k = 0 , E0⊥ = γ E + × B = 0
c
(7.44)
2 1/2
B2 − E
 
1
B0k = 0 , B0⊥ = B= B
c B2
In the frame K 0 the only field acting is a static magnetic field B0 which points
in the same direction as B but is weaker by a factor 1/γ . Thus the motion in
K 0 is the same as in the previous section, namely spiraling around the lines
of force.
As viewed from the original frame, this gyration is accompanied by a
uniform “drift” u perpendicular to E and B.
The direction of the drift is independent of the sign of the charge of the
particle. The drift can be understood by noting that a particle that starts

Fig. 7.2. E × B-drift of charged particles in perpendicular fields

gyrating around B, is accelerated by the electric field, gains energy and so


moves in a path with a larger radius for roughly half of its cycle. On the other
half the electric field decelarates it, causing it to lose energy and so move in a
tighter arc. The combination of arcs produces a translation perpendicular to
E and B as shown in the Figure.

CASE: |E| > |B|


The electric field is so strong that the particle is continually accelerated
in the direction of E and its average energy continues to increase with time.
124 7 Dynamics of Relativistic Particles and EM Fields (**)

If we consider a Lorentz transformation to a system K 00 moving relative


to the first with velocity
E×B
u0 = c (7.45)
E2
we get in the K 00
1/2
E2 − B2

1
E00k = 0, E00⊥ = E= E
γ E2
(7.46)
u0
 
B00k = 0 , E00⊥ = γ 0 B − ×E =0
c

Thus the particle, in the system K 00 , is acted on by a purely electrostatic field


which causes hyperbolic motion with ever-increasing velocity.

7.4 Lowest Order Relativistic Corrections to the


Lagrangian...

The interaction Lagrangian was given by (7.11). In its simpler form the non-
relativistic Lagrangian for two charged particles
q1 q2
LN R
int = (7.47)
r
including lowest order relativistic effects is
  
q1 q2 1 (v1 · r)(v2 · r)
Lint = −1 + 2 v1 · v2 + (7.48)
r 2c r2

For a system of interacting charged particles the complete Darwin Lagrangian


correct to order 1/c2 can be written by expanding the free-particle Lagrangian
(7.7) for each particle and summing up all the interaction terms of the form
(7.48)
1X 1 X
LDarwin = mi vi2 + 2 mi vi4
2 i 8c i
∼ ∼
1 X qi qj 1 X qi qj
− + 2 [vi · vj + (vi · r̂ij )(vj · r̂ij )] (7.49)
2 i,j rij 4c i,j rij

where rij = |xi − xj |, and r̂ij is the unit vector in the direction xi − xj and
the “tilde” (∼) in the summation indicates omission of the self-energy terms
i = j.
7.5 Lagrangian for the Electromagnetic Field 125

7.5 Lagrangian for the Electromagnetic Field


We now examine a Lagrangian description of the EM field in interaction with
specified external sources of charge and current.
• The Lagrangian approach to to continuous fields is similar to the
approach used for discrete point particles.
• The finite number of coordinates qi (t) and q̇i (t) are replaced by an
infinite number of degrees of freedom.
• The coordinate qi is replaced by a continuous field φk (x) with a discrete
index (k = 1, 2, . . . , n) and a continuous index (xα ), i.e.

i → xα , k , qi → φk (x) q̇i → ∂ α φk (x)


X Z
L= Li (qi , q̇1 ) → L(φk , ∂ α φk )d3 x (7.50)
i
 
d ∂L ∂L ∂L ∂L
= → ∂β =
dt ∂ q̇i ∂qi ∂(∂ β φk ) ∂φk

where L is the Lagrangian density, corresponding to a definite point in


space-time and equivalent to the individual terms in a discrete particle La-
grangian like (7.49). For the EM-field the “coordinates” are Aα and the “ve-
locities” ∂ β Aα .
The action integral take sthe form
Z Z Z
A= L d3 x dt = L d4 x (7.51)

and it will be Lorentz-invariant if the Lagrangian density L is a Lorentz scalar.


In analogy with the situation with discrete particles, we expect the free-field
Lagrangian at least to be quadratic in the velocities, that is, ∂ β Aα or F αβ .
The only Lorentz invariant quadratic forms are Fαβ F αβ and Fαβ F αβ (the
last term is pseudoscalar under inversion).
Thus the Lfree will be a multipole of Fαβ F αβ and the Lint according to (7.10)
will be a multipole of Jα Aα . Thus the EM Lagrangian density is:
1 1
L=− Fαβ F αβ − Jα Aα (7.52)
16π c
If we want to use it for the Euler-Lagrange equations given in (7.50) we get
1  1
L=− gλµ gνσ (∂ µ Aσ − ∂ σ Aµ ) ∂ λ Aν − ∂ ν Aλ − Jα Aα (7.53)
16π c

∂L
The term ∂(∂ β Aα )
in the Euler-Lagrange equations becomes (how?)

∂L 1 1
β α
= − Fβα = Fαβ (7.54)
∂(∂ A ) 4π 4π
126 7 Dynamics of Relativistic Particles and EM Fields (**)

The other part of the Euler-Lagrange equations is


∂L 1
= − Jα (7.55)
∂Aα c
Thus the equations of motion for the EM field are
1 β 1
∂ Fβα = Jα (7.56)
4π c
which is a covariant form of the inhomogeneous Maxwell equations (6.45).
The conservation of the source current density can be obtained from (7.56)
1 α β 1
∂ ∂ Fβα = ∂ α Jα
4π c
The left hand side has a differential operator which is symmetric in α and β,
while Fαβ is antisymmetric. Again the contraction vanishes (why?) and we
have:
∂ α Jα = 0 . (7.57)

7.6 Proca Lagrangian; Photon Mass Effect


If we assume that the photon is not massless then the Lagrangian (7.52) has to
be modified by the addition of a “ mass” term, this is called Proca Lagrangian

1 µ2 1
LProca = − Fαβ F αβ + Aα Aα − J α Aα (7.58)
16π 8π c
The parameter µ has dimensions of inverse length and is the reciprocal Comp-
ton wavelength of the photon (µ = mγ c/~). Instead of (7.56) the Proca equa-
tions of motion are

∂ β Fβα + µ2 Aα = Jα (7.59)
c
with the same homogeneous equations ∂α F αβ = 0 as in Maxwell theory. In
contrast to the Maxwell equations the potentials have real physical (observ-
able) significance through the mass term. In the Lorentz gauge (7.59) can be
written

Aα + µ2 Aα = Jα (7.60)
c

In the static limit takes the form



∇2 Aα − µ2 Aα = − Jα (7.61)
c
7.7 Conservation Laws : Canonical Stress Tensor 127

If the source is a point charge q at rest in the origin then the only non-vanishing
component is A0 = Φ. the solution will be the spherically symmetric Yukawa
potential
q
Φ(x) = e−µr (7.62)
r
i.e. we observe an exponential falloff of the static potentials and fields, with
1/e distance equal to 1/µ.
Notice that the exponential factor alters the character of the Earth’s (and
other planets) magnetic fields sufficiently to permit us to set quite stringent
limits on the photon mass from geomagnetic data.

µ2 r2 e−µr 2 2 e−µr
 
B(x) = [3r̂(r̂ · m) − m] 1 + µr + − µ m 3
3 r3 3 r
The result shows that the Earth’s magnetic field will appear as a dipole an-
gular distribution plus an added constant magnetic field antiparallel to m.
Measurements show that this “ external” field is less than 0.004 times the
dipole field at the magnetic equator which leads to µ < 4 × 10−10 cm−1 or
mγ < 8 × 10−49 g.

7.7 Conservation Laws : Canonical Stress Tensor


In particle mechanics the transition to the Hamilton formulation and conser-
vation of energy is made by first defining the canonical momentum variables
∂L
pi =
∂ q̇i
and then introducing the Hamiltonian
X
H= pi q̇i − L (7.63)
i

where dH/dt = 0 if ∂L/∂t = 0.


For fields the Hamiltonian is the volume integral over the 3-D space of
Hamiltonian density H i.e. Z
H = Hd3 x .

It is necessary that the Hamiltonian density H transform as the tt component


of a 2nd-rank tensor.
If the Lagrangian density for some fields is a function of the field variables
φk (x) and ∂ α φk (x) with k = 1, 2, . . . , n the Hamiltonian density is defined in
analogy to (7.63) as
X ∂L ∂φk
H= −L (7.64)
∂(∂φk /∂t) ∂t
k
128 7 Dynamics of Relativistic Particles and EM Fields (**)

The first factor in the sum is the field momentum canonically conjugate to
φk (x) and ∂ α φk (x) is equivalent to the velocity q̇i .
The Lorentz transformation properties of H suggest that the covariant
generalization of the Hamiltonian density is the canonical stress tensor:
X ∂L
T αβ = ∂ β φk − g αβ L (7.65)
∂(∂α φk )
k

For the free EM field Lagrangian:


1
Lem = − Fµν F µν
16π
the canonical stress tensor is
∂Lem β λ
T αβ = ∂ A − g αβ Lem
∂(∂α Aλ )

By using equation (7.54) we find


1 αµ
T αβ = − g Fµλ ∂ β Aλ − g αβ Lem (7.66)


With L = E2 − B2 /8π and (6.42) we find (how?)

1 1
T 00 = E2 + B2 +

∇ · (ΦE)
8π 4π
1 1
T 0i = (E × B)i + ∇ · (Ai E) (7.67)
4π 4π  
1 1 ∂
T i0 = (E × B)i + (∇ × ΦB)i − (ΦEi )
4π 4π ∂x0
If the fields are localized in some finite region of space the integrals over
all 3-space at fixed time in some inertial frame of the components T 00 and
T 0i can be interpreted as the total energy and c times the total momentum
of the EM fields in that frame:
Z Z
1
T 00 d3 x = E2 + B2 d3 x = Efield


(7.68)
Z Z
1
T i0 d3 x = (E × B)i d3 x = cPfield
i

The previous definitions of field energy and momentum densities suggests


that there should be a covariant generalization of the differential conservation
law (2.75) of Poynting’s theorem. That is:

∂α T αβ = 0 (7.69)
7.8 Conservation Laws : Symmetric Stress Tensor 129

Consider
 
αβ
X ∂L
∂α T = ∂α ∂ φk − ∂ β L
β
∂(∂α φk )
k
X ∂L ∂L

β
= ∂α ∂ φk + ∂ α ∂ φk − ∂ β L
β
∂(∂α φk ) ∂(∂α φk )
k

but because of the equation of motion (7.50) the first term can be transformed
so that
X  ∂L ∂L

αβ β
∂α T = ∂ φk + ∂ (∂α φk ) − ∂ β L
β
∂φk ∂(∂α φk )
k
α
since L = L(φk , ∂ φk ) the term in the square bracket is an implicit differen-
tiation, hence
∂α T αβ = ∂ β L(φk , ∂ α φk ) − ∂ β L = 0

The conservation law (or continuity equation) (7.69) yields the conserva-
tion of total energy and momentum upon integration over all of 3-space at
fixed time Z Z Z
0 = ∂α T αβ d3 x = ∂0 T 0β d3 x + ∂i T iβ d3 x

If the fields are localized the 2nd integral (divergence) gives no contribution.
Then with the identification (7.68) we get

d d
Efield = 0 , Pfield = 0 (7.70)
dt dt
The results are valid for an observer at rest in the frame in which the fields
are specified.

7.8 Conservation Laws : Symmetric Stress Tensor


The canonical stress tensor T αβ has a number of deficiencies for example lack
of symmetry ! ( see T 0i and T i0 ). This affects the consideration of the angular
momentum of the field
Z
1
Lfield = x × (E × B)d3 x
4πc
The angular momentum density is expressed in terms of a 3rd-rank tensor

M αβγ = T αβ xγ − T αγ xβ (7.71)

Then as (7.69) implies conservation of energy and momentum (7.70) we expect


the vanishing of the 4-divergence
130 7 Dynamics of Relativistic Particles and EM Fields (**)

∂α M αβγ = 0 (7.72)

to imply conservation of the total angular momentum of the field.


Direct evaluation gives

0 = (∂α T αβ )xγ + T γβ − (∂α T αγ )xβ − T βγ

then by using (7.69) one can see that the conservation of angular momentum
requires that T αβ be symmetric. Finally, the involvement of the potentials in
(7.66) makes it not gauge invariant.
We will construct a symmetric, traceless, gauge-invariant stress tensor Θαβ
from the canonical stress tensor T αβ of (7.66).
We substitute ∂ β Aλ = −F λβ + ∂ λ Aβ and obtain
 
1 1 1 αµ
T αβ = g αµ Fµλ F λβ + g αβ Fµν F µν − g Fµλ ∂ λ Aβ (7.73)
4π 4 4π
The first term in (7.73) is symmetric and gauge invariant while the last term
of (7.73), with the help of the source-free Maxwell equations, can be written
(how?)

αβ 1 αµ 1 λα
TD ≡− g Fµλ ∂ λ Aβ = F ∂ λ Aβ
4π 4π
1 1
F λα ∂λ Aβ + Aβ ∂λ F λα = ∂λ F λα Aβ
 
= (7.74)
4π 4π
with the following properties: (why?)
Z
αβ 0β 3
(i) ∂α TD = 0, (ii) TD d x=0

Thus the differential conservation law (7.69) will hold for the difference (T αβ −
αβ
TD ) if it holds for T αβ .
If we define the symmetric stress tensor Θαβ
 
αβ 1 1
Θαβ = T αβ − TD = g αµ Fµλ F λβ + g αβ Fµν F µν (7.75)
4π 4

then the differential conservation law (7.69) will hold if it holds for T αβ . The
explicit components of Θαβ are:
1
Θ00 = E2 + B2


1 i
Θ0i = (E × B) (7.76)
4π  
1 1 ij
Θij i j i j 2 2

=− E E +B B − δ E +B
4π 2

The tensor Θαβ can be written in the schematic matrix form


7.9 Conservation Laws for EM fields interacting with Charged Particles 131
!
u cg
Θαβ = (M ) (7.77)
cg −Tij

where the time-time component is the energy density (2.73) the time-space
component is the momentum density (2.85) while the space-space compo-
nents are the negative of the Maxwell stress tensor (2.86).
The various covariant or mixed forms of the stress tensor are
! !
u −cg α u −cg
Θαβ = (M ) Θ β= (M )
−cg −Tij cg Tij
!
β u cg
Θα = (M )
−cg Tij
The differential conservation law

∂α Θαβ = 0 (7.78)

embodies Poynting’s theorem and conservation of momentum for free fields.


For example, for β = 0 we have
 
α0 1 ∂u
0 = ∂α Θ = +∇·S
c ∂t

where S = c2 g is the Poynting vector, this is the source-free form of (2.75).


Similarly for β = i
3
∂gi X ∂ (M )
0 = ∂α Θαi = − T
∂t j=1
∂xj ij

which is equivalent to (2.89).


The conservation of the field angular momentum defined by

Θαβγ = Θαβ xγ − Θαγ xβ (7.79)

is assured by (7.78) and the symmetry of Θαβ .

7.9 Conservation Laws for EM fields interacting with


Charged Particles
When external forces are present the Lagrangian for the Maxwell equations
is (7.52), the symmetric stress tensor for the EM field retains its form (7.75)
but the coupling to the external force makes its divergence non-vanishing.
132 7 Dynamics of Relativistic Particles and EM Fields (**)
 
αβ 1 µ λβ
 1 β µλ

∂α Θ = ∂ Fµλ F + ∂ Fµλ F
4π 4
 
1 µ λβ µ λβ 1 β µλ
= (∂ Fµλ )F + Fµλ ∂ F + Fµλ ∂ F
4π 2
the 1st term can be transformed via (7.56) and we get
1 1
∂α Θαβ + F βλ Jλ = Fµλ ∂ µ F λβ + ∂ µ F λβ + ∂ β F µλ

c 8π
the last two terms (in blue) can be transformed via the homogeneous Maxwell
equation ∂ µ F λβ + ∂ β F µλ + ∂ λ F βµ = 0 by −∂ λ F βµ = ∂ λ F µβ
1 1
∂α Θαβ + F βλ Jλ = Fµλ ∂ µ F λβ + ∂ λ F µβ

c 8π

The right-hand side is zero (why?) and thus the divergence of the stress
tensor is
1
∂α Θαβ = − F βλ Jλ (7.80)
c
The time and space components of this equation are
 
1 ∂u 1
+∇·S =− J·E (7.81)
c ∂t c
and
3  
∂gi X ∂ (M ) 1
− T = − ρEi + (J × B)i (7.82)
∂t j=1
∂xj ij c
These are the conservation of energy and momentum equations for EM fields
interacting with sources described by J α = (cρ, J).
The negative of the right hand side term in (7.80) is called the Lorentz
force density,
 
1 1 1
f β ≡ F βλ Jλ = J · E, ρE + J × B (7.83)
c c c
If the sources are a number of charged particles then the volume integral
of f β leads through the Lorents force equation (7.1) to the time rate of change
of the sum of the energies or momenta of all particles:
β
dPparticles
Z
f β d3 x =
dt
The conservation of the 4-momentum for the combined system of particle
and fields:
Z
d  β β

d3 x ∂α Θαβ + f β =

Pfield + Pparticles =0 (7.84)
dt
8
A Short Introduction to Tensor Analysis

8.1 Scalars and Vectors (****)


An n-dim manifold is a space M on every point of which we can assign n
numbers (x1 ,x2 ,...,xn ) - the coordinates - in such a way that there will be a
one to one correspondence between the points and the n numbers.
The manifold cannot be always covered by a single system of coordinates
and there is not a preferable one either.

The coordinates of the point P are connected by relations of  the form: 


0 0 0 0 0
xµ = xµ x1 , x2 , ..., xn for µ0 = 1, ..., n and their inverse xµ = xµ x1 , x2 , ..., xn


for µ = 1, ..., n. If there exist


0
0 ∂xµ ∂xν µ0
Aµ ν = and Aν µ0 = 0 ⇒ det |A ν| (8.1)
∂xν ∂xµ

then the manifold is called differential.


Any physical quantity, e.g. the velocity of a particle, is determined by a
set of numerical values - its components - which depend on the coordinate
system.
134 8 A Short Introduction to Tensor Analysis

Studying the way in which these values change with the coordinate system
leads to the concept of tensor.
With the help of this concept we can express the physical laws by tensor
equations, which have the same form in every coordinate system.
• ¡1-¿ Scalar field : is any physical quantity determined by a single numer-
ical value i.e. just one component which is independent of the coordinate
system (mass, charge,...)
• ¡2-¿ Vector field (contravariant): an example is the infinitesimal dis-
placement vector, leading from a point A with coordinates xµ to a neigh-
bouring point A0 with coordinates xµ + dxµ . The components of such a
vector are the differentials dxµ .

8.1.1 Vector Transformations

From the infinitesimal vector AA0 with components dxµ we can construct a
finite vector v µ defined at A. This will be the tangent vector of the curve
xµ = f µ (λ) where the points A and A0 correspond to the values λ and λ + dλ
of the parameter. Then
dxµ
vµ = (8.2)

Any transformation from xµ to x̃µ (xµ → x̃µ ) will be determined by n equa-
tions of the form: x̃µ = f µ (xν ) where µ , ν = 1, 2, ..., n.
This means that :
X ∂ x̃µ X ∂f µ
dx̃µ = ν
dxν = ν
dxν for ν = 1, ..., n (8.3)
ν
∂x ν
∂x

and
dx̃µ X ∂ x̃µ dxν X ∂ x̃µ
ṽ µ = = ν
= ν
vν (8.4)
dλ ν
∂x dλ ν
∂x

8.1.2 Contravariant and Covariant Vectors

Contravariant Vector: is a quantity with n components depending on the


coordinate system in such a way that the components aµ in the coordinate
system xµ are related to the components ãµ in x̃µ by a relation of the form
X ∂ x̃µ
ãµ = aν (8.5)
ν
∂xν

Covariant Vector: eg. bµ , is an object with n components which depend on


the coordinate system on such a way that if aµ is any contravariant vector,
the following sums are scalars
X X
bµ aµ = b̃µ ãµ = φ for any xµ → x̃µ [ Scalar Product] (8.6)
µ µ
8.2 Tensors: at last (****) 135

The covariant vector will transform as:


X ∂xν X ∂ x̃ν
b̃µ = bν or b µ = b̃ν (8.7)
ν
∂ x̃µ ν
∂xµ

What is Einstein’s summation convention?

8.2 Tensors: at last (****)


A contravariant tensor of order 2 is a quantity having n2 components T µν
which transforms (xµ → x̃µ ) in such a way that, if aµ and bµ are arbitrary
covariant vectors the following sums are scalars:

T λµ aµ bλ = T̃ λµ ãλ b̃µ ≡ φ for any xµ → x̃µ (8.8)

Then the transformation formulae for the components of the tensors of order
2 are (why?):

∂ x̃α ∂ x̃β µν α ∂ x̃α ∂xν µ ∂xµ ∂xν


T̃ αβ = T , T̃β = T ν & T̃αβ = Tµν
∂xµ ∂xν ∂xµ ∂ x̃β ∂ x̃α ∂ x̃β
The Kronecker symbol
(
λ 0 if λ 6= µ ,
δ µ =
1 if λ = µ .

is a mixed tensor having frame independent values for its components.


Tensors of higher order: T αβγ... µνλ...

8.2.1 Tensor algebra

Tensor addition : Tensors of the same order (p, q) can be added, their sum
being again a tensor of the same order. For example:
∂ x̃ν µ
ãν + b̃ν = (a + bµ ) (8.9)
∂xµ

Tensor multiplication : The product of two vectors is a tensor of order


2, because
∂ x̃α ∂ x̃β µ ν
ãα b̃β = a b (8.10)
∂xµ ∂xν
in general:

T µν = Aµ B ν or T µ ν = Aµ Bν or Tµν = Aµ Bν (8.11)
136 8 A Short Introduction to Tensor Analysis

• Contraction: for any mixed tensor of order (p, q) leads to a tensor of


order (p − 1, q − 1) (prove it!)

T λµν λα = T µν α (8.12)

• Symmetric Tensor : Tλµ = Tµλ orT(λµ) ,


Tνλµ = Tνµλ or Tν(λµ)
• Antisymmetric : Tλµ = −Tµλ or T[λµ] ,
Tνλµ = −Tνµλ or Tν[λµ]
Number of independent components :
Symmetric : n(n + 1)/2,
Antisymmetric : n(n − 1)/2

8.2.2 Tensors: Differentiation

We consider a region V of the space in which some tensor, e.g. a covariant


vector aλ , is given at each point P (xα ) i.e.

aλ = aλ (xα )

We say then that we are given a tensor field in V .


The simplest tensor field is a scalar field φ = φ(xα ) and its derivatives are
the components of a covariant tensor!
∂φ ∂xα ∂φ ∂φ
= we will use: = φ,α ≡ ∂α φ (8.13)
∂ x̃λ ∂ x̃λ ∂xα ∂xα
i.e. φ,α is the gradient of the scalar field φ.
The derivative of a contravariant vector field Aµ is :
∂Aµ
 µ 
∂ x̃ρ ∂
 µ 
∂ ∂x ν ∂x ν
Aµ ,α ≡ = Ã = Ã
∂xα ∂xα ∂ x̃ν ∂xα ∂ x̃ρ ∂ x̃ν
∂ 2 xµ ∂ x̃ρ ν ∂xµ ∂ x̃ρ ∂ Ãν
= Ã + (8.14)
∂ x̃ν ∂ x̃ρ ∂xα ∂ x̃ν ∂xα ∂ x̃ρ
Without the first term in the right hand side this equation would be the
transformation formula for a mixed tensor of order 2.

8.2.3 Tensors: Connections

The transformation (xµ → x̃µ ) of the derivative of a vector is:

∂xµ ∂ x̃ρ ν ∂ 2 xκ ∂ x̃ν σ


Aµ ,α = ν α
[Ã,ρ + σ x̃ρ ∂xκ
à ] (8.15)
∂ x̃ ∂x |∂ x̃ ∂{z }
ν
Γ̃σρ
8.2 Tensors: at last (****) 137

in another coordinate (xµ → x0µ ) we get again:

∂xµ ∂x0ρ 0ν
Aµ ,α = [A ,ρ + Γ 0ν σρ A0σ ] . (8.16)
∂x0ν ∂xα
Suggesting that the transformation (x̃µ → x0µ ) will be:

µ ∂ x̃µ ∂x0ρ 0ν
õ,α + Γ̃αλ Ãλ = (A ,ρ + Γ 0ν σρ A0σ ) (8.17)
∂x0ν ∂ x̃α
The necessary and sufficient condition for Aµ ,α to be a tensor is:

∂ 2 xµ ∂x0λ ∂xκ ∂xσ ∂x0λ µ


Γ 0λ ρν = 0ν 0ρ
+ Γ κσ . (8.18)
∂x ∂x ∂x µ ∂x0ρ ∂x0ν ∂xµ
Γ λ ρν is the called the connection of the space and it is not tensor.

8.2.4 Covariant Derivative

According to the previous assumptions, the following quantity transforms as


a tensor of order 2

Aµ ;α = Aµ ,α + Γ µ αλ Aλ or ∇α Aµ = ∂α Aµ + Γ µ αλ Aλ (8.19)

and is called covariant derivative of the contravariant vector Aµ .


In similar way we get (how?) :

φ;λ = φ,λ (8.20)


ρ
Aλ;µ = Aλ,µ − Γµλ aρ (8.21)
λµ
T ;ν = T ,ν + Γαν T αµ + Γαν
λµ λ µ
T λα (8.22)
λ
T µ;ν = T λ µ,ν + Γαν λ
T α µ − Γµνα λ
T α (8.23)
α α
Tλµ;ν = Tλµ,ν − Γλν Tµα − Γµν Tλα (8.24)
λµ···
T νρ··· ;σ = T λµ··· νρ··· ,σ
λ
+ Γασ T αµ··· νρ··· + Γασµ
T λα··· νρ··· + · · ·
α
− Γνσ T λµ··· αρ··· − Γρσ
α λµ···
T να··· − · · · (8.25)

8.2.5 Parallel Transport of a vector


λ
The connection Γµν helps is determining a vector Aµ = aµ + δaµ , at a point
0
P , which can be considered as “equivalent” to the vector aµ given at P .

∆aµ = aµ (P 0 ) − aµ (P ) = aµ (P ) + aµ,ν dxν − aµ (P ) = aµ,ν dxν


138 8 A Short Introduction to Tensor Analysis

Fig. 8.1.

aµ (P 0 ) − Aµ (P 0 ) = aµ + ∆aµ − (aµ + δaµ ) = ∆aµ − δaµ


| {z } | {z } | {z } | {z }
vector at point P at point P vector
ν λ
 ν λ
= aµ,ν dx − δaµ = aµ,ν − Cµν aλ dx i.e. δaµ = Cµν aλ dxν
| {z }
vector

δaµ = Γ λ µν aλ dxν for covariant vectors (8.26)


µ µ λ ν
δa = −Γ λν a dx for contravariant vectors (8.27)

The connection Γ λ µν allows to define the transport of a vector aλ from a


point P to a neighbouring point P 0 (Parallel Transport).

•The parallel transport of a scalar field is zero! δφ = 0 (why?)

8.2.6 Curvature Tensor (**)

The ”trip” of a parallel transported vector along a closed path The parallel
transport of the vector aλ from the point P to A leads to a change of the
vector by a quantity Γ λ µν (P )aµ dxν and the new vector is:

aλ (A) = aλ (P ) − Γµν
λ
(P )aµ dxν (8.28)

A further parallel transport to the point B will lead the vector

aλ (B) = aλ (A) − Γ λ ρσ (A)aρ (A)δxσ


= aλ (P ) − Γ λ µν (P )aµ dxν − Γ λ ρσ (A) aρ (P ) − Γ ρ βν (P )aβ dxν δxσ
 

Since both dxν and δxν assumed to be small we can use the following
expression
8.2 Tensors: at last (****) 139

Fig. 8.2.

Γ λ ρσ (A) ≈ Γ λ ρσ (P ) + Γ λ ρσ,µ (P )dxµ . (8.29)


λ
Thus we have estimated the total change of the vector a from the point P
to B via A (all terms are defined at the point P ).
ρ β
aλ (B) = aλ − Γµν
λ µ
a dxν − Γρσ
λ ρ
a δxσ + Γ λ ρσ Γβν a dxν δxβ + Γρσ,τ
λ
aρ dxτ δxσ
− Γ λ ρσ,τ Γ ρ βν aβ dxτ dxν δxσ

If we follow the path P → C → B we get:

aλ (B) = aλ −Γ λ µν aµ δxν −Γ λ ρσ aρ dxσ +Γ λ ρσ Γ ρ βν aβ δxν dxβ +Γ λ ρσ,τ aρ δxτ dxσ

and the effect on the vector will be


 
ρ
δaλ ≡ aλ (B) − aλ (B) = aβ (dxν δxσ − dxσ δxν ) Γ λ ρσ Γβν + Γ λ βσ,ν (8.30)

By exchanging the indices ν → σ and σ → ν we construct a similar relation

δaλ = aβ (dxσ δxν − dxν δxσ ) Γ λ ρν Γ ρ βσ + Γ λ βν,σ



(8.31)

and the total change will be given by the following relation:


1
δaλ = − aβ Rλ βνσ (dxσ δxν − dxν δxσ ) (8.32)
2
where

Rλ βνσ = −Γ λ βν,σ + Γ λ βσ,ν − Γ µ βν Γ λ µσ + Γ µ βσ Γ λ µν (8.33)

is the curvature tensor.


140 8 A Short Introduction to Tensor Analysis

8.3 Geodesics (***)


• For a vector uλ at point P we apply the parallel transport along a curve
on an n-dimensional space which will be given by n equations of the form:
xµ = f µ (λ); µ = 1, 2, ..., n
µ
• If uµ = dx
dλ is the tangent vector at P the parallel transport of this vector
will determine at another point of the curve a vector which will not be in
general tangent to the curve.
• If the transported vector is tangent to any point of the curve then this
curve is a geodesic curve of this space and is given by the equation :
duρ
+ Γ ρ µν uµ uν = 0 . (8.34)

• Geodesic curves are the shortest curves connecting two points on a curved
space.

8.4 Metric Tensor (****)


A space is called a metric space if a prescription is given attributing a scalar
distance to each pair of neighbouring points
The distance ds of two points P (xµ ) and P 0 (xµ + dxµ ) is given by
2 2 2
ds2 = dx1 + dx2 + dx3 (8.35)
In another coordinate system, x̃µ , we will get
∂xν α
dxν = dx̃ (8.36)
∂ x̃α
which leads to:
ds2 = g̃µν dx̃µ dx̃ν = gαβ dxα dxβ . (8.37)
This gives the following transformation relation (why?):
∂xα ∂xβ
g̃µν = gαβ (8.38)
∂ x̃µ ∂ x̃ν
suggesting that the quantity gµν is a symmetric tensor, the so called metric
tensor.
The relation (8.37) characterises a Riemannian space: This is a metric
space in which the distance between neighbouring points is given by (8.37).
• If at some point P there are given 2 infinitesimal displacements d(1) xα
and d(2) xα , the metric tensor allows to construct the scalar
gαβ d(1) xα d(2) xβ
which shall call scalar product of the two vectors.
• Properties:
gµν Aµ = Aν , gµν T µα = T α ν , gµν T µ α = Tαν , gµν gασ T µα = Tνσ
8.4 Metric Tensor (****) 141

• Metric element for Minkowski spacetime


ds2 = −dt2 + dx2 + dy 2 + dz 2 (8.39)
2 2 2 2 2 2 2 2
ds = −dt + dr + r dθ + r sin θdφ (8.40)
• For a sphere with radius R :
ds2 = R2 dθ2 + sin2 θdφ2

(8.41)
• The metric element of a torus with radii a and b
2
ds2 = a2 dφ2 + (b + a sin φ) dθ2 (8.42)
• The contravariant form of the metric tensor:
1
gµα g αβ = δµβ where g αβ = Gαβ ← minor determinant (8.43)
det |gµν |
With g µν we can now raise lower indices of tensors
Aµ = g µν Aν , T µν = g µρ T ν ρ = g µρ g νσ Tρσ (8.44)
• The angle, ψ, between two infinitesimal vectors d(1) xα and d(2) xα is:

gαβ d(1) xα d(2) xβ


cos(ψ) = p p . (8.45)
gρσ d(1) xρ d(1) xσ gµν d(2) xµ d(2) xν

8.4.1 The Determinant of gµν (**)

The quantity g ≡ det |gµν | is the determinant of the metric tensor. The deter-
minant transforms as :
∂xα ∂xβ
   α  β
∂x ∂x
g̃ = det g̃µν = det gαβ µ ν = det gαβ · det · det
∂ x̃ ∂ x̃ ∂ x̃µ ∂ x̃ν
2
∂xα
 
= det µ g = J 2g (8.46)
∂ x̃
where J is the Jacobian of the transformation.
This relation can be written also as:
p p
|g̃| = J |g| (8.47)
p
i.e. the quantity |g| is a scalar density of weight 1.
• The quantity
p p
|g|δV ≡ |g|dx1 dx2 . . . dxn (8.48)
is the invariant volume element of the Riemannian space.
• If the determinant vanishes at a point P the invariant volume is zero
and this point will be called a singular point.
142 8 A Short Introduction to Tensor Analysis

8.4.2 Christoffel Symbols (****)


In Riemannian space there is a special connection derived directly from the
metric tensor. This is based on a suggestion originating from Euclidean ge-
ometry that is :
“ If a vector aλ is given at some point P , its length must remain unchanged
under parallel transport to neighboring points P 0 ”.

Fig. 8.3.

|a|2P = |a|2P 0 or gµν (P )aµ (P )aν (P ) = gµν (P 0 )aµ (P 0 )aν (P 0 ) (8.49)


0 ρ
Since the distance between P and P is dx we can get
gµν (P 0 ) ≈ gµν (P ) + gµν,ρ (P )dxρ (8.50)
µ 0 µ µ
a (P ) ≈ a (P ) − Γσρ (P )aσ (P )dxρ (8.51)

By substituting these two relation into equation (8.49) we get (how?)


σ σ
 µ ν ρ
gµν,ρ − gµσ Γνρ − gσν Γµρ a a dx = 0 (8.52)
This relations must by valid for any vector aν and any displacement dxν
which leads to the conclusion the the relation in the parenthesis is zero. Closer
observation shows that this is the covariant derivative of the metric tensor !
σ σ
gµν;ρ = gµν,ρ − gµσ Γνρ − gσν Γµρ = 0. (8.53)
i.e. gµν is covariantly constant.
This leads to a unique determination of the connections of the space (Rie-
mannian space) which will have the form (why?)

α 1 αν
Γµρ = g (gµν,ρ + gνρ,µ − gρµ,ν ) (8.54)
2
8.5 Geodesics in a Riemann Space (***) 143

and will be called Christoffel Symbols .


It is obvious that Γ α µρ = Γ α ρµ .

8.5 Geodesics in a Riemann Space (***)


The geodesics of a Riemannian space have the following important property.
If a geodesic is connecting two points A and B is distinguished from the
neighboring lines connecting these points as the line of minimum or maximum
length. The length of a curve, xµ (s), connecting A and B is:
B B µ ν 1/2
Z Z  
α dx dx
S= ds = gµν (x ) ds
A A ds ds

Fig. 8.4.

A neighboring curve x̃µ (s) connecting the same points will be described
by the equation:
x̃µ (s) = xµ (s) +  ξ µ (s) (8.55)
where ξ µ (A) = ξ µ (B) = 0. The length of the new curve will be:
B µ ν 1/2
Z  
α dx̃ dx̃
S̃ = gµν (x̃ ) ds (8.56)
A ds ds

For simplicity we set:


1/2 1/2
f (xα , uα ) = [gµν (xα )uµ uν ] and f˜(x̃α , ũα ) = [gµν (x̃α )ũµ ũν ]
µ
uµ = ẋµ = dxµ /ds and uµ = x̃˙ = dx̃µ /ds = ẋ + ξ˙µ (8.57)
144 8 A Short Introduction to Tensor Analysis

Then we can create the difference δS = S̃ − S 1


Z B Z B 
δS = δf ds = f˜ − f ds
A A
Z B   Z B  
∂f d ∂f α d ∂f α
= − ξ ds +  ξ ds
A ∂xα ds ∂uα A ds ∂uα

The last term does not contribute and the condition for the length of S to be
an extremum will be expressed by the relation:
Z B  
∂f d ∂f
δS =  − ξ α ds = 0 (8.58)
A ∂xα ds ∂uα
Since ξ α is arbitrary, we must have for each point of S:
 
d ∂f ∂f
µ
− =0 (8.59)
ds ∂u ∂xµ

Notice that the Langrangian of a freely moving particle with mass m = 2, is:
L = gµν uµ uν ≡ f 2 this leads to the following relations

∂f 1 ∂L −1/2 ∂f 1 ∂L −1/2
= L and = L (8.60)
∂uα 2 ∂uα ∂xα 2 ∂xα
and by substitution in (8.59) we come to the condition

 
d ∂L ∂L
− = 0. (8.61)
ds ∂uµ ∂xµ

Since L = gµν uµ uν we get:


∂L ∂uµ ν µ ∂u
ν
µ ν µ ν µ
= gµν u + gµν u = gµν δα u + gµν u δα = 2gµα u (8.62)
∂uα ∂uα ∂uα
∂L µ ν
= gµν,α u u (8.63)
∂xα

thus
d dgµα µ duµ duµ
(2gµα uµ ) = 2 u + 2gµα = 2gµα,ν uν uµ + 2gµα
ds ds ds ds
ν µ µ ν duµ
= gµα,ν u u + gνα,µ u u + 2gµα (8.64)
ds
1
We make use of the following relations:

!
α α α α α α α α α ∂f α ∂f 2
f (x̃ , ũ ) = f (x + ξ ,u + ξ̇ ) = f (x , u ) +  ξ + ξ̇ + O( )
∂xα ∂uα

! !
d ∂f α ∂f α d ∂f α
ξ = ξ̇ + ξ
ds ∂uα ∂uα ds ∂uα
8.5 Geodesics in a Riemann Space (***) 145

and by substitution in (8.61) we get

duµ 1
gµα + [gµα,ν + gαµ,ν − gµν,α ] uµ uν = 0
ds 2
if we multiply with g ρα we the geodesic equations
duρ ρ µ ν
+ Γµν u u = 0, or uρ ;ν uν = 0 (8.65)
ds
because duρ /ds = uρ ,µ uµ .

8.5.1 Euler-Lagrange Eqns vs Geodesic Eqns (***)

The Lagrangian for a freely moving particle is: L = gµν uµ uν and the Euler-
Lagrange equations:  
d ∂L ∂L
− =0
ds ∂uµ ∂xµ
are equivalent to the geodesic equation

duρ ρ µ ν d2 xρ µ
ρ dx dx
ν
+ Γµν u u = 0 or 2
+ Γµν =0
ds ds ds ds
Notice that if the metric tensor does not depend from a specific coordinate
e.g. xk then  
d ∂L
=0
ds ∂ ẋκ
which means that the quantity ∂L/∂ ẋκ is constant along the geodesic. Then
eq (8.62) implies that ∂∂L
ẋκ = gµκ u
µ
that is the κ component of the generalized
µ
momentum pκ = gµκ u remains constant along the geodesic.

8.5.2 Tensors: Geodesics (***)

If we know the tangent vector uρ at a given point of a known space we can


determine the geodesic curve. Which will be characterized as:
• timelike if |u|2 < 0
• null if |u|2 = 0
• spacelike if |u|2 > 0
where |u|2 = gµν uµ uν
If gµν 6= ηµν then the light cone is affected by the curvature of the space-
time.
For example, in a space with metric ds2 = −f (t, p x)dt2 + g(t, x)dx2 the light
cone will be drawn from the relation dt/dx = ± g/f which leads to STR
results for f , g → 1.
146 8 A Short Introduction to Tensor Analysis

Fig. 8.5.

8.5.3 Null Geodesics (***)

For null geodesics ds = 0 and the proper length s cannot be used to


parametrize the geodesic curves. Instead we will use another parameter λ
and the equations will be written as:

d2 xκ µ
κ dx dx
ν

2
+ Γµν =0 (8.66)
dλ dλ dλ
and obiously:
dxµ dxν
gµν = 0. (8.67)
dλ dλ

8.5.4 Geodesic Eqns & Affine Parameter (**)

In deriving the geodesic equations we have chosen to parametrize the curve


via the proper length s. This choice simplifies the form of the equation but
it is not a unique choice. If we chose a new parameter, σ then the geodesic
equations will be written:

d2 xµ α
µ dx dx
β
d2 σ/ds2 dxµ
2
+ Γαβ =− (8.68)
dσ dσ dσ (dσ/ds)2 dσ
where we have used
2
dxµ dxµ dσ d2 xµ d2 xµ dxµ d2 σ


= and = + (8.69)
ds dσ ds ds2 dσ 2 ds dσ ds2

The new geodesic equation (8.68), reduces to the original equation(8.66) when
the right hand side is zero. This is possible if
8.6 Riemann Tensor (***) 147

d2 σ
=0 (8.70)
ds2
which leads to a linear relation between s and σ i.e. σ = αs + β where α and
β are arbitrary constants. σ is called affine parameter.

8.6 Riemann Tensor (***)


When in a space we define a metric then is called metric space or Riemann
space. For such a space the curvature tensor

Rλ βνµ = −Γβν,µ
λ λ
+ Γβµ,ν σ
− Γβν λ
Γσµ σ
+ Γβµ λ
Γσν (8.71)

is called Riemann Tensor and can be also written as:


1
Rκβνµ = gκλ Rλ βνµ = (gκµ,βν + gβν,κµ − gκν,βµ − gβµ,κν )
2  
α ρ α ρ
+ gαρ Γκµ Γβν − Γκν Γβµ

Properties of the Riemann Tensor:

Rκβνµ = −Rκβµν , Rκβνµ = −Rβκνµ , Rκβνµ = Rνµκβ , Rκ[βµν] = 0

Thus in an n-dim space the number of independent components is:

n2 (n2 − 1)/12 (8.72)

For a 4-dimensional space only 20 independent components

8.6.1 The Ricci and Einstein Tensors (***)

The contraction of the Riemann tensor leads to Ricci Tensor

Rαβ = Rλ αλβ = g λµ Rλαµβ


µ µ µ ν µ ν
= Γαβ,µ − Γαµ,β + Γαβ Γνµ − Γαν Γβµ (8.73)

which is symmetric Rαβ = Rβα . Further contraction leads to the Ricci or


Curvature Scalar

R = Rα α = g αβ Rαβ = g αβ g µν Rµανβ . (8.74)

The following combination of Riemann and Ricci tensors is called Einstein


Tensor
1
Gµν = Rµν − gµν R (8.75)
2
148 8 A Short Introduction to Tensor Analysis

with the very important property:

 
1
Gµ ν;µ = Rµ ν − δ µ ν R = 0. (8.76)
2 ;µ

This results from the following important identity (Bianchi Identity)

Rλ µ[νρ;σ] = 0 (8.77)

8.6.2 Flat & Empty Spacetimes

• When Rαβµν = 0 the spacetime is flat


• When Rµν = 0 the spacetime is empty

Fig. 8.6.

Prove that :
aλ ;µ;ν − aλ ;ν;µ = −Rλ κµν aκ

8.7 Weyl Tensor (-)


Important relations can be obtained when we try to express the Riemann or
Ricci tensor in terms of trace-free quantities.
1
Sµν = Rµν − gµν R ⇒ S = S µ µ = g µν Sµν = 0 . (8.78)
4
8.8 Tensors : An example for parallel transport (****) 149

or the Weyl tensor Cλµνρ :


1
Rλµνρ = Cλµνρ + (gλρ Sµν + gµν Sλρ − gλν Sµρ − gµρ Sλν )
2
1
+ R (gλρ gµν − gλν gµρ ) (8.79)
12
1
Cλµνρ = Rλµνρ − (gλρ Rµν + gµν Rλρ − gλν Rµρ − gµρ Rλν )
2
1
+ R (gλρ gµν − gλν gµρ ) . (8.80)
6
and we can prove (how?) that :

g λρ Cλµνρ = 0 (8.81)

The Weyl tensor is called also conformal curvature tensor because it


has the following property :
• Consider besides the Riemannian space M with metric gµν a second Rie-
mannian space M̃ with metric

g̃µν = e2A gµν

where A is a function of the coordinates. The space M̃ is said to be conformal


to M .
• One can prove that
α α α α
Rβγδ 6= R̃βγδ while Cβγδ = C̃βγδ (8.82)

i.e. a “conformal transformation” does not change the Weyl tensor.


• It can be verified at once from equation (8.80) that the Weyl tensor
has the same symmetries as the Riemann tensor. Thus it should have 20
independent components, but because it is traceless [condition (8.81)] there
are 10 more conditions for the components therefore the Weyl tensor has 10
independent components.

8.8 Tensors : An example for parallel transport (****)


A vector A = A1 eθ + A2 eφ be parallel transported along a closed line on the
surface of a sphere with metric ds2 = dθ2 + sin2 dφ2 and Christoffel symbols
1 2
Γ22 = − sin θ cos θ and Γ12 = cot θ .
The eqns δA = −Γµν Aµ dxν for parallel transport will be written as:
α α

∂A1 1 ∂A1
= −Γ22 A2 ⇒ = sin θ cos θA2
∂x2 ∂φ
∂A2 2 ∂A2
= −Γ12 A1 ⇒ = − cot θA1
∂x2 ∂φ
150 8 A Short Introduction to Tensor Analysis

Fig. 8.7.

The solutions will be:


∂ 2 A1
= − cos2 θA1 ⇒ A1 = α cos(φ cos θ) + β sin(φ cos θ)
∂φ2
⇒ A2 = − [α sin(φ cos θ) − β cos(φ cos θ)] sin−1 θ

and for an initial unit vector (A1 , A2 ) = (1, 0) at (θ, φ) = (θ0 , 0) the integration
constants will be α = 1 and β = 0. The solution is:

sin(2π cos θ)
A = A1 eθ + A2 eφ = cos(2π cos θ)eθ − eφ
sin θ
i.e. different components but the measure is still the same
2 2
|A|2 = gµν Aµ Aν = A1 + sin2 θ A2
sin2 (2π cos θ)
= cos2 (2π cos θ) + sin2 θ =1
sin2 θ

Question : What is the condition for the path followed by the vector to
be a geodesic?
9
Physics on Curved Spaces

9.1 The electromagnetic (EM) force of a moving charge


(***)
In an inertial frame S, the 3-force on a particle of charge q moving in an EM
field is:
f = q (E + u × B)
This equation suggests that the force is proportional to the velocity and this
means that the extension of this equation in a 4-dim spacetime should be of
the form:
f = qF · u ⇒ fµ = qFµν uν (9.1)
where F (or Fµν ) is the electromagnetic field tensor, the scalar q is some
property of the particle that determines the strength of the EM force upon it
(i.e. the charge).
In order that the rest mass of a particle is not altered by the action of the EM
force we require the latter to be a pure force or rest-mass preserving. This
requires that u · f = 0 (why?) or

fµ uµ = qFµν uµ uν = 0

This implies that electromagnetic field tensor must be antisymmetric

Fµν = −Fνµ

9.2 The 4-current density (***)


• In order to construct the field equations of the theory which will determine
F(x) at any point in spacetime in terms of charges and currents we must find
a covariant way of expressing the source term.
152 9 Physics on Curved Spaces

• Let us consider some general time-dependent charge distribution. At


each event P of the spacetime we give the charge density ρ and 3-velocity u
as measured in some inertial frame e.g. the frame in which u = 0 at P .
• In this frame the proper charge density is given by ρ0 = qn0 , where
q is the charge of each particle and n0 is the number of particles in a unit
volume.

Fig. 9.1.

• In some other frame S 0 moving with speed v relative to S the volume


will be Lorentz contracted along the direction of motion hence the number
density of particles will be n0 = γv n0 and

ρ0 = γv ρ0

i.e the charge density is not a scalar but transforms like the component of a
vector.
Thus the obvious choice of a source term for the EM field equations should
be a 4-vector
j = ρ0 (x)u(x) and j · j = ρ20 c2
In an inertial frame S the components of the 4-current density j

j µ = ρ0 γu (c, u) = (cρ, j)

Thus we see that c2 ρ2 − j 2 is a Lorentz invariant.

9.3 The EM field equations (***)


The general form of the field equations will be

∇ · F = kj.

In Cartesian inertial coordinates xµ in some inertial frame S


9.4 Electromagnetism in the Lorenz gauge (***) 153

F µν ,µ = k j ν (9.2)

This equation can be used to find a law for conservation of charges

F µν ,µν = k j ν ,ν ⇒ j ν ,ν = 0 (why?)

In a 3-vector notation we may write this in a well known form:


∂ρ
+∇·j=0
∂t
which is similar to the non-relativistic equation of charge continuity, but we
should use the relativistic expressions for ρ and j.
—————————- Do we have a viable theory? ———————
The field equations are given by eqn (9.2). But Fµν has 6 independent
components and from eqn (9.2) we get only 4 field equations.
Our theory is underdetermined as it stands and this suggests that one can
use a 4-vector “potential” Aµ to construct Fµν . In Cartesian coordinates we
may write:
Fµν = Aµ,ν − Aν,µ (9.3)
i.e. Fµν is antisymmetric by construction and contains ONLY 4 indepen-
dent fields Aµ .
Thus we can write the field equations (9.2) in terms of the 4-vector poten-
tial Aµ
η µσ (Aλ,σµ − Aσ,λµ ) = k jλ (9.4)
Alternatively, we can express electromagnetism entirely in terms of the EM
field tensor and in this case, we require the 2 field equations:

F µν ,µ = k j µ (9.5)
F[µν,σ] ≡ Fµν,σ + Fσµ,ν + Fνσ,µ = 0 (9.6)

The last equation can be easily derived from equation (9.3) (how?).
Finally, the constant k can be found by demanding consistency with the
standard Maxwell equations. In SI units we have k = µ0 , where 0 µ0 = 1/c2 .

9.4 Electromagnetism in the Lorenz gauge (***)


We can always add an arbitrary 4-vector Qµ to the 4-potential Aµ . Thus in
Cartesian inertial coordinates xµ we have

A(new)
µ = Aµ + Qµ (9.7)

This is not a coordinate transformation. The new EM field tensor will be:
(new)
Fµν = A(new)
ν,µ
(new)
− Aµ,ν = Aν,µ − Aµ,ν + Qν,µ − Qµ,ν (9.8)
154 9 Physics on Curved Spaces

and the original EM field tensor will be recovered if Qν,µ = Qµ,ν .


This is possible if Qµ is the gradient of a scaler field i.e. Qµ = ψ,µ .
This is called gauge freedom of the theory i.e. we are free to add the gradient
of any scalar field ψ to the 4-vector potential Aµ i.e.

A(new)
µ = Aµ + ψ,µ (9.9)

and still recover the same EM field tensor and the same EM field equations.
This is an example of a gauge transformation, which is not a coordinate
transformation.
In the field equations

η µσ (Aλ,σµ − Aσ,λµ ) = µ0 jλ (9.10)

the 2nd term in the left hand side can be written as Aµ ,µλ thus this term can
become zero by choosing a scalar field ψ such that

Aµ ,µ = 0 (9.11)

This condition is called Lorentz gauge1 and simplifies the EM field equations

η µσ Aλ,σµ = Aλ ,µ ,µ = Aλ = µ0 jλ (9.12)

• The last 2 equations are the EM field equations in the Lorentz gauge.
• In the absence of charges and currents Aµ has wave solutions traveling
with the speed of light.
• This is also true for the components of Fµν since in this case we also have
Fµν = 0.

9.5 Electric and Magnetic Fields in inertial frames (***)


It is useful to define the components of F µν and Aµ in terms of the familiar
electric and magnetic 3-vector fields E and B in some inertial frame S. Thus
the components of Aµ can be written in terms of the electrostatic potential φ
and the 3-dimesional vector potential A:

Aµ = (φ/c, A)

Then the Lorentz gauge condition becomes (why?):


1 ∂φ
∇·A+ =0
c ∂t
2
and in this gauge, the field equations become
1
Prove that this condition is preserved by any further gauge transformation Aµ →
Aµ + ψ,µ if and only if η νµ ψ,νµ ≡ ψ,µ ,µ = 0.
2
0 : permittivity of free space, µ0 magnetic permeability of free space, both are
universal constants. SI UNITS: henries per meter, or newtons per ampere squared.
9.6 Electromagnetism in arbitrary coordinates (***) 155
ρ
A = µ0 j and φ = (9.13)
0
In terms of φ and A, the electric and magnetic fields in S are given by:
∂A
B=∇×A and E = −∇φ − (9.14)
∂t

These equations lead to Maxwell equations in the familiar form (prove


it):

ρ ∂B
∇·E = , ∇×E=− (9.15)
0 ∂t
∂E
∇ · B = 0, ∇ × B = µ0 j + µ0 0 (9.16)
∂t
Finally, by using eqns (9.3) and (9.14) we can find (how?) that the covariant
components of F µν in S are:

E 1 /c E 2 /c E 3 /c
 
0
 −E 1 /c 0 −B 3 B 2 
F µν =  −E 2 /c B 3

0 −B 1 
−E 3 /c −B 2 B 1 0

The corresponding electric and magnetic fields E0 and B0 in some other


Cartesian inertial frame S 0 are most easily obtained by calculating the com-
ponents of the EM field tensor F µν or the 4-potential Aµ in this frame.

9.6 Electromagnetism in arbitrary coordinates (***)


The EM field equations in Cartesian inertial coordinates are given by:

F µν ,µ = µ0 j µ (9.17)
F[µν,σ] ≡ Fµν,σ + Fσµ,ν + Fνσ,µ = 0 (9.18)

In such a coordinate system, the partial derivative is identical to the covariant


derivative and we can rewrite the equations as:

F µν ;µ = µ0 j µ (9.19)
F[µν;σ] ≡ Fµν;σ + Fσµ;ν + Fνσ;µ = 0 (9.20)

These new equations are fully covariant and if they are valid in one system of
coordinates then they will be valid in all coordinate systems.
In the same way we can write the EM field equations in terms of the
4-potential Aµ

η µσ (Aλ,σµ − Aσ,λµ ) = k jλ ⇒ g µσ (Aλ;σµ − Aσ;λµ ) = µ0 jλ (9.21)


156 9 Physics on Curved Spaces

which is again a fully covariant equation.


Gauge transformations are also permitted in arbitrary coordinates:
Aµ(new) = Aµ + ψ,µ ≡ Aµ + ψ;µ (9.22)
The Lorentz gauge condition in abritrary coordinates will be written as:
Aµ ;µ = 0
and in this case the EM field equations will be written as:
Aµ = µ0 jµ
where the d’Alembertian operator is:  = g µν ∇µ ∇ν and in vacuuo we may
again have:
Aµ = 0 and Fµν = 0
Finally, the charge conservation in arbitrary coordinates will be expressed
as:
j µ ;µ = 0 .
Obviously, the components of F µν and Aµ in two different arbitrary coordinate
systems xµ and x0µ are related by:
∂x0µ ν ∂x0µ ∂x0µ σρ
A0µ = A and F 0µν = F (9.23)
∂xν ∂xσ ∂xρ

9.7 Equations of motion for charged particles (***)


The equation of motion of a charged particle in an EM field in a coordinate
invariant form is:
dp du
= m0 = qF·u
dτ dτ
where m0 is the rest mass of the particle, p is its 4-momentum, u is its 4-
velocity and τ is the proper time measured along its worldline.
In Cartesian inertial coordinates this will be written
duµ
m0 = q F µ ν uν

in an arbitrary coordinate system the derivative in left hand side must be
replaced with the absolute (or intrinsic) derivative:
Duµ
 µ 
du µ ν σ
m0 = m0 + Γνσ u u = q F µ ν uν
Dτ dτ
where uµ = dxµ /dτ , since the 4-velocity is tangent to the particle’s worldline
xµ (τ ). In arbitrary coordinates will be:
d2 xµ ν
µ dx dx
σ
q µ dx
ν
+ Γ νσ = F ν (9.24)
dτ 2 dτ dτ m0 dτ
9.8 Energy-Momentum Tensor(***) 157

Conclusion
The general procedure for converting an equation valid in Cartesian inertial
coordinates into one that is valid in an arbitrary coordinate system is as
follows:
• replace partial with covariant derivatives
• replace ordinary derivatives along curves with absolute (intrinsic) deriva-
tives
• replace ηµν with gµν .

9.8 Energy-Momentum Tensor(***)


The energy momentum tensor T µν in Cartesian coortinates will be
 
1 1
T µν = F µα F νβ ηαβ − η µν Fαβ F αβ (9.25)
4π 4

The explicit form of the various components is:


1
T 00 = |E|2 + |B|2


1 1 i
T 0i δ ij εjkm E k B m =

= (E × B)
4π 4π
1 
T ij |E|2 + |B|2 δ ij − 2 E i E j + B i B j
 
= (9.26)

• The component T 00 is the electromagnetic energy density
• The component T 0i = T i0 gives the momentum flux density of the EM
field, referred as the Poynting vector.
c
P= (E × B)

• The spacelike diagonal component with i = j
1 
T ii = |E|2 + |B|2 − 2 E i E i + B i B i
 

gives the pressure normal to the face of a small cube of EM field
• The off-diagonal spacelike component with i 6= j
1
T ij = − EiEj + BiBj


gives the shear stresses that act parallel to the faces of an infinitesimal
region of the EM field
158 9 Physics on Curved Spaces

EXAMPLE: The EM stress energy tensor for a constant electric field in the
x direction will be (prove it):
E2 E2 E2
T 00 = , T xx = − , T yy = T zz =
8π 8π 8π
0
what will be the components of this tensor in a frame S moving with a velocity
v at the x-direction?
To construct the gravitational field equations, we must find a covariant
way of expressing the source term.
We must find a tensor that describes the matter distribution at each event
in spacetime.
• Let’s start with dust, this is a time-dependent distribution of electrically
neutral non-interacting particles, each of rest mass m0 .
• At each event P of the spacetime we can characterize the distribution
completely by giving the matter density ρ and 3-velocity u as measured in
some inertial frame.
• For example in an instantaneous rest frame S at a given point P , u = 0
while the proper density is given by ρ0 = m0 n0 , where m0 is the rest mass of
each particle and n0 is the number of particles in the unit volume.

• In a moving frame S 0 , the volume containing a fixed number of particles


is Lorentz contracted along the direction of the motion thus n0 = γv n and the
mass of each particle in S 0 is m0 = γv m0 thus the matter density in S 0 is
ρ0 = γv2 ρ0
• This means that the matter density is not a scalar but does transform as
the component of rank-2 tensor. This suggests that the source term in the
gravitational field equations should be a rank-2 tensor
T µν = ρ0 uµ uν (9.27)
In a local Cartesian coordinate frame the components of the 4-velocity for the
fluid is uµ = γu (c, u). In this frame the components of the enrgy-momentum
tensor are:
9.8 Energy-Momentum Tensor(***) 159

Energy - Momentum Tensor

T 00 = ρu0 u0 = γu2 ρc2 energy density of the particles


T 0k = ρ0 u0 uk = γu2 ρcuk energy flux ×c−1 in the k-direction
k0
T = ρ0 u u = γu2 ρcuk
0 k
momentum density ×c in the k-direction
ij
T = ρui uj = γu2 ρui uj the rate of flow of the i-component
of momentum per unit area in the j-direction

Due to these identifications the tensor T is known as the energy - momentum


or stress - energy tensor.

9.8.1 The Energy-Momentum tensor of a perfect fluid(**)

A perfect fluid is defined as one for which there are no forces between the
particles and no heat conduction or viscosity. Thus in an inertial reference
frame the components of T for a perfect fluid are given by:
 2 
ρc 0 0 0
 0 p 0 0
T µν = 
 0 0 p 0

0 00p

while we can show that for inertial observers (or in STR)

T µν = (ρ + p/c2 )uµ uν + η µν p (9.28)

while in an arbitrary coordinate system it will be:

T µν = (ρ + p/c2 )uµ uν + g µν p (9.29)

Note : the T µν is symmetric and is made up from 2 scalar fields ρ and p


and the vector field u that uniquely characterize the perfect fluid.

9.8.2 The Energy-Momentum tensor of a “real fluid” (*)

For the general real fluid we have to take into account that:
• besides the bulk motion of the fluid, each particle has some random (ther-
mal) velocity

• there may be various forces between the particles that contribute potential
energies to the total.
In this case the components of the energy momentum tensor will be:
• T 00 : is the total energy density, including any potential energy contribu-
tions from forces between the particles and kinetic energy from the random
thermal motions.
160 9 Physics on Curved Spaces

• T 0i : although there is NO bulk motion, energy might transmitted by


heat conduction i.e. it is a heat conduction term in the inertial frame of
reference
• T i0 : if heat is being conducted then the energy will carry momentum
• T ij : the random thermal motions of the particles will give rise to momen-
tum flow, so that T ii is the isotropic pressure in the i-direction and T ij
(i 6= j) are the viscous stresses in the fluid.

9.9 Conservation of Energy and Momentum for a perfect


fluid (**)
By analogy with equation j µ ;µ = 0 for the conservation of charge, the conser-
vation of energy and momentum is represented by

T µν ,ν = 0 (9.30)

We will show that this relation leads to well known equations of motion and
continuity for a fluid in the Newtonian limit. Substituting eqn (9.28) into
equation (9.30) we get:

ρ + p/c2 ,ν uµ uν + ρ + p/c2 [(uµ ,µ )uν + uµ (uν ,µ )] − η µν p,µ = 0 (9.31)


 

from the normalization condition for the 4-velocity uν uν = c2 by differen-


tiation we get the relativistic equation of continuity for a perfect fluid in
local inertial coordinates at a point P :

(ρuµ )µ + (p/c2 )uµ µ = 0 (9.32)

and equation (9.31) becomes the the relativistic equation of motion for
a perfect fluid in local inertial coordinates at some point P :

ρ + p/c2 (uµ ,µ )uν = η µν − uµ uν /c2 p,µ


 
(9.33)

A slowly moving fluid is the one for which we may neglect u/c and so take
γu ≈ 1 and uµ ≈ (c, u) and the proper density becomes the “normal” density.
In this limit we get
∂ρ
T 0ν ,ν = 0 → eqn (9.32) ⇒ + ∇ · (ρu) = 0 (9.34)
∂t
which is the classical equation of continuity for a fluid, and
∂u 1
T jν ,ν = 0 → eqn (9.33) ⇒ + (u · ∇) u = − ∇p (9.35)
∂t ρ
which Euler’s equations of motion for a perfect fluid.
9.10 Linear Field Equations for Gravitation (**) 161

————–
We can easily obtain the condition for energy and momentum conservation
in arbitrary coordinates by replacing the partial with the covariant derivative,
and we get:
T µν ;ν = 0 (9.36)
This is a fundamental equation which imposes tight restrictions on the pos-
sible forms that the gravitational field equations may take.

9.10 Linear Field Equations for Gravitation (**)

• In the linear approximation for the gravitational field, we neglect the


effects of the gravitational field on itself. 3
• The approximation applies to all phenomena that lie in the region of
overlap between Newton’s and Einstein’s theories.
• Although, it is true that the most spectacular results of gravitational
theory depend in a crucial way on the nonlinearity of the field equations,
almost all of the results that have been the subject of experimental investi-
gation can be described by the linear approximation.
• The deflection of the light, the retardation of the light, the grav. time
dilation and gravitational radiation emerge in the linear approximation.
• It will be instructive if we follow the same path as for constructing
the Electromagnetic field equations which are linear. The most general form
of the EM field equations will have the form:

∂µ ∂ µ Aν + b∂ ν ∂µ Aµ + aAν = κj ν (9.37)

where a, b and κ(= 4π) are constants.


• We will assume that our spacetime is the flat spacetime of STR and
we look for a suitable field equation for the gravitational field.
• The first thing that we should decide is the source of the gravitation,
and from Newtonian theory we know that the energy density is the best
candidate. However, it is impossible to construct a Lorentz invariant theory
of gravitation. As we have seen earlier, the energy density is just a component
of the the energy density tensor T µν and if in one system of coordinates there
exists only one component the T 00 this will not be true in any system. This
means that the whole T µν will be the source of the gravitation field. 4
• Since the source of gravitation is a 2nd-rank symmetric tensor the
obvious choice is a 2nd rank symmetric field tensor lets say hµν . Then the
most general field equation, linear in hµν is of 2nd differential order and
contains the T µν as source term:
3
From ”Gravitation & Spacetime” H. Ohanian & R. Ruffini, W.W. Norton (1994)
4
Can you find a good reason why the trace of T µν alone cannot be the source of
the gravitational field?
162 9 Physics on Curved Spaces
 ,ν ,µ

hµν ,λ ,λ + αh,µν − β hµλ ,λ + hνλ ,λ + γη µν h,λ ,λ + δη µν hλσ ,λσ = −κT µν
(9.38)
where α, β, γ , δ and κ are arbitrary constants.
• If we operate on both sides with ∂/∂xµ the right hand side should
vanish due to energy-momentum conservation (T µν ,µ = 0).
• Then by rearranging the terms on the left hand side we get the
following set of equations for the constants: −β + δ = 0, 1 − β = 0 and
α + γ = 0. (How?)
• Thus β = δ = 1 & γ = −α and the field equations will be written as
 ,ν ,µ

hµν ,λ ,λ + αh,µν − hµλ ,λ + hνλ ,λ − αη µν h,λ ,λ + η µν hλσ ,λσ = −κT µν
(9.39)
even α can be be eliminated if we introduce a new tensor of the form hµν =
h̃µν − Cη µν h̃ by choosing C = (1 − α)/(2 − 4α) (prove it?).
• Thus the relativistic field equation for gravitation in the linear ap-
proximation becomes:
 ,ν ,µ

hµν ,λ ,λ + h,µν − hµλ ,λ + hνλ ,λ − η µν h,λ ,λ + η µν hλσ ,λσ = −κT µν (9.40)

while κ is the coupling constant of the theory and will be estimated from the
Newtonian limit of the above equation.
• We will call hµν gravitational tensor potential in analogy to EM
potential Aµ .
• Besides the tensor field hµν one may consider extra scalar, vector
or even tensor fields and to construct endlessly complicated theories.
• Jordan (1951) and Brans-Dicke (1961) suggested the least complicated
scalar-tensor theory containing only an extra scalar field. This new scalar field
has the effect of making the gravitational constant dependent on position. The
experimental evidence (up to now) speaks against the scalar-tensor theory
• In EM we got a simplified field equation by taking advantage of the
gauge invariance. The same can be done for gravitation and it can be shown
that the field equations are invariant under the gauge transformation

hµν
(new) = h
µν
+ Qµ,ν + Qν,µ (9.41)

where Q is an arbitrary vector field.


• In analogy to EM (Lorentz gauge) using the gauge condition known
as Hilbert gauge5  
µν 1 µν
h − η h =0 (9.42)
2 ,µ

If the Hilbert gauge condition holds; the field equations simplifies to:
5
Show that if Qν is solution of the differential equation Qν ,µ ,µ = −hµν ,µ + 12 h,ν
then hµν
(new) will satisfy the Hilbert condition.
9.11 Interaction of Gravitation and Matter (*) 163
  ,λ
1
hµν − η µν h = −κT µν (9.43)
2 ,λ

By substituting
1
φµν = hµν − η µν h (9.44)
2
the field equation becomes

φµν = φµν ,λ ,λ = −κT µν (9.45)

and the gauge condition


φµν ,µ = 0 (9.46)

9.11 Interaction of Gravitation and Matter (*)


In writing field eqn (9.45) we assumed that T µν was the energy - momentum
tensor of matter. I.e in order to obtain the linear field equations we left out
the effect of the gravitational field upon itself.
Due to this our linear field equations have two drawbacks:
• According to eqn (9.45) matter acts on the gravitational fields and changes
the fields, but there is no reciprocal action of the gravitational fields on
matter, and the energy-momentum of matter is conserved T µν ,ν = 0 ! This
is an incosistency.
• Gravitational energy does not act as source of gravitation, in contradiction
to the principle of equivalence.
Thus eqn (9.45) is only an approximation for weak gravitational fields.
The way to correct is to include the energy-momentum of the gravitational
field in T µν i.e.
µν
T µν = T(m) + tµν (9.47)
µν
where T(m) and tµν are the energy momentum tensors of of matter and grav-
itation.
The field equations become:
 
φµν = φµν ,λ ,λ = −κ T(m)µν + tµν (9.48)

If we act on both side with ∂/∂xν we get (because of the Hilbert gauge)
 
T(m)µν + tµν =0 (9.49)

which expresses the conservation of the total energy-momentum, and raises


the inconsistencies mentioned earlier.
164 9 Physics on Curved Spaces

But, tµν is still unknown and we can find it only approximately.


If we assume a solution of the above eqn free of sources i.e. φµν ,λ ,λ = 0
from field theory we get a first order approximation to the energy-momentum
tensor of the gravitational field
  
1 1
t(1) µν = 2φαβ,µ φαβ ,ν − φ,µ φ,ν − η µν φαβ,σ φαβ,σ − φ,σ φ,σ (9.50)
4 2
Thus the new approximate form of the grav. field equations will be:
φµν = φµν ,λ ,λ = −κ T µν + t(1) µν

(9.51)

The corresponding equation for the conservation of the total energy-


momentum is
T(m) µν + t(1) µν ,ν ≈ 0

(9.52)
this can be used to derive the equation of motion.
Thus by expanding the above equation we get
1  αβ,µ  κ
t(1) µν ,ν = φαβ ,ν,ν − φ,µ φ,ν ,ν ≈ − 2φαβ,µ T(m)αβ − φ,µ T(m) +. . .


4 4
(9.53)
we can obtain (how?):
 
µν κ αβ,µ 1 αβ ,µ
T(m) ,ν − T(m)αβ φ − η φ =0 (9.54)
2 2
or
κ
T(m) µ ν,ν − hαβ,µ T(m)αβ = 0 (9.55)
2
This equation tells how much energy and momentum the gravitational field
transfers to the matter on which it acts.
In other words it determines the rate at which the momentum of a particle
changes and therefore it determines the equation of motion of a particle
acted upon a gravitational field.
To derive this equation of motion we integrate equation (9.55) over the
volume of the “particle” (a small system of finite size).
The 1st term gives:
Z
d d
T(m)µ0 d3 x = Pµ (9.56)
dt dt
For the 2nd term we should remember that:
T(m)αβ = ρ0 uα uβ + . . . (9.57)

then6 Z
κ κp
− hαβ,µ T(m)αβ d3 x = − 1 − u2 mhαβ,µ uα uβ (9.58)
2 2
Thus we finally get (how?)
6

ρ0 uα uβ d3 x = muα uβ 1 − u2
R
Show that
9.12 Local Cartesian coordinates(-) 165

d κ
Pµ = − mhαβ,µ uα uβ
dτ 2
which has the right form for an equation of motion
If we assume that :
∂L κ ∂L
= − mhαβ,µ uα uβ and = Pµ (9.59)
∂xµ 2 ∂uµ
we get the Euler -Lagrange form of equations of motion.
Actually, in the absence of forces the Lagrangian will have the form
1
L0 = m ηαβ uα uβ
2
while in the presence of a field becomes:
1  
L= m ηαβ + κh̃αβ uα uβ
2

notice that h̃αβ is a quantity related to the spacetime while hαβ is the ”tensor
potential” of our theory. But because of the Euler-Lagrange equations these
two quantities are identical and this suggests that a relation between the
“geometrical” quantity h̃αβ and the “ physical” quantity” hαβ !
NOTE 1 : We have created a relation between physical and geometrical
quantities
Alternatively, if in the geodesic equations one assumes that gµν = ηµν +
κhµν then it comes to the equation

d κ
(uµ + κhµα uα ) − hαβ,µ uα uβ = 0 . (9.60)
dτ 2
NOTE 2 : That the equation of motion of a particle in a gravitational
field is independent of the mass of the particle, in agreement with Galileo
principle.

NOTE 3 : The equations of motion of a particle is no more than a state-


ment about the exchange of momentum between particle and field. Hence the
equations of motion cannot be independent of the conservation law for the
energy-momentum tensor of the fields

9.12 Local Cartesian coordinates(-)


• For a general n-dimensional Riemannian manifold, it is not possible to
perform a coordinate transformation xa → x0a that will take the line ele-
ment ds2 = gαβ (x)dxα dxβ into ds2 = δαβ (x)dx0α dx0β at every point in the
manifold.
166 9 Physics on Curved Spaces

• This can be understood by counting how many transformation equa-


tions are available, n and how many are the unknown independent components
of gµν (x) i.e. n(n + 1)/2.
• However, it is always possible to make a coordinate transformation
such that in the neighborhood of some specific point P the line element takes
the Euclidean form. In other words, we can always find coordinates x0a such
0
that at the point P the new metric functions gαβ (x0 ) satisfy:
0
∂gαβ
 
0
gαβ (P ) = δαβ , and = 0. (9.61)
∂x0µ P

Thus, in the neighborhood of P , we have


0
(x0 ) = δαβ + O (x0 − x0P )2
 
gαβ (9.62)

Such coordinates are called local Cartesian coordinates at P .


The Taylor expansion about a point P of the transformation is:
 α 
∂x 
xα (x0 ) = xα
P + 0β
x0β − x0β P
∂x
 2 α
P 
1 ∂ x 0β 0β

0γ 0γ 
+ x − x P x − x P (9.63)
2 ∂x0β ∂x0γ P
∂ 3 xα
 
1 



x0γ − x0γ
 0δ
x − x0δ

+ x − x P P P + ...
6 ∂x0β ∂x0γ ∂x0δ P

Fig. 9.2. Construction of local geodesics coordinates near a point P on the surface
of a sphere. A small flat patch is placed tangent to the sphere at P . The rectangular
coordinates give us the local coordinates.

∂xα

• N 2 independent values
∂x0β2 α
P 
∂ x
• 0β ∂x0γ N 2 (N + 1)/2 independent values
 ∂x P 
∂ 3 xα
• ∂x0β ∂x0γ ∂x0δ
N 2 (N + 1)(N + 2)/6 independent values
P

The quantities that we want to fix are:


9.12 Local Cartesian coordinates(-) 167

0
• gαβ (P ) N (N + 1)/2 independent values
 0 
∂gαβ
• ∂x0µ N 2 (N + 1)/2 independent values
P
0
∂ 2 gαβ

• ∂x0µ x0ν N 2 (N + 1)2 /4 independent values
P

9.12.1 Local geodesic coordinates and Cartesian coordinates(-)

We will show how one can get s system of local Cartesian coordinates starting
from a general system of coordinate xα via equations
0
gαβ (P ) = δαβ (9.64)
0 
∂gαβ

= 0. (9.65)
∂x0µ P
if the first of the two relations is valid (and not the 2nd) in the new coordinate
system then from the definition of the Christoffel symbols they have to vanish
i.e.
Γ 0α βγ (P ) = 0. (9.66)
If the 2nd of the above relations is valid (and not the 1st) then since gµα,β =
Γ ν γβ gνα + Γ ν αβ gµν (remember that gµα;β = 0) this leads again to equation
(9.66).
Coordinates for which equation (9.66) holds are referred to as geodesic
coordinates about P .
If we start with some arbitrary coordinates xα and lets a point P has

coordinates xα P and then we define a new coordinate system x by
1 α µ
x0α = xα − xα ν ν µ
P + Γ νµ (P ) (x − xP ) (x − xP ) (9.67)
2
by differentiation we get
∂x0α
= δνα + Γνµ
α
(P ) (xµ − xµP )
∂xν
thus at the point P we get: ∂x0α /∂xν = δνα (or ∂xα /∂x0ν = δνα ).
One more differentiation will give:
∂ 2 x0 α
= Γ α µν (P )δλµ = Γ α λν (P )
∂xλ ∂xν
and if we substitute these two relations in the formula for the transformation
of connections (Chapter 1):

Γ 0α βµ (P ) = δνα δµλ δβσ Γ ν λσ (P ) − δβν δµλ Γ α νλ (P ) = Γ 0α βµ (P ) − Γ 0α βµ (P ) = 0

Thus in the new primed coordinates the Christoffel symbols are zero at P
and thanks to condition (9.66) we constructed a system of geodesic coordinates
at P .
168 9 Physics on Curved Spaces

The metric function at this coordinate system will not satisfy necessarily
the first condition (9.64). Thus one more transformation of the form

x00α = Aα β x0β
00
which is linear and thus the coefficients of Aα β will bring the metric gαβ (P )
in these coordinates into the form (9.64) without affecting the derivatives and
the equation (9.65) will still be satisfied.

9.13 Geodesic deviation (-)


In a curved spacetime two geodesics that can be “parallel” initially will either
converge or diverge depending on the local curvature. Consider two neigh-
bouring geodesics G given by xα (σ) and G̃ given by x̃α (σ) where σ is an affine
parameter. If ξ α (σ) is a small vector connecting points of the two geodesics
for the same values of σ i.e.

x̃α (σ) = xα (σ) + ξ α (σ)

If we construct local geodesic coordinates about the point P , the Christof-


fel symbols will vanish but its derivatives will be non-zero there.

Fig. 9.3. (Left) Two neighbouring geodesics. (Right) Converging geodesics on the
surface of a sphere.

In this coordinate system we will get


 2 α  2 α
dx̃µ dx̃ν

d x d x̃ α
= 0 , + Γ µν =0 (9.68)
dσ 2 P dσ 2 dσ dσ Q

But since ξ α is small:

Γ α µν (Q) = Γ α µν (P ) + [Γ α µν,λ ]P ξ λ = [Γ α µν,λ ]P ξ λ

by subtracting the two equations in (??) we get (to 1st order, at P ):

d2 ξ α dxµ dxν λ
2
+ Γ α µν,λ ξ =0
dσ dσ dσ
9.14 Tidal forces in a curved spacetime (-) 169

However, in our geodesic coordinates the 2nd order absolute (intrinsic) deriva-
tive of ξ α at P is:

D2 ξ α d dξ α d2 ξ α dxµ dxλ ν
 
α µ ν α
= + Γ µν ξ x = + Γ µν,λ ξ
Dσ 2 dσ dσ dσ 2 dσ dσ
α
where we have used the fact that Γµν (P ) = 0.
By combining the last two equations we get:

D2 ξ α α α
µ
ν dx dx
λ
+ [Γ µλ,ν − Γ µν,λ ]ξ =0
Dσ 2 dσ dσ
which will give

D2 ξ α α
µ
ν dx dx
ν
+ R µνλ ξ =0 (9.69)
Dσ 2 dσ dσ
because the term in the square brackets is the Riemann tensor in local geodesic
coordinates.

9.14 Tidal forces in a curved spacetime (-)

Tidal forces deform the shape of bodies as they freely move in a gravitational
field.
Thus two nearby particles with trajectories xi (t) and x̃i (t) (in Cartesian
coordinates) will be separated by a vector ζ i = xi − x̃i
d2 ζ
 2 
∂ Φ
= − ζj (9.70)
dt2 ∂xi ∂xj
(why?) where Φ is the Newtonian gravitational potential.

Fig. 9.4. Tidal effects on a cloud of particles

Tidal effects can be also estimated in GR for two particles moving along
timelike geodesics xµ (τ ) and x̃µ (τ ) (τ is the proper time of the 1st particle).
170 9 Physics on Curved Spaces

The separation vector between the worldlines of the 2 particles is ξ µ (τ ) =


x̃ − xµ :
µ
D2 ξ µ
= Rµ σρν uσ uρ ξ ν ≡ S µ ν ξ ν (9.71)
Dτ 2
where S µ ν is the so called tidal stress tensor and uσ = duσ /dτ . This is a
fully covariant tensor equation and holds in any coordinate system.

Fig. 9.5. The basis vectors of the instantaneous rest frame (IRF) at P .
• êα is a set of orthonormal basis vectors at P that define the IRF of the first
particle (observer) with êα · êβ = ηαβ .
• ξ is a general connecting vector with ξ α̂ ≡ êα · ξ = (êα )µ ξ µ
• ζ is the orthogonal connecting vector.

For an observer sitting on the one of the particles it can be shown that in
any orthonormal freely falling frame becomes:

d2 ξ µ̂
= c2 Rµ̂ 0̂0̂ν̂ ξ ν̂ . (9.72)
dτ 2
Newtonian limit (we will discuss the details later)

∂2Φ
Rµ̂ 0̂0̂ν̂ → (9.73)
∂xi ∂xj
10
Einstein’s Theory of Gravity

10.1 Newtonian Gravity (***)


Poisson equation
Z
ρ(x)
2
∇ U (x) = 4πGρ(x) → U (x) = −G d3 x0
|x − x0 |
For a spherically symmetric mass distribution of radius R

1 R 02 0 0
Z
U (r) = − r ρ(r )dr for r > R
r 0
1 r 02 0 0
Z Z R
U (r) = − r ρ(r )dr − r0 ρ(r0 )dr0 for r < R
r 0 r

Fig. 10.1.

For a non-spherical distribution the term 1/|x − x0 | can be expanded as


172 10 Einstein’s Theory of Gravity

1 1 X xk x0k 1 XX 0k 0l
 k l
02 l x x
= + + 3x x − r δk + ...
|x − x0 | r r3 2 r5
k k l

GM G X k k G X kl xk xl
U (x) = − − 3 x D − Q + ...
r r 2 r5
k kl

Gravitational Multipoles
Z
M = ρ(x0 )d3 x0 Mass
Z
k
D = x0k ρ(x0 )d3 x0 Mass Dipole moment1
Z
Qkl = 3x0k x0l − r02 δkl ρ(x0 )d3 x0

Mass Quadrupole tensor2

The Earth’s polar and equatorial diameters differ by 3/1000. This devia-
tion produces a quadrupole term in the gravitational potential, which causes
perturbations in the elliptical Kepler orbits of satellites. Usually, we define
the dimensionless parameter

Q33
J2 = − (10.1)
2M R2
as a convenient measure of the oblateness of a nearly spherical body. For the
Sun the oblateness due to rotation gives J2 ≈ 10−7 .
The main perturbation is the precession of Kepler’s ellipse and this can
be used for precise determinations of the multipole moments and the mass
distribution in the Earth.

Fig. 10.2.

What about the Sun?


10.2 Equivalence Principle (****) 173

10.2 Equivalence Principle (****)


In GR gravitational phenomena arise not from forces and fields, but from the
curvature of the 4-dim spacetime. The starting point for this consideration is
the Equivalence Principle which states that the Gravitational and the Inertial
masses are equal. The equality mG = mI is one of the most accurately tested
principles in physics!
Now it is known experimentally that: γ = mGm−m G
I
< 10−13 Experimen-
tal verification
• Galileo (1610) , Newton (1680) γ < 10−3
• Bessel (19th century) γ < 2 × 10−5
• Eötvös (1890) & (1908) γ < 3 × 10−9
• Dicke et al. (1964) γ < 3 × 10−11
• Braginsky et al. (1971) γ < 9 × 10−13 ,
• Kuroda and Mio (1989) γ < 8 × 10−10 ,
• Adelberger et al. (1990) γ < 1 × 10−11
• Su et al. (1994) γ < 1 × 10−12 (torsional)
• Williams et al. (‘96), Anderson & Williams (‘01) γ < 1 × 10−13

• Weak Equivalence Principle : The motion of a neutral test body released


at a given point in space-time is independent of its composition
• Strong Equivalence Principle :
– The results of all local experiments in a frame in free fall are indepen-
dent of the motion
– The results are the same for all such frames at all places and all times
– The results of local experiments in free fall are consistent with STR

10.2.1 Equivalence Principle : Dicke’s Experiment

The experiment is based on measuring the effect of the gravitational field on


two masses of different material in a torsional pendulum.

Fig. 10.3.

(E) (E) (E)


m v2

GM mG GM
2 mG
= I ⇒ v =
R2 R R mI
174 10 Einstein’s Theory of Gravity

The forces acting on both masses are:


" (j) #
(j) (j) (j) (E)
m v2

(j) GM mG (j) GM mI mG mG
F + 2
= I ⇒ F = −
R R R2 mI mI
h i
 GM (1) (2)
and the total torque applied is: L = F (1) − F (2) ` = R2 mG − mG `
(1) (2)
here we assumed that : mI = mI = m.

10.3 Einstein’s Equations (****)

Einstein’s equivalence principle: Gravitational and Inertial forces/accelerations


are equivalent and they cannot be distinguished by any physical experiment.
This statement has the 3 implications:
1. Gravitational accelerations are described in the same way as
the inertial ones.
• This means that the motion of a freely moving particle, observed from
2 µ
an inertial frame, will be described by ddtx2 = 0 while from a non-inertial
frame its movement will be described by the geodesic equation

d2 xµ ρ
µ dx dx
σ
+ Γ ρσ = 0.
ds2 ds ds
• The 2nd term appeared due to the use of a non-inertial frame, i.e. the
inertial accelerations will be described by the Christoffel symbols.
• But according to Einstein the gravitational accelerations as well will
be described by the Christoffel symbols.
• This leads to the following conclusion: The metric tensor should play
µ
the role of the gravitational potential since Γρσ is a function of the metric
tensor and its derivatives.
2. When gravitational accelerations are present the space can-
not be flat since the Christoffel symbols are non-zero and the Riemann tensor
is not zero as well.
In other words, the presence of the gravitational field forces the space to
be curved, i.e. there is a direct link between the presence of a gravitational
field and the geometry of the space.
3. Consequence: if gravity is present there cannot exist inertial
frames.
If it was possible then one would have been able to discriminate among
inertial and gravitational accelerations which against the “generalized equiv-
alence principle” of Einstein.
The absence of “special” coordinate frames (like the inertial) and their
substitution from “general” (non-inertial) coordinate systems lead in naming
Einstein’s theory for gravity “General Theory of Relativity”.
10.4 Solutions of Einstein’s Equations (*) 175

Since the source of the gravitational field is a tensor (Tµν ) the field should
be also described by a 2nd order tensor e.g. Fµν . Since the role of the gravita-
tional potential is played by the metric tensor then Fµν should be a function of
the metric tensor gµν and its 1st and 2nd order derivatives. Moreover, the law
of energy-momenum conservation implies that T µν ;µ = 0 which suggests that
F µν ;µ = 0 Then since F µν should be a linear function of the 2nd derivative
of gµν we come to the following form of the field equations (how & why?):

F µν = Rµν + ag µν R + bg µν = κT µν (10.2)
8πG
where κ = c4 . Then since F µν ;µ = 0 there should be

(Rµν + ag µν R + bg µν );µ = 0 (10.3)

which is possible only for a = −1/2. Thus the final form of Einstein’s equations
is:
1
Rµν − g µν R + Λg µν = κT µν . (10.4)
2
8πG
where Λ = c2 ρv is the so called cosmological constant.

10.3.1 Newtonian Limit (**)

In the absence of strong gravitational fields and for small velocities both Ein-
stein & geodesic equations reduce to the Newtonian ones.
• Geodesic equations:

d2 xµ α
µ dx dx
β
d2 t/ds2 dxµ d2 xj
+ Γαβ = − ⇒ ≈ g00,j (10.5)
dt2 dt dt (dt/ds)2 dt dt2
2 k
If g00 ≈ η00 + h00 = 1 + h00 = 1 + 2 cU2 then ddtx2 ≈ ∂x
∂U
k

• Einstein equations
 
1 1
Rµν = −κ Tµν − gµν T ⇒ ∇2 U = κρ (10.6)
2 2

where κ = 8πG.
Here we have used the following approximations:

j 1 j
Γ00 ≈ g00,j and R00 ≈ Γ00,j ≈ ∇2 U
2

10.4 Solutions of Einstein’s Equations (*)


• A static spacetime is one for which a timelike coordinate x0 has the
following properties:
176 10 Einstein’s Theory of Gravity

1. all metric components gµν are independent of x0


2. the line element ds2 is invariant under the transformation x0 → −x0 .
Note that the 1st property does not imply the 2nd (e.g. the time reversal
on a rotating star changes the sense of rotation, but the metric components
are constant in time).
• A spacetime that satisfies (I) but not (II) is called stationary.
• The line element ds2 of a static metric depends only on rotational in-
variants of the spacelike coordinates xi and their differentials, i.e. the metric
is isotropic.
The only rotational invariants of the spacelike coordinates xi and their
differentials are

x · x ≡ r2 , x · dx ≡ rdr, dx · dx ≡ dr2 + r2 dθ2 + r2 sin2 θdφ2

Thus the more general form of a spatially isotropic metric is:


2
ds2 = A(t, r)dt2 − B(t, r)dt (x · dx) − C(t, r) (x · dx) − D(t, r)dx2
= A(t, r)dt2 − B(t, r)r dt dr − C(t, r)r2 dr2
−D(t, r) dr2 + r2 dθ2 + r2 sin2 θdφ2

(10.7)
2 2 2 2 2

= A(t, r)dt − B̃(t, r)dt dr − C̃(t, r)dr − D̃(t, r) dθ + sin θdφ
= A0 (t, r̃)dt2 − B 0 (t, r̃)dt dr̃ − C 0 (t, r̃)dr̃2 − r̃2 dθ2 + sin2 θdφ2 (10.8)


where we have set r̃2 = D(t, r) and we redefined the A, B̃ and C̃. The next
step will be to introduce a new timelike coordinate t̃ as
 
0 1 0
dt̃ = Φ(t, r̃) A (t, r̃)dt − B (t, r̃)dr̃
2

where Φ(t, r̃) is an integrating factor that makes the right-hand side an exact
differential.
By squaring we obtain
 
1
dt̃2 = Φ2 A02 dt2 − A0 B 0 dtdr̃ + B 02 dr̃2
4

from which we find


1 B0 2
A0 dt2 − B 0 dtdr̃ = dt̃2 − dr̃
A0 Φ2 4A0

Thus by defining the new functions  = 1/(A0 Φ)2 and B̂ = C + B 0 /(4A0 ) the
metric 10.8 becomes diagonal

ds2 = Â(t̃, r̃)dt̃2 − B̂(t̃, r̃)dr̃2 − r̃2 dθ2 + sin2 θdφ2



(10.9)

or by dropping the ‘hats’ and ‘tildes’


10.5 Schwarzschild Solution (****) 177

ds2 = A(t, r)dt2 − B(t, r)dr2 − r2 dθ2 + sin2 θdφ2



(10.10)

• Thus the general isotropic metric is specified by two functions of t and r,


namely A(t, r) and B(t, r).
• Also, surfaces for t and r constant are 2-spheres (isotropy of the metric).
• Since B(t, r) is not unity we cannot assume that r is the radial distance.

10.5 Schwarzschild Solution (****)


A typical solution of Einstein’s equations describing spherically symmetric
spacetimes has the form:

ds2 = eν(t,r) dt2 − eλ(t,r) dr2 − r2 dθ2 + sin2 θdφ2



(10.11)

Then the components of the Ricci tensor will be:


" #
ν 00 ν 02 ν 0 λ0 λ0 λ−ν λ̈ λ̇2 λ̇ν̇
R11 = − − + + +e + − (10.12)
2 4 4 r 2 4 4
 00
ν 02 ν 0 λ0 ν0 λ̈ λ̇2

ν λ̇ν̇
R00 = eν−λ + − + − − + (10.13)
2 4 4 r 2 4 4
R10 = λ̇/r (10.14)
h r i
R22 = = −e−λ 1 + (ν 0 − λ0 ) + 1 (10.15)
2
2
R33 = sin θR22 (10.16)

in the absence of matter (outside the source) Rµν = 0 we can prove that
λ(r) = −ν(r), i.e. the solution is independent of time (how and why?).

   −1
2GM 2 2 2GM
2
dr2 − r2 dθ2 + sin2 θdφ2

ds = 1 − c dt − 1 −
rc2 rc2

• Sun: M ≈ 2 × 1033 gr and R = 696.000km


2GM
≈ 4 × 10−6
rc2
• Neutron star : M ≈ 1.4M and R ≈ 10 − 15km
2GM
≈ 0.3 − 0.5
rc2
• Neutonian limit:
2U GM
g00 ≈ η00 + h00 = 1 + ⇒ U=
c2 r
178 10 Einstein’s Theory of Gravity

10.5.1 Schwarzschild Solution: Geodesics


1 1
r̈ − ν 0 ṙ2 − reν θ̇2 − reν sin2 θφ̇2 + e2ν ν 0 ṫ2 =0 (10.17)
2 2
2
θ̈ + ṙθ̇ − sin θ cos θφ̇2 =0 (10.18)
r
2 cos θ
φ̈ + ṙφ̇ + 2 θ̇φ̇ =0 (10.19)
r sin θ
ẗ + ν 0 ṙṫ =0 (10.20)

and from gλµ ẋλ ẋµ = 1 (here we assume a massive particle) we get :

−eν ṫ2 + eλ ṙ2 + r2 θ̇ + r2 sin2 θφ̇2 = 1. (10.21)

If a geodesic is passing through a point P in the equatorial plane (θ = π/2)


and has a tangent at P situated also in this plane (θ̇ = 0 at P ) then from
(10.18) we get θ̈ = 0 at P and all higher derivatives are also vanishing at P .
That is, the geodesic lies entirely in the plane defined by P , the tangent at
P and the center of symmetry of the space. Since the symmetry planes are
equivalent to each other, it will be sufficient to discuss the geodesics lying on
one of these planes e.g. the equatorial plane θ = π/2.
The geodesics on the equatorial plane are:
1 1
r̈ − ν 0 ṙ2 − reν φ̇2 + e2ν ν 0 ṫ2 = 0 (10.22)
2 2
2
φ̈ + ṙφ̇ = 0 (10.23)
r
ẗ + ν 0 ṙṫ = 0 (10.24)
−eν ṫ2 + eλ ṙ2 + r2 φ̇2 = 1 (10.25)

Then from equations (10.23) and (10.24) we can easily prove (how?) that:
d  2 
r φ̇ = 0 ⇒ r2 φ̇ = L = const : (Angular Momentum) (10.26)

d ν 
e ṫ = 0 ⇒ eν ṫ = E = const : (Energy) (10.27)

• For a massive particle with unit rest mass, by assuming that τ is the
affine parameter for its motion we get pµ = ẋµ . Thus

p0 = g00 ṫ = eν ṫ = E and pφ = gφφ φ̇ = −r2 φ̇ = −L (10.28)

• An observer with 4-velocity U µ will find that the energy of a particle


with 4-momentum pµ is:
E = pµ U µ
An observer at infinity, U µ = (1, 0, 0, 0) will find E = p0 = E.
Actually, for a particle with rest mass m0 we get E = E/(m0 c2 ) i.e. it is
the energy per unit rest-mass.
10.5 Schwarzschild Solution (****) 179

————
Finally, by substituting eqns (10.27) and (10.26) into (10.25) we get the
“energy equation” for the r-coordinate
eν 2
ṙ2 +
L + eν = E 2 (10.29)
r2
which suggests that at r → ∞ we get that E = 1.
By combining the 2 integrals of motion and eqn (10.25) we can eliminate
the proper time τ to derive an equation for a 3-d path of the particle (how?)
2
E 2 − 1 2M u

d
u + u2 = + + 2M u3 (10.30)
dφ L2 L2

where we have used u = 1/r and

dr dr dφ L dr
ṙ = = = 2
dτ dφ dτ r dφ
The previous equation can be written in a form similar to Kepler’s equation
of Newtonian mechanics (how?) i.e.

d2 M
u + u = 2 + 3M u2 (10.31)
dφ2 L

The term 3M u2 is the relativistic correction to the Newtonian equation.


This term for the trajectories of planets in the solar system, where M =
M = 1.47664 km, and for orbital radius r ≈ 4.6 × 107 km (Mercury) and
r ≈ 1.5 × 108 km (Earth) gets verys small values ≈ 3 × 10−8 and ≈ 10−8
correspondingly.

10.5.2 Radial motion of massive particles (**)

For the radial motion φ is constant, which implies that L = 0 and eqn (10.29)
reduces to
ṙ2 = E 2 − eν (10.32)
and by differentiation we get an equation which reminds the equivalent one
of Newtonian gravity i.e.
M
r̈ = 2 . (10.33)
r
• If a particle is dropped from the rest at r = R we get that E 2 =
ν(R)
e = 1 − 2M/R and (10.32) will be written
 
2 1 1
ṙ = 2M − (10.34)
r R
180 10 Einstein’s Theory of Gravity

which is again similar to the Newtonian formula for the gain of kinetic energy
due to the loss in gravitational potential energy for a particle (of unit mass)
falling from rest at r = R.
• For a particle dropped from the rest at infinity E = 1 and the
geodesic equations are simplified
r
dt −ν dr 2M
=e and =− (10.35)
dτ dτ r
The component of the 4-velocity will be
r !
µ dxµ −ν 2M
u = = e ,− ,0,0 (10.36)
dτ r

Then by integrating the second of (10.35) and by assuming that at τ = τ0


that r = r0 we get r r
2 r03 2 r3
τ= − (10.37)
3 2M 3 2M
q
r03
which suggests that for r = 0 we get τ → 32 2M ie the particle takes finite
proper time to reach r = 0.
If we want to map the trajectory of the particle in the (r, t) coordinates
we need to solve the equation
r
dr dr dτ −ν 2M
= = −e (10.38)
dt dτ dt r
The integration leads to the relation
r r !
r03
r r
r3

2 r0 r
t= − + 4M −
3 2M 2M 2M 2M
p ! p !
r/2M + 1 r0 /2M − 1
+ 2M ln p (10.39)

p
r/2M − 1 r0 /2M + 1

where we have chosen that for t = 0 to set r = r0 .


Notice that when r → 2M then t → ∞. In other words, for an observer at
infinity, it takes infinite time for a particle to reach r = 2M .
———-
QUESTION : Can you find what will be the velocity of the radially
falling particle for a stationary observer at coordinate radius r̃?
r r
2 r03 2 r3
τ= −
3 2M 3 2M
0v v 1
u 3
u r3
u s s ! ˛ p ! p !˛
2 But r0 r0 r ˛ r/2M + 1 r0 /2M − 1 ˛
t = − A + 4M − + 2M ln ˛
t C ˛ ˛
@ p p ˛
3 2M 2M 2M 2M ˛ r/2M − 1 r0 /2M + 1 ˛
10.5 Schwarzschild Solution (****) 181

Fig. 10.4. Radial fall from rest towards a Schwarschild BH as described by a co-
moving observer (proper time τ ) and by a distant observer (Schwarschild co-
ordinate time t)

10.5.3 Circular motion of massive particles (**)

The motion of massive particles in the equatorial plane is described by eqn


(10.31)
d2 M
u + u = 2 + 3M u2 (10.40)
dφ2 L
For circular motions r = 1/u=const and ṙ = r̈ = 0. Thus we get

r2 M
L2 = (10.41)
r − 3M
if we also put ṙ = 0 in eqn (10.29) we get :
1 − 2M/r
E=p (10.42)
1 − 3M/r

Circular orbits will be bound for E/(m0 c2 ) = E < 1, so the limit on r for an
orbit to be bound is given by E = 1 which leads to
2
(1 − 2M/r) = 1 − 3M/r true when r = 4M or r = ∞ (10.43)

Thus over the range 4M < r < ∞ circular orbits are bound.
From the integral of motion r2 φ̇ = L and eqn (10.41) we get
 2
dφ M
= 2 (10.44)
dτ r (r − 3M )
NOTICE : This equation cannot be satisfied for circular orbits with r <
3M . Such orbits cannot be geodesics and cannot be followed by freely falling
particles.
182 10 Einstein’s Theory of Gravity

Fig. 10.5. The variation of E = E/(m0 c2 ) as a function of r/M for a circular orbit
of a massive particle in the Schwarzchild geometry

Fig. 10.6. The shape of a bound orbit outside a spherical star or a black-hole

We can also calculate an expression for Ω = dφ/dt


2 2 2
e2ν
  
dφ dφ dτ dφ M
Ω2 = = = = (10.45)
dt dτ dt E2 dτ r3

which is equivalent to Kepler’s law in Newtonian gravity.

10.5.4 Stability of massive particle orbits (-)

According to the previous discussion the closest bound orbit around a massive
body is at r = 4M , however we cannot yet determine whether this orbit is
stable.
In Newtonian theory the particle motion in a central potential is described
by:
10.5 Schwarzschild Solution (****) 183
 2
1 dr
+ Veff (r) = E 2 (10.46)
2 dt
M L2
Veff (r) = − + 2 (10.47)
r 2r

Fig. 10.7.

The bound orbits have two turning points while the circular orbit corre-
sponds to the special case where the particle sits in the minimum of the of
the effective potential.
In GR the ‘energy’ equation (10.29) is
eν 2
ṙ2 + L + eν = E 2 (10.48)
r2
which leads to an effective potential of the form
eν 2 M L2 M L2
Veff (r) = 2
L + eν = − + 2− 3 (10.49)
r r 2r r
Circular orbits occur where dVeff /dr = 0 that is:

dVeff M L2 3M L2
= 2 − 3 + (10.50)
dr r r r4
so the extrema are located at the solutions of the eqn M r2 − Lr + 3M L2 = 0
which occur at
L  p 
r= L ± L2 − 12M 2
2M
√ √
Note that if L = 12M = 2 3M then there is only one extremum and no
turning points in the orbit for lower values of L. Thus the innermost stable
circular orbit (ISCO) has

rISCO = 6M and L = 2 3M
184 10 Einstein’s Theory of Gravity

Fig. 10.8. The dots indicate the locations of stable circular orbits which occur
at the local minimum of the potential. The local maxima in the potential are the
locations of the unstable circular orbits

Fig. 10.9. Orbits for L = 4.3 and different values of E.


10.5 Schwarzschild Solution (****) 185

and it is unique in satisfying both dVeff /dr = 0 and d2 Veff /dr2 = 0, the latter
is the condition for marginal stability of the orbit.
• The upper shows circular orbits. A stable (outer) and an unstable (in-
ner) one.
• The lower shows bound orbits , the particle moves between two turning
points marked by dotted circles.

Fig. 10.10. Orbits for L = 4.3 and different values of E.

• The upper shows a scattering orbit the particle comes in from infinity
passes around the center of attraction and moves out to infinity again.
• The lower shows plunge orbit , in which the particle comes in from
infinity.

10.5.5 Trajectories of photons (-)

Photons as any zero rest mass particle move on null geodesics. In this case
we cannot use the proper time τ as the parameter to characterize the motion
and thus we will use some affine parameter σ.
We will study photon orbits on the equatorial plane and the equation of
motion will be in this case

eν ṫ = E (10.51)
ν 2 −ν 2 2 2
e ṫ − e ṙ − r φ̇ = 0 (10.52)
2
r φ̇ = L (10.53)

The equivalent to eqn (10.29) for photons is


186 10 Einstein’s Theory of Gravity

eν 2
ṙ2 + L = E2 (10.54)
r2
while the equivalent of eqn (10.31) is
d2
u + u = 3M u2 (10.55)
dφ2

10.5.6 Radial motion of photons(-)


For radial motion φ̇ = 0 and we get
eν ṫ2 − e−ν ṙ2 = 0
from which we obtain  
dr 2M
=± 1− (10.56)
dt r
integration leads to:
Outgoing Photon
r
t = r + 2M ln − 1 + const

2M
Incoming Photon
r
t = −r − 2M ln − 1 + const

2M

10.5.7 Circular motion of photons (-)


For circular orbits we have r=constant and thus from eqn (10.55) we see that
the only possible radius for a circular photon orbit is:
 
3GM
r = 3M or r = (10.57)
c2
There are no such orbits around typical stars because their radius is much
larger than 3M (in geometrical units). But outside the black hole there can
be such an orbit.

10.5.8 Stability of photon orbits (-)


We can rewrite the “energy” equation (10.54) as
ṙ2 eν 1
2
+ 2
= 2 (10.58)
L r D
where
WeDcan
= L/E = eνa/r
andVVefefff has
see that single maximum at r = 3M where the value
2
of the potential is 1/(27M ). Thus the circular orbit r = 3M is unstable.
There are no stable circular photon orbits in the Schwarzschild
geometry.
10.6 The slow-rotation limit : Dragging of inertial frames (-) 187

Fig. 10.11. Radially infalling particle emitting a radially outgoing photon.

Fig. 10.12. The effective potential for photon orbits.

10.6 The slow-rotation limit : Dragging of inertial


frames (-)
4M a
ds2 = ds2Schw + sin2 θdtdφ (10.59)
r
where J = M a is the angular momentum.
The contravariant components of the particles 4-momentum will be

pφ = g φµ pµ = g φt pt + g φφ pφ and pt = g tµ pµ = g tt pt + g tφ pφ

If we assume a particle with zero angular momentum, i.e. pφ = 0 along the


geodesic then the particle’s trajectory is such that

dφ pφ g tφ 2M a
= t = tt = 3 = ω(r)
dt p g r
188 10 Einstein’s Theory of Gravity

which is the coordinate angular velocity of a zero-angular-momentum


particle.
A particle dropped “straight in” from infinity ( pφ = 0) is dragged just by
the influence of gravity so that acquires an angular velocity in the same
sense as that of the source of the metric.
The effect weakens with the distance and makes the angular momentum
of the source measurable in practice.

Useful constants
Useful Constants in geometrical units

Speed of light c =299,792.458 km/s = 1


Planck’s constant ~ = 1.05 × 10−27 erg ·s = 2.612 × 10−66 cm2
Gravitation constant G = 6.67 × 10−8 cm3 /g ·s2 =1
Energy eV=1.602×10−12 erg = 1.16 × 104 K
=1.7823×10−33 g =1.324×10−56 km
Distance 1 pc=3.09×1013 km=3.26 ly
Time 1 yr = 3.156×107 sec
Light year 1 ly = 9.46×1012 km
Astronomical unit (AU) 1AU = 1.5 × 108 km
Earth’s mass M⊕ = 5.97 × 1027 g
Earth’s radius (equator) R⊕ = 6378 km
Solar Mass M = 1.99 × 1033 g =1.47664 km
Solar Radius R = 6.96 × 105 km

10.7 The Classical Tests


10.7.1 The Classical Tests: Perihelion Advance (***)

For a given value of angular momentum Kepler’s equation (10.31)

d2 M
u + u = 2 + 3M u2 (10.31)
dφ2 L

(without the GR term 3M u2 ) admits a solution given by

M
u= [1 + e cos(φ + φ0 )] (10.60)
L2
where e and φ0 are integration constants.
• We can set φ0 = 0 by rotating the coordinate system by φ0 .
• The other constant e is the eccentricity of the orbit which is an
ellipse if e < 1.
10.7 The Classical Tests 189

Now since the term 3M u2 is small we can use perturbation theory to get
a solution of equation (10.31).

d2 M 3M 3 2 M 3M 3 6eM 3
2
u + u ≈ 2 + 4 [1 + e cos(φ)] ≈ 2 + 4 + cos(φ)
dφ L L L L L4

Because, 3M 3 /L4  M/L2 and can be omitted and its corrections will be
small periodic elongations of the semiaxis of the ellipse.
The term 6eM 3 /L4 cos(φ) is also small but has an accumulative effect
which can be measured. Thus the solution of the relativistic form of Kepler’s
equation (10.61) becomes (k = 3M 2 /L2 )

3eM 2
 
M M
u = 2 1 + e cos φ + 2
φ sin φ ≈ 2 {1 + e cos[φ(1 − k)]} (10.61)
L L L

Fig. 10.13.

The perihelion of the orbit can be found be maximizing u that is when


cos[φ(1 − k)] = 1 or better if φ(1 − k) = 2nπ. This mean that after each
rotation around the Sun the angle of the perihelion will increase by

2nπ 2π 6πM 2
φn = ⇒ φn+1 − φn = ≈ 2π(1 + k) = 2π + (10.62)
1−k 1−k L2

i.e. in the relativistic orbit the perihelion is no longer a fixed point, as it


was in the Newtonian elliptic orbit but it moves in the direction of the motion
of the planet, advancing by the angle

6πM 2
δφ ≈ (10.63)
L2
190 10 Einstein’s Theory of Gravity

By using the well know relation from Newtonian Celestial Mechanics L2 =


M r0 (1 − e2 ) connecting the perihelion distance r0 with L we derive a more
convenient form for the perihelion advance
6πM
δφ ≈ (10.64)
r0 (1 − e2 )
Solar system measurements In binary pulsars separated by 106 km the

Mercury (43.11 ± 0.45)00 43.0300


Venus (8.4 ± 4.8)00 8.600
00
Earth (5.0 ± 1.2) 3.800

perihelion advance is extremely important and it can be up to ∼ 2o /year


(about 1000 orbital rotations)
Sources of the precession of perihelion for Mercury

Amount (arcsec/century) Gause


5025.6 Coordinate ( precession of the equinoxes)3
531.4 Gravitational tugs of the other planets
0.0254 Oblateness of the Sun
42 ± 0.04 General relativity
5600.0 Total
5599.7 Observed

10.7.2 The Classical Tests : Deflection of Light Rays (***)

Deflection of light rays due to presence of a gravitational field is a prediction


of Einstein dated even before the GR and has been verified in 1919.
Photons follow null geodesics, this means that ds = 0 and the integrals of
motion (E and L) are divergent but not their ratio L/E. Thus the photon’s
equation of motion on the equatorial plane of Schwarszchild spacetime will
be:
d2 u
+ u = 3M u2 . (10.65)
dφ2
with an approximate solution (we omit at the moment the term 3M u2 )
1
u= cos(φ + φ0 ) (10.66)
b
which describes a straight line, where b and φ0 are integration constants.
Actually, with an appropriate rotation φ0 = 0 and the length b is the distance
of the line from the origin.
10.7 The Classical Tests 191

Fig. 10.14.

Fig. 10.15.

Since the term 3M u2 is very small we can substitute u with the Newtonian
solution (10.66) and we need to solve the non-homogeneous ODE
d2 u 3M
+ u = 2 cos2 φ . (10.67)
dφ2 b
admitting a solution of the form
cos φ M
+ 2 1 + sin2 φ

u= (10.68)
b b
which for distant observers (r → ∞) i.e. u → 0, and we get a relation between
φ, M and the parameter b
cos φ M
+ 2 1 + sin2 φ = 0

(10.69)
b b
and since r → ∞, this means that φ → π/2 +  and cos φ → 0 +  and
sin φ → 1 −  we get:
192 10 Einstein’s Theory of Gravity

2M
≈− (10.70)
b
Since, φ → π/2 + 2M/b for r → ∞ on the one side & φ → 3π/2 − 2M/b on
the other side the total deviation will be the sum of the two i.e.
4M
δφ = . (10.71)
b
For a light ray tracing the surface of the Sun gives a deflection of ∼ 1.7500 .
The deflection of light rays is a quite common phenomenon in Astronomy
and has many applications. We typically observe “crosses” or “rings”

Fig. 10.16. Einstein Cross (G2237+030) is the most characteristic case of gravi-
tational lens where a galaxy at a distance 5 × 108 lys focuses the light from a quasar
who is behind it in a distance of 8×109 lys. The focusing creates 4 symmetric images
of the same quasar. The system has been discovered by John Huchra.

Fig. 10.17. “Einstein rings” are observed when the source, the focusing body and
Earth are on the same line of sight. This ring has been discovered by Hubble space
telescope.

10.7.3 The Classical Tests : Gravitational Redshift (***)


The clocks of the 3 observers ticking with different rates. The clocks of the two
closer to the source are ticking slower than the clock of the observer at infinity
10.7 The Classical Tests 193

Fig. 10.18. Let’s assume 3 static observers on a Schwarszchild spacetime, one very close to the source of
the field the other in a medium distance from the source and the third at infinity.

 1/2
2M
who measures the so called ‘ coordinate time” i.e. dτ1 = 1 − r1 dt and
 1/2
dτ2 = 1 − 2M
r2 dt this means that
 1/2
2M
dτ2 1− r2 M M
dτ1
= 1/2 ≈ 1 + r − r . (10.72)
2M 1 2
1− r1

If the 1st observer sends light signals on a specific wavelegth λ1 from


c = λ/τ we get a relation between the wavelength of the emitted and received
signals
λ2 M M λ2 − λ1 ∆λ M M
≈1+ − ⇒ = = − . (10.73)
λ1 r1 r2 λ1 λ1 r1 r2
A similar relation can be found for the frequency of the emitted signal:
 1/2
2M
ν1 1 − r2
= 1/2 . (10.74)
ν2
1 − 2Mr1

While the photon redshift z is defined by


ν1
1+z = (10.75)
ν2
QUESTION : What will be the redshift for signals emitted from the
surface of the Sun, a neutron star and a black hole?

10.7.4 The Classical Tests : Radar Delay (***)

A more recent test (late 60s) where the delay of the radar signals caused
by the gravitational field of Sun was measured. This experiment suggested
and performed by I.I. Shapiro and his collaborators. The line element ds2 =
gµν dxµ dxν for the light rays i.e. for ds = 0 on the equatorial plane has the
form
   −1  2  2
2M 2M dr dφ
0= 1− − 1− − r2 (10.76)
r r dt dt
194 10 Einstein’s Theory of Gravity

Fig. 10.19.

For the study of the radial motion we should substitute the term dφ/dt from
the integrals of motion (10.26) and (10.27) i.e. we can create the quantity:
 −1
L 2M dφ
D= = r2 1 − (10.77)
E r dt

Then eqn (10.76) becomes


−1  2 2
D2
   
2M 2M dr 2M
1− − 1− − 2 1− = 0. (10.78)
r r dt r r

At the point of the closest approach to the Sun, r0 , there should be dr/dt = 0
r02
and thus we get the value of D2 = 1−2M/r 0
. Leading to an equation for the
radial motion:

dr

2M
  r 2 1 − 2M/r 1/2
0
= 1− 1− (10.79)
dt r r 1 − 2M/r0

leading to
Z r1
dr
t1 = q 2 1−2M/r
1 − 2M 1 − rr0 1−2M/r

r0
r 0
" p #
2 2
r
r1 r1 − r0 r1 − r0
q
2 2
= r1 − r0 + 2M ln +M
r0 r1 + r0
 
2r1
≈ r1 + 2M ln +M
r0

For flat space we have t1 = r1 i.e. the term t̃1 = 2M ln(2r1 /r0 ) + M is
the relativistic correction for the first part of the orbit in same way we get
a similar contribution as the signal returns to Earth. Thus the total “extra
time” is:   
 4r1 r2
∆T = 2 t̃1 + t̃2 = 4M 1 + ln . (10.80)
r02
10.7 The Classical Tests 195

Fig. 10.20. Comparison of the experimental results with the prediction of the
theory. The results are from I.I. Shapiro’s experiment (1970) using Venus as reflector.

Fig. 10.21.
11
Solutions of Einstein’s Equations & Black
Holes

11.1 Schwarzschild Solution: Black Holes (***)


The light cone in a Schwarzschild spacetime
   −1  
2 2M 2 2M dr 2M
ds ≡ 0 = 1 − dt − 1 − dr2 ⇒ =± 1− .
r r dt r
r = 2M is a null surface and is called horizon or infinite redshift surface.

Fig. 11.1. The light cone for Schwarzschild spacetime.

11.1.1 Eddington-Filkenstein coordinates

The Schwarzschild solution


   
2M 2M 4M
ds2 = 1 − dt02 − 1 + dr2 ± drdt0 − r2 dθ2 + sin2 θdφ2

r r r
198 11 Solutions of Einstein’s Equations & Black Holes

Fig. 11.2. Collapse Diagram

Fig. 11.3. Embedding diagrams for Schwarzschild spacetime.

in a new set of coordinates (t0 , r0 , θ0 , φ0 )


 0 
r
t = t0 ± ln − 1 , r > 2M
2M
r0
 
t = t0 ± ln 1 − , r < 2M
2M
r = r 0 , θ = θ 0 , φ = φ0

while the
   
2M 02 2M 4M
2
ds ≡ 0 = 1 − dt − 1 + dr2 ± drdt0
r r r
    
0 2M 0 2M
(dt + dr) 1− dt − 1 + dr = 0
r r
11.1 Schwarzschild Solution: Black Holes (***) 199

dr dr r − 2M
= −1 and =
dt0 dt0 r + 2M

Fig. 11.4. Light-cones in Eddington-Filkenstein coordinates

A manifold is said to be geodesically complete if all geodesics are of infinite


length. This means that in a complete specetime manifold, every particle has
been in the spacetime forever and remains in it forever.

11.1.2 Kruskal - Szekeres Coordinates : Maximal Extension (*)


 r 1/2 r  
t
v= −1 e sinh
4M
2M 4M
 r 1/2 r  
t
u= −1 e cosh
4M
2M 4M
 r 
− 1 er/2M = u2 − v 2
2M
t v
= tanh−1
4M u

32M 3 −r/2M
ds2 = dv 2 − du2 − r2 dΩ 2

e
r
du
du − dv 2 = 0 ⇒
2
= ±1
dv
In a maximal spacetime, particles can appear or disappear, but only at sin-
gularities.
200 11 Solutions of Einstein’s Equations & Black Holes

Fig. 11.5. Spacetime diagram in Kruskal coordinates

11.1.3 Wormholes (*)

A hypothetical ”tunnel” connecting two different points in spacetime in such


a way that a trip through the wormhole could take much less time than a
journey between the same starting and ending points in normal space. The
ends of a wormhole could, in theory, be intra-universe (i.e. both exist in the
same universe) or inter-universe (exist in different universes, and thus serve
as a connecting passage between the two).

Fig. 11.6. Embedding diagrams for wormholes

11.2 Reissner-Nordstrøm Solution (1916-18) (*)

Solution of Einstein-Maxwell equations with a static radial with a potential


11.3 Kerr Solution (1963) (**) 201
 
Q
Aµ = , 0, 0, 0 (11.1)
r
where Q is the total charge measured by a distant observer
−1
Q2 Q2
  
2M 2M
ds2 = − 1 − + 2 dt2 + 1 − + 2 dr2 + r2 dΩ 2 (11.2)
r r r r
The horizon is for g00 = 0
p
r± = M ± M 2 − Q2 (11.3)

two horizons r− and r+ .

11.3 Kerr Solution (1963) (**)


∆ 2 2 2 sin2 θ  2 2 ρ2 2
ds2 = − 2
+ dr + ρ2 dθ2

dt − a sin θdφ + (r + a )dφ − adt
ρ2 ρ2 ∆
1
where
∆ = r2 − 2M r + a2 and ρ2 = r2 + a2 cos2 θ (11.4)
here a = J/M = GJ/M c3 is the angular momentum per unit mass. For our
Sun J = 1.6 × 1048 g cm2 /s which corresponds to a = 0.185. Obviously for
a = 0 Kerr metric reduces to Schwarszchild.
The coordinates are known as Boyer-Lindquist coordinates and they are
related to Cartesian coordinates

x = (r2 + a2 )1/2 sin θ cos φ


y = (r2 + a2 )1/2 sin θ sin φ
z = r cos θ

11.3.1 Infinite redshift surface


2M r
g00 = 1 − ρ2 = 0 i.e. r2 + a2 cos2 θ − 2M r = 0 which leads to
p
R± = M ± M 2 − a2 cos2 θ

ds2 ≡ 0 = g00 dt2 + grr dr2 + gφφ dφ2 + 2g0φ dφdt at θ = π/2
 2  
dφ dφ
for constant r i.e. dr = 0 we get 0 = g00 + gφφ dt + 2g0φ dt then at
g00 = 0 we will get
1
If ∆ = r2 − 2M r + a2 + Q2 the we get the so called Kerr-Newman solution which
describes a stationary, axially symmetric and charged spacetime.
202 11 Solutions of Einstein’s Equations & Black Holes

Fig. 11.7. The light cone in the Kerr spacetime

dφ dφ gφφ
=0 and =− (11.5)
dt dt 2g0φ
The light cone in the radial direction:
 2
2 2 dr g00 ∆(ρ2 − 2M r)
0 = g00 dt + grr dr ⇒ =− = (11.6)
dt grr ρ4
thus the horizon(s) will be the surfaces:
p
∆ = r2 − 2M r + a2 = 0 ⇒ r± = M ± M 2 − a2 (11.7)

Fig. 11.8. Schematic Diagram of the Kerr black hole

The area between r+ ≤ r ≤ R+ is called ergosphere.

11.3.2 Penrose Process (1969)


Particles inside the ergosphere may have negative energy
11.3 Kerr Solution (1963) (**) 203

E = p0 = mg0µ uµ = m g00 u0 + g0φ uφ



(11.8)
Inside the ergosphere g00 < 0 while g0φ > 0 thus with an appropriate choice
of u0 and uφ it is possible to have particles with negative energy.
This can lead into energy extraction from a Kerr BH. Actually, we can
extract rotational energy till the BH reduces to a Schwarzschild one.
Starting from outside the ergosphere particle m0 enters into the ergosphere
and splits into two parts. One of them follows a trajectory of negative energy
and falls into the BH, the second escapes at infinity carrying more energy
than the initial particle!

Fig. 11.9. Schematic Diagram of the Penrose process

11.3.3 Black-Hole: Mechanics


There is a lower limit on the mass that can remain after a continuous extrac-
tion of rotational energy via Penrose process. Christodoulou (1969) named it
irreducible mass Mir
 p 
2 2
4Mir = r+ + a2 ≡ M + M 2 − a2 + a2 = 2M r+ (11.9)

while Hawking (‘70) suggested that the area of the BH horizon cannot be
reduced via the Penrose process. The two statement are equivalent!

EXAMPLE: Calculate the area of the horizon


(r2 + a2 )2 sin2 θ
ds2 = gθθ dθ2 + gφφ dφ2 = r+
2
+ a2 cos2 θ dθ2 + +2

r+ + a2 cos2 θ
thus the horizon area is
Z 2π Z π 2π π
√ √
Z Z

A= gdθdφ = gφφ gθθ dθdφ
0 0 0 0
Z 2π Z π
2
= (r+ + a2 ) sin θdθdφ = 4π(r+
2
+ a2 ) = 8πM r+ = 16πMir
2
0 0
204 11 Solutions of Einstein’s Equations & Black Holes

11.4 The Tolman-Oppenheimer-Volkov (TOV) Solution


(-)

The energy momentum tensor for a perfect fluid is:


dxµ
T µν = (ρ + p)uµ uν − g µν p where uµ =
ds
Then the law for conservation of energy and momentum leads to:
∂p µν ∂
T µν ;ν = g + [(p + ρ)uµ uν ] + (ρ + p)Γ µ νλ uµ uλ = 0 (11.10)
∂xν ∂xν
For a spherically symmetric and static solution

ds2 = eν(r) dt2 − eλ(r) dr2 − r2 dθ2 + sin2 θdφ2




the fluid velocity is: uµ = (e−ν/2 , 0, 0, 0) and thus (11.10) becomes

∂p µν
T µν ;ν = g + (ρ + p)Γ µ 00 u0 u0 = 0 (11.11)
∂xν
where Γ µ 00 = − 12 g µν g00,ν = − 21 eν ν 0
...multiplying with gµλ we get

∂p 1 ∂ν dp 1 dν
λ
= − (ρ + p) λ ⇒ = − (ρ + p) (11.12)
∂x 2 ∂x dr 2 dr
which is the relativistic version of the equations for hydrodynamical equilib-
rium, since in the Newtonian limit g00 = eν ≈ 1 + 2U and ρ >> p which
leads to:
∂p ∂U
= −ρ (11.13)
∂r ∂r
Still we need to find a way, via Einstein’s equations
 
1
Rµν = −8π Tµν − gµν Tλλ (11.14)
2

to estimate ν(r) in the same way that we need to solve Poisson equation to
estimate the gravitaional potential U .

h r i
{θθ} : 1 − e−λ 1 + (ν 0 − λ0 ) = −4πr2 (ρ − p) (11.15)
2
ν 00 (ν 0 )2 ν 0 λ0 λ0
{rr} : + − − = 4πeλ (ρ − p) (11.16)
2 4 4 r
ν 00 (ν 0 )2 ν 0 λ0 ν0
{tt} : + − + = −4πeλ (3ρ + p) (11.17)
2 4 4 r
11.4 The Tolman-Oppenheimer-Volkov (TOV) Solution (-) 205
0
which leads to the following ODE rλ0 + eλ − 1 = 8πeλ r2 ρ or re−λ =
1 − 8πρr2 leading to
Z r
−1
eλ(r) = (1 − 2M/r) where M (r) = 4π ρ(r0 )r02 dr0 (11.18)
0

which together with (11.12) give

dp M + 4πr3 p dν M + 4πr3 p
= −(ρ + p) or =2 . (11.19)
dr r(r − 2M ) dr r(r − 2M )

11.4.1 TOV : A uniform density star

For the special case ρ = const there is an analytic solution. For example the
mass function will become:
4 3 4 3
M (r) = πr ρ for r≤R and M (r) = πR ρ for r≥R
3 3
The we get:
p (1 − 2M r2 /R2 )1/2 − (1 − 2M/R)1/2
= . (11.20)
ρ 3(1 − 2M/R)1/2 − (1 − 2M r2 /R3 )1/2
But substituting the above relations in (11.12) we get an analytic solution
for g00 i.e.
1/2 1/2
2M r2
 
3 2M 1
eν/2 = 1− − 1− . (11.21)
2 R 2 R3
Maximum allowed mass when p(r = 0) → ∞ :

M 4
=
R 9
A
Useful constants in geometrical units

Speed of light c =299,792.458 km/s = 1


Planck’s constant ~ = 1.05 × 10−27 erg ·s = 2.612 × 10−66 cm2
Gravitation constant G = 6.67 × 10−8 cm3 /g ·s2 =1
Energy eV=1.602×10−12 erg = 1.16 × 104 K
=1.7823×10−33 g =1.324×10−56 km
Distance 1 pc=3.09×1013 km=3.26 ly
Time 1 yr = 3.156×107 sec
Light year 1 ly = 9.46×1012 km
Astronomical unit (AU) 1AU = 1.5 × 108 km
Earth’s mass M⊕ = 5.97 × 1027 g
Earth’s radius (equator) R⊕ = 6378 km
Solar Mass M = 1.99 × 1033 g =1.47664 km
Solar Radius R = 6.96 × 105 km

Das könnte Ihnen auch gefallen