Você está na página 1de 187

Lecture Notes

On
General Relativity

yvind Grn

Oslo College, Department of engineering, Cort Adelers gt. 30, N-0254 Oslo,
Norway
and
Department of Physics, University of Oslo, Box 1048 Blindern, N-0316, Norway
May 6, 2005
Preface
These notes are a transcript of lectures delivered by yvind Grn during the
spring of 1997 at the University of Oslo. Two compendia, (Grn and Fl 1984)
and (Ravndal 1978) were provided by Grn as additional reference material
during the lectures.
The present version of this document is an extended and corrected version of
a set of Lecture Notes which were typesetted by S. Bard, Andreas O. Jaunsen,
Frode Hansen and Ragnvald J. Irgens using LATEX2. Svend E. Hjelmeland has
made many useful suggestions which have improved the text.
While we hope that these typeset notes are of benefit particularly to stu-
dents of general relativity and look forward to their comments, we welcome all
interested readers and accept all feedback with thanks.
All comment may be sent to the author either by e-mail or snail mail.
yvind Grn
Fysisk Institutt
Universitetet i Oslo
P.O.Boks 1048, Blindern
0315 OSLO
E-mail: Oyvind.Gron@iu.hio.no
Contents

List of Figures v

List of Definitions ix

List of Examples xi

1 Newtons law of universal gravitation 1


1.1 The force law of gravitation . . . . . . . . . . . . . . . . . . . . . 1
1.2 Newtons law of gravitation in its local form . . . . . . . . . . . . 2
1.3 Tidal Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 The Principle of Equivalence . . . . . . . . . . . . . . . . . . . . 9
1.5 The general principle of relativity . . . . . . . . . . . . . . . . . . 10
1.6 The covariance principle . . . . . . . . . . . . . . . . . . . . . . . 11
1.7 Machs principle . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Vectors, Tensors and Forms 13


2.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 4-vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.2 Tangent vector fields and coordinate vectors . . . . . . . . 17
2.1.3 Coordinate transformations . . . . . . . . . . . . . . . . . 20
2.1.4 Structure coefficients . . . . . . . . . . . . . . . . . . . . . 23
2.2 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.1 Transformation of tensor components . . . . . . . . . . . . 27
2.2.2 Transformation of basis 1-forms . . . . . . . . . . . . . . . 27
2.2.3 The metric tensor . . . . . . . . . . . . . . . . . . . . . . 28
2.3 Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3 Accelerated Reference Frames 34


3.1 Rotating reference frames . . . . . . . . . . . . . . . . . . . . . . 34
3.1.1 Space geometry . . . . . . . . . . . . . . . . . . . . . . . . 34
3.1.2 Angular acceleration in the rotating frame . . . . . . . . . 38
3.1.3 Gravitational time dilation . . . . . . . . . . . . . . . . . 40
3.1.4 Path of photons emitted from axes in the rotating refer-
ence frame (RF) . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.5 The Sagnac effect . . . . . . . . . . . . . . . . . . . . . . . 42
3.2 Hyperbolically accelerated reference frames . . . . . . . . . . . . 43

i
4 Covariant Differentiation 49
4.1 Differentiation of forms . . . . . . . . . . . . . . . . . . . . . . . . 49
4.1.1 Exterior differentiation . . . . . . . . . . . . . . . . . . . . 49
4.1.2 Covariant derivative . . . . . . . . . . . . . . . . . . . . . 51
4.2 The Christoffel Symbols . . . . . . . . . . . . . . . . . . . . . . . 53
4.3 Geodesic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.4 The covariant Euler-Lagrange equations . . . . . . . . . . . . . . 57
4.5 Application of the Lagrangian formalism to free particles . . . . . 59
4.5.1 Equation of motion from Lagranges equation . . . . . . . 60
4.5.2 Geodesic world lines in spacetime . . . . . . . . . . . . . . 61
4.5.3 Gravitational Doppler effect . . . . . . . . . . . . . . . . . 70
4.6 The Koszul connection . . . . . . . . . . . . . . . . . . . . . . . . 71
4.7 Connection coefficients and structure coefficients c in ... . 74
4.8 Covariant differentiation of vectors, forms and tensors . . . . . . 75
4.8.1 Covariant differentiation of a vector in an arbitrary basis . 75
4.8.2 Covariant differentiation of forms . . . . . . . . . . . . . . 75
4.8.3 Generalization for tensors of higher rank . . . . . . . . . . 77
4.9 The Cartan connection . . . . . . . . . . . . . . . . . . . . . . . . 77

5 Curvature 81
5.1 The Riemann curvature tensor . . . . . . . . . . . . . . . . . . . 81
5.2 Differential geometry of surfaces . . . . . . . . . . . . . . . . . . . 86
5.2.1 Surface curvature, using the Cartan formalism . . . . . . . 89
5.3 The Ricci identity . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.4 Bianchis 1st identity . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.5 Bianchis 2nd identity . . . . . . . . . . . . . . . . . . . . . . . . 91

6 Einsteins Field Equations 93


6.1 Energy-momentum conservation . . . . . . . . . . . . . . . . . . . 93
6.1.1 Newtonian fluid . . . . . . . . . . . . . . . . . . . . . . . . 93
6.1.2 Perfect fluids . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.2 Einsteins curvature tensor . . . . . . . . . . . . . . . . . . . . . . 95
6.3 Einsteins field equations . . . . . . . . . . . . . . . . . . . . . . . 96
6.4 The geodesic postulate as a consequence of the field equations . 98

7 The Schwarzschild spacetime 100


7.1 Schwarzschilds exterior solution . . . . . . . . . . . . . . . . . . . 100
7.2 Radial free fall in Schwarzschild spacetime . . . . . . . . . . . . . 104
7.3 Light cones in Schwarzschild spacetime . . . . . . . . . . . . . . . 106
7.4 Analytical extension of the Schwarzschild spacetime . . . . . . . . 108
7.5 Embedding of the Schwarzschild metric . . . . . . . . . . . . . . . 110
7.6 Deceleration of light . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.7 Particle trajectories in Schwarzschild 3-space . . . . . . . . . . . 112
7.7.1 Motion in the equatorial plane . . . . . . . . . . . . . . . 114
7.8 Classical tests of Einsteins general theory of relativity . . . . . . 116
7.8.1 The Hafele-Keating experiment . . . . . . . . . . . . . . . 116

ii
7.8.2 Mercurys perihelion precession . . . . . . . . . . . . . . . 118
7.8.3 Deflection of light . . . . . . . . . . . . . . . . . . . . . . . 119

8 Black Holes 122


8.1 Surface gravity:gravitational acceleration on the horizon of a
black hole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
8.2 Hawking radiation:radiation from a black hole (1973) . . . . . . . 123
8.3 Rotating Black Holes: The Kerr metric . . . . . . . . . . . . . . . 124
8.3.1 Zero-angular-momentum-observers (ZAMOs) . . . . . . . 125
8.3.2 Does the Kerr space have a horizon? . . . . . . . . . . . . 126

9 Schwarzschilds Interior Solution 128


9.1 Newtonian incompressible star . . . . . . . . . . . . . . . . . . . 128
9.2 The pressure contribution to the gravitational mass of a static,
spherical symmetric system . . . . . . . . . . . . . . . . . . . . . 130
9.3 The Tolman-Oppenheimer-Volkov equation . . . . . . . . . . . . 131
9.4 An exact solution for incompressible stars - Schwarzschilds inte-
rior solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

10 Cosmology 135
10.1 Comoving coordinate system . . . . . . . . . . . . . . . . . . . . 135
10.2 Curvature isotropy - the Robertson-Walker metric . . . . . . . . 136
10.3 Cosmic dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
10.3.1 Hubbles law . . . . . . . . . . . . . . . . . . . . . . . . . . 137
10.3.2 Cosmological redshift of light . . . . . . . . . . . . . . . . 137
10.3.3 Cosmic fluids . . . . . . . . . . . . . . . . . . . . . . . . . 139
10.3.4 Isotropic and homogeneous universe models . . . . . . . . 140
10.4 Some cosmological models . . . . . . . . . . . . . . . . . . . . . . 143
10.4.1 Radiation dominated model . . . . . . . . . . . . . . . . . 143
10.4.2 Dust dominated model . . . . . . . . . . . . . . . . . . . . 144
10.4.3 Friedmann-Lematre model . . . . . . . . . . . . . . . . . 148
10.5 Inflationary Cosmology . . . . . . . . . . . . . . . . . . . . . . . . 158
10.5.1 Problems with the Big Bang Models . . . . . . . . . . . . 158
10.5.2 Cosmic Inflation . . . . . . . . . . . . . . . . . . . . . . . 161

Bibliography 166

iii
List of Figures

1.1 Newtons law of universal gravitation . . . . . . . . . . . . . . . . 1


1.2 Newtons law of gravitation in its local form . . . . . . . . . . . . 2
1.3 The definition of solid angle d . . . . . . . . . . . . . . . . . . . 5
1.4 Tidal Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 A small Cartesian coordinate system at a distance R from a mass
M. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 An elastic, circular ring falling freely in the Earths gravitational
field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 Closed polygon (linearly dependent) . . . . . . . . . . . . . . . . 13


2.2 Carriage at rest (top) and with velocity ~v (bottom) . . . . . . . . 14
2.3 World-lines in a Minkowski diagram . . . . . . . . . . . . . . . . 16
2.4 No position vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Tangentplane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Proper time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.7 Coordinate transformation, flat space. . . . . . . . . . . . . . . . 22
2.8 Basis-vectors ~e1 and ~e2 . . . . . . . . . . . . . . . . . . . . . . . . 29
2.9 The covariant- and contravariant components of a vector . . . . . 30

3.1 Simultaneity in rotating frames . . . . . . . . . . . . . . . . . . . 36


3.2 Rotating system: Distance between points on the circumference . 36
3.3 Rotating system: Discontinuity in simultaneity . . . . . . . . . . 37
3.4 Rotating system: Angular acceleration . . . . . . . . . . . . . . . 39
3.5 Rotating system: Distance increase . . . . . . . . . . . . . . . . . 39
3.6 Rotating system: Lorentz contraction . . . . . . . . . . . . . . . . 40
3.7 The Sagnac effect . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.8 Hyperbolic acceleration . . . . . . . . . . . . . . . . . . . . . . . 44
3.9 Simultaneity and hyperbolic acceleration . . . . . . . . . . . . . . 46
3.10 The hyperbolically accelerated reference system . . . . . . . . . . 48

4.1 Parallel transport . . . . . . . . . . . . . . . . . . . . . . . . . . . 55


4.2 Different world-lines connecting P 1 and P2 in a Minkowski diagram 57
4.3 Geodesic on a flat surface . . . . . . . . . . . . . . . . . . . . . . 59
4.4 Geodesic on a sphere . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.5 Timelike geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.6 Projectiles in 3-space . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.7 Geodesics in rotating reference frames . . . . . . . . . . . . . . . 64

v
4.8 Coordinates on a rotating disc . . . . . . . . . . . . . . . . . . . . 65
4.9 Projectiles in accelerated frames . . . . . . . . . . . . . . . . . . . 66
4.10 The twin paradox . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.11 Rotating coordinate system . . . . . . . . . . . . . . . . . . . . . 72

5.1 ~ . . . . . . . . .
Parallel transport of A . . . . . . . . . . . . . . . 81
5.2 Parallel transport of a vector along a triangle of angles 90 is
rotated 90 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.3 Geometry of parallel transport . . . . . . . . . . . . . . . . . . . 83
5.4 Surface geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

7.1 Light cones in Schwarzschild spacetime . . . . . . . . . . . . . . . 107


7.2 Light cones in Schwarzschild spacetime . . . . . . . . . . . . . . . 107
7.3 Embedding of the Schwarzschild metric . . . . . . . . . . . . . . . 111
7.4 Deceleration of light . . . . . . . . . . . . . . . . . . . . . . . . . 111
7.5 Newtonian centrifugal barrier . . . . . . . . . . . . . . . . . . . . 115
7.6 Gravitational collapse . . . . . . . . . . . . . . . . . . . . . . . . 116
7.7 Deflection of light . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

8.1 Static border and horizon of a Kerr black hole . . . . . . . . . . . 127

9.1 Hydrostatic equilibrium . . . . . . . . . . . . . . . . . . . . . . . 129

10.1 Schematic representation of cosmological redshift . . . . . . . . . 138


10.2 Expansion of a radiation dominated universe . . . . . . . . . . . 144
10.3 The size of the universe . . . . . . . . . . . . . . . . . . . . . . . 147
10.4 Expansion factor . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
10.5 The expansion factor as function of cosmic time in units of the
age of the universe. . . . . . . . . . . . . . . . . . . . . . . . . . . 151
10.6 The Hubble parameter as function of cosmic time. . . . . . . . . 151
10.7 ....... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
10.8 The deceleration parameter as function of cosmic time. . . . . . . 153
10.9 The ratio of the point of time when cosmic decelerations turn
over to acceleration to the age of the universe. . . . . . . . . . . . 154
10.10The cosmic red shift of light emitted at the turnover time from
deceleration to acceleration as function of the present relative
density of vacuum energy. . . . . . . . . . . . . . . . . . . . . . . 155
10.11The critical density in units of the constant density of the vacuum
energy as function of time. . . . . . . . . . . . . . . . . . . . . . . 155
10.12The relative density of the vacuum energy density as function of
time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
10.13The density of matter in units of the density of vacuum energy
as function of time. . . . . . . . . . . . . . . . . . . . . . . . . . . 157
10.14The relative density of matter as function of time. . . . . . . . . 157
10.15Rate of change of as function of ln( tt0 ). The value ln( tt0 ) =
40 corresponds to the cosmic point of time t 0 1s. . . . . . . . 159
10.16The shape of the potential depends on the sign of 2 . . . . . . . . 162

vi
10.17The temperature dependence of a Higgs potential with a first
order phase transition. . . . . . . . . . . . . . . . . . . . . . . . . 163

vii
List of Definitions

1.2.1 Solid angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4


2.1.1 4-velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.2 4-momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.3 4-acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.4 Reference frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.5 Coordinate system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.6 Comoving coordinate system . . . . . . . . . . . . . . . . . . . . . . 18
2.1.7 Orthonormal basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.8 Coordinate basis vectors. . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.9 Coordinate basis vectors. . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.10Orthonormal basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.11Commutators between vectors . . . . . . . . . . . . . . . . . . . . . . 23
2.1.12Structure coefficients c . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2.1 Multilinear function, tensors . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.2 Tensor product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.3 The metric tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2.4 Contravariant components . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3.1 p-form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2.1 Born-stiff motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3.1 Geodesic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.6.1 Koszuls connection coeffecients in an arbitrary basis . . . . . . . . . 71
4.8.1 Covariant derivative of a vector . . . . . . . . . . . . . . . . . . . . . 75
4.8.2 Covariant directional derivative of a one-form field . . . . . . . . . . 75
4.8.3 Covariant derivative of a one-form . . . . . . . . . . . . . . . . . . . 76
4.8.4 Covariant derivative of a tensor . . . . . . . . . . . . . . . . . . . . . 77
4.9.1 Exterior derivative of a basis vector . . . . . . . . . . . . . . . . . . . 77
4.9.2 Connection forms . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.9.3 Scalar product between vector and 1-form . . . . . . . . . . . . . . . 78
5.5.1 Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
7.1.1 Physical singularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.1.2 Coordinate singularity . . . . . . . . . . . . . . . . . . . . . . . . . . 104
8.3.1 Horizon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

ix
List of Examples

2.1.1 Photon clock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14


2.1.2 Coordinate transformation . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.3 Relativistic Doppler Effect . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1.4 Structure coefficients in planar polar coordinates . . . . . . . . . . . 25
2.2.1 Example of a tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.2 A mixed tensor of rank 3 . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2.3 Cartesian coordinates in a plane . . . . . . . . . . . . . . . . . . . . 29
2.2.4 Basis-vectors in plane polar-coordinates . . . . . . . . . . . . . . . . 29
2.2.5 Non-diagonal basis-vectors . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.6 Cartesian coordinates in a plane . . . . . . . . . . . . . . . . . . . . 31
2.2.7 Plane polar coordinates . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.3.1 antisymmetric combinations . . . . . . . . . . . . . . . . . . . . . . . 32
2.3.2 antisymmetric combinations . . . . . . . . . . . . . . . . . . . . . . . 32
2.3.3 A 2-form in 3-space . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.1.1 Outer product of 1-forms in 3-space . . . . . . . . . . . . . . . . . . 50
4.1.2 The derivative of a vector field with rotation . . . . . . . . . . . . . . 51
4.2.1 The Christoffel symbols in plane polar coordinates . . . . . . . . . . 54
4.3.1 vertical motion of free particle in hyperb. acc. ref. frame . . . . . . . 56
4.5.1 How geodesics in spacetime can give parabolas in space . . . . . . . 61
4.5.2 Spatial geodesics described in the reference frame of a rotating disc. 62
4.5.3 Christoffel symbols in a hyperbolically accelerated reference frame . 65
4.5.4 Vertical projectile motion in a hyperbolically accelerated reference
frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.5.5 The twin paradox . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.5.6 Measurements of gravitational Doppler effects (Pound and Rebka 1960) 71
4.6.1 The connection coefficients in a rotating reference frame. . . . . . . . 72
4.6.2 Acceleration in a non-rotating reference frame (Newton) . . . . . . . 73
4.6.3 The acceleration of a particle, relative to the rotating reference frame 73
4.9.1 Cartan-connection in an orthonormal basis field in plane polar coord. 79
6.1.1 Energy momentum tensor for a Newtonian fluid . . . . . . . . . . . . 94
10.4.1Age-redshift relation for dust dominated universe with k = 0 . . . . 146

xi
Chapter 1

Newtons law of universal


gravitation

1.1 The force law of gravitation

r
F

Figure 1.1: Newtons law of universal gravitation states that the force between
two masses is attractive, acts along the line joining them and is inversely pro-
portional to the distance separating the masses.

M M
F~ = mG 3 ~r = mG 2 ~er (1.1)
r r
Let V be the potential energy of m (see figure 1.1). Then

V
F~ = V (~r), Fi = (1.2)
xi

For a spherical mass distribution: V (~r) = mG M r , with zero potential


infinitely far from the center of M . Newtons law of gravitation is valid for
small velocities, i.e. velocities much smaller than the velocity of light and
weak fields. Weak fields are fields in which the gravitational potential energy
of a test particle is very small compared to its rest mass energy. (Note that
here one is interested only in the absolute values of the above quantities and
not their sign).
M GM
mG  mc2 r  2 . (1.3)
r c

1
2 Chapter 1. Newtons law of universal gravitation

The Schwarzschild radius for an object of mass M is R s = 2GM c2


. Far
outside the Schwarzschild radius we have a weak field. To get a feeling for
magnitudes consider that Rs u 1 cm for the Earth which is to be compared
with RE u 6400 km. That is, the gravitational field at the Earths surface can
be said to be weak! This explains, in part, the success of the Newtonian theory.

1.2 Newtons law of gravitation in its local form

Let P be a point in the field (see figure 1.2) with position vector ~r = x i~ei and
0
let the gravitating point source be at r~0 = xi ~ei0 . Newtons law of gravitation
for a continuous distribution of mass is

Z
~r r~0 3 0
F~ = mG (r~0 ) d r
r |~r r~0 |3 (1.4)
= V (~r)

See figure (1.2) for symbol definitions.

P
~r r~0

~r

r~0

Figure 1.2: Newtons law of gravitation in its local form.


1.2 Newtons law of gravitation in its local form 3

Lets consider equation (1.4) term by term.


1 1
= ~ei
|~r r~0 | xi (xj xj 0 )(xj xj 0 )1/2
h j 0
i1/2
= ~ei (x xj )(xj xj 0 )
xi
1 xj h k 0
i3/2
= ~ei 2(xj xj 0 ) (x xk )(xk xk0 )
2 xi
0
(x xj ) ij
j (1.5)
= ~ei 3/2
[(xk xk0 )(xk xk0 )]
0
(xi xi )
= ~ei  3/2
(xj xj 0 )(xj xj 0 )
~r r~0
=
|~r r~0 |3
Now equations (1.4) and (1.5) together
Z
(r~0 ) 3 0
V (~r) = mG d r (1.6)
|~r r~0 |
Gravitational potential at point P :
Z
V (~r) (r~0 ) 3 0
(~r) = G d r
m |~r r~0 |
Z
~r r~0 3 0
(~r) = G (r~0 ) d r (1.7)
|~r r~0 |3
Z
~r r~0 3 0
(~r) = G (r~0 )
2
d r
|~r r~0 |3
The above equation simplifies considerably if we calculate the divergence in the
integrand. Note that
operates on ~r
~r r~0 ~r 1 only!
= + (~r r~0 )
|~r r~0 |3 |~r r~0 |3 |~r r~0 |3
3 3(~r r~0 )
= (~r r~0 )
|~r r~0 |3 |~r r~0 |5 (1.8)
3 3
=
|~r r~0 |3 |~r r~0 |3
=0 ~r 6= r~0
We conclude that the Newtonian gravitational potential at a point in a gravi-
tational field outside a mass distribution satisfies Laplaces equation

2 = 0 (1.9)
4 Chapter 1. Newtons law of universal gravitation

Digression 1.2.1 (Diracs delta function)


The Dirac delta function has the following properties:

1. (~r r~0 ) = 0 ~r 6= r~0


R
2. (~r r~0 )d3 r 0 = 1 when ~r = r~0 is contained in the integration domain. The
integral is identically zero otherwise.
R
3. f (r~0 )(~r r~0 )d3 r 0 = f (~r)

R ~0
A calculation of the integral ~r~r0 3 d3 r 0 which is valid also in the case where
|~rr |
the field point is inside the mass distribution is obtained through the use of
Gauss integral theorem:
Z I
~ 3 0
Ad r = A ~ d~s, (1.10)
v s

where s is the boundary of v (s = v is an area).

Definition 1.2.1 (Solid angle)

ds0
d (1.11)
|~r r~0 |2

where ds0 is the projection of the area ds0 normal to the line of sight. ds~ 0 is the
component vector of ds ~ 0 along the line of sight which is equal to the normal vector
0
of ds (see figure (1.3)).

Now, lets apply Gauss integral theorem.


Z I I I
~r r~0 3 0 ~r r~0 ~ ds0
d r = 0
ds = = d (1.12)
|~r r~0 |3 |~r r~0 |3 |~r r~0 |2
v s s s

So that,
Z (
~r r~0 3 0 4 if P is inside the mass distribution,
d r = (1.13)
|~r r~0 |3 0 if P is outside the mass distribution.
v

The above relation is written concisely in terms of the Dirac delta function:

~r r~0
= 4(~r r~0 ) (1.14)
|~r r~0 |3
1.3 Tidal Forces 5

ds~0

P
~r r~0
~rr~0 ~r
ds~0 = ds~0 d
|~rr~0 |

r~0

ds~0 normal to bounding surface

Figure 1.3: The solid angle d is defined such that the surface of a sphere
subtends 4 at the center

We now have
Z
~r r~0 3 0
2 (~r) = G (r~0 ) d r
|~r r~0 |3
Z
(1.15)
=G (r~0 )4(~r r~0 )d3 r 0

= 4G(~r )

Newtons theory of gravitation can now be expressed very succinctly indeed!

1. Mass generates gravitational potential according to

2 = 4G (1.16)

2. Gravitational potential generates motion according to

~g = (1.17)

where ~g is the field strength of the gravitational field.

1.3 Tidal Forces


Tidal force is difference of gravitational force on two neighboring particles in a
gravitational field. The tidal force is due to the inhomogeneity of a gravitational
field.
In figure 1.4 two points have a separation vector .~ The position vectors of 1
~ ~
and 2 are ~r and ~r + , respectively, where ||  |~r|. The gravitational forces on
6 Chapter 1. Newtons law of universal gravitation

2

1

F2

F1

Figure 1.4: Tidal Forces

a mass m at 1 and at 2 are F~ (~r) and F~ (~r + ~). By means of a Taylor expansion
~ we get for the i-component of the tidal force
to lowest order in ||
 
~ Fi
fi = Fi (~r + ) Fi (~r) = j . (1.18)
xj ~r

The corresponding vector equation is

f~ = (~ )~r F~ . (1.19)

Using that
F~ = m, (1.20)
the tidal force may be expressed in terms of the gravitational potential according
to
f~ = m(~ ). (1.21)
It follows that in a local Cartesian coordinate system, the i-coordinate of the
relative acceleration of the particles is
 2 
d2 i
2
= j . (1.22)
dt xi xj ~r

Let us look at a few simple examples. In the first one ~ has the same direction
as ~g . Consider a small Cartesian coordinate system at a distance R from a mass
M (see figure 1.5). If we place a particle of mass m at a point (0, 0, +z), it will,
according to eq. (1.1) be acted upon by a force

GM
Fz (+z) = m (1.23)
(R + z)2

while an identical particle at the origin will be acted upon by the force
GM
Fz (0) = m . (1.24)
R2
1.3 Tidal Forces 7

z


m 

Fz (+z)
     

    

     
    
























y
     

    

   

Fz (0)  


R


M


 

 

Figure 1.5: A small Cartesian coordinate system at a distance R from a mass


M.

If this little coordinate system is falling freely towards M , an observer at


the origin will say that the particle at (0, 0, +z) is acted upon by a force
GM
fz = Fz (z) Fz (0) 2mz (1.25)
R3
directed away from the origin, along the positive z-axis. We have assumed
z  R. This is the tidal force.
In the same way particles at the points (+x, 0, 0) and (0, +y, 0) are attracted
towards the origin by tidal forces
GM
fx = mx , (1.26)
R3
GM
fy = my 3 . (1.27)
R
Eqs. (1.25)(1.27) have among others the following consequence: If an elastic,
circular ring is falling freely in the Earths gravitational field, as shown in figure
1.6, it will be stretched in the vertical direction and compressed in the horizontal
direction.
In general, tidal forces cause changes of shape.
The tidal forces from the Sun and the Moon cause flood and ebb on the
Earth. Let us consider the effect due to the Moon. We then let M be the mass
of the Moon, and choose a coordinate system with origin at the Earths center.
The tidal force per unit mass at a point is the negative gradient of the tidal
potential
 
GM 1 2 1 2 GM
(~r) = 3 z x y = 3 r 2 (3 cos2 1),
2
(1.28)
R 2 2 2R
8 Chapter 1. Newtons law of universal gravitation

Figure 1.6: An elastic, circular ring falling freely in the Earths gravitational
field

where we have introduced spherical coordinates, z = r cos , x 2 + y 2 = r 2 sin2 ,


R is the distance between the Earth and the Moon, and the radius r of the
spherical coordinate is equal to the radius of the Earth.
The potential at a height h above the surface of the Earth has one term,
mgh, due to the attraction of the Earth and one given by eq. (1.28), due to the
attraction of the Moon. Thus,

GM 2
(r) = gh r (3 cos2 1). (1.29)
2R3

At equilibrium, the surface of the Earth will be an equipotential surface,


given by = constant. The height of the water at flood, = 0 or = , is
therefore
GM  r 2
hflood = h0 + , (1.30)
gR R
where h0 is an unknown constant. The height of the water at ebb ( = 2 or
= 3
2 ) is
1 GM  r 2
hebb = h0 . (1.31)
2 gR R
The height difference between flood and ebb is therefore

3 GM  r 2
h = . (1.32)
2 gR R

For a numerical result we need the following values:

MMoon = 7.35 1025 g, g = 9.81m/s2 , (1.33)


5
R = 3.85 10 km, rEarth = 6378km. (1.34)

With these values we find h = 53cm, which is typical of tidal height differences.
1.4 The Principle of Equivalence 9

1.4 The Principle of Equivalence


Galilei investigated experimentally the motion of freely falling bodies. He found
that they moved in the same way, regardless what sort of material they consisted
of and what mass they had.
In Newtons theory of gravitation mass appears in two different ways; as
gravitational mass, mG , in the law of gravitation, analogously to charge in
Coulombs law, and as inertial mass, m I in Newtons 2nd law.
The equation of motion of a freely falling particle in the field of gravity from
a spherical body with mass M then takes the form
d2~r mG M
2
= G ~r. (1.35)
dt mI r 3
The results of Galileis measurements imply that the quotient between gravita-
tional and inertial mass must be the same for all bodies. With a suitable choice
of units, we then obtain
mG = m I . (1.36)
Measurements performed by the Hungarian baron Etvs around the turn
of the century indicated that this equality holds with an accuracy better than
mI
108 . More recent experiments have given the result | m G
1| < 9 1013 .
Einstein assumed the exact validity of eq.(1.52). He did not consider this as
an accidental coincidence, but rather as an expression of a fundamental principle,
called the principle of equivalence.
A consequence of this principle is the possibility of removing the effect of
a gravitational force by being in free fall. In order to clarify this, Einstein
considered a homogeneous gravitational field in which the acceleration of gravity,
g, is independent of the position. In a freely falling, non-rotating reference frame
in this field, all free particles move according to
d2~r
mI = (mG mI )~g = 0, (1.37)
dt2
where eq. (1.36) has been used.
This means that an observer in such a freely falling reference frame will say
that the particles around him are not acted upon by forces. They move with
constant velocities along straight paths. In other words, such a reference frame
is inertial.
Einsteins heuristic reasoning suggests equivalence between inertial frames in
regions far from mass distributions, where there are no gravitational fields, and
inertial frames falling freely in a gravitational field. This equivalence between all
types of inertial frames is so intimately connected with the equivalence between
gravitational and inertial mass, that the term principle of equivalence is used
whether one talks about masses or inertial frames. The equivalence of different
types of inertial frames encompasses all types of physical phenomena, not only
particles in free fall.
The principle of equivalence has also been formulated in an opposite way.
An observer at rest in a homogeneous gravitational field, and an observer in
10 Chapter 1. Newtons law of universal gravitation

an accelerated reference frame in a region far from any mass distributions, will
obtain identical results when they perform similar experiments. An inertial
field caused by the acceleration of the reference frame, is equivalent to a field of
gravity caused by a mass distribution, as far is tidal effects can be ignored.

1.5 The general principle of relativity


The principle of equivalence led Einstein to a generalization of the special princi-
ple of relativity. In his general theory of relativity Einstein formulated a general
principle of relativity, which says that not only velocities are relative, but accel-
erations, too.
Consider two formulations of the special principle of relativity.

S1 All laws of Nature are the same (may be formulated in the same way) in all
inertial frames.

S2 Every inertial observer can consider himself to be at rest.

These two formulations may be interpreted as different formulations of a


single principle. But the generalization of S1 and S2 to the general case, which
encompasses accelerated motion and non-inertial frames, leads to two different
principles G1 and G2.

G1 The laws of Nature are the same in all reference frames.

G2 Every observer can consider himself to he at rest.

In the literature both G1 and G2 are mentioned as the general principle of


relativity. But G2 is a stronger principle (i.e. stronger restriction on natural
phenomena) than G1. Generally the course of events of a physical process
in a certain reference frame, depends upon the laws of physics, the boundary
conditions, the motion of the reference frame and the geometry of space-time.
The two latter properties are described by means of a metrical tensor. By
formulating the physical laws in a metric independent way, one obtains that G1
is valid for all types of physical phenomena.
Even if the laws of Nature are the same in all reference frames, the course of
events of a physical process will, as mentioned above, depend upon the motion
of the reference frame. As to the spreading of light, for example, the law is that
light follows null-geodesic curves (see ch. 4). This law implies that the path of
a light particle is curved in non-inertial reference frames and straight in inertial
frames.
The question whether G2 is true in the general theory of relativity has been
thoroughly discussed recently, and the answer is not clear yet.
1.6 The covariance principle 11

1.6 The covariance principle


The principle of relativity is a physical principle. It is concerned with physical
phenomena. This principle motivates the introduction of a formal principle,
called the covariance principle: The equations of a physical theory shall have
the same form in every coordinate system.
This principle is not concerned directly with physical phenomena. The
principle may be fulfilled for every theory by writing the equations in a form-
invariant i.e. covariant way. This may he done by using tensor (vector) quanti-
ties, only, in the mathematical formulation of the theory.
The covariance principle and the equivalence principle may be used to obtain
a description of what happens in the presence of gravitation. We then start
with the physical laws as formulated in the special theory of relativity. Then
the laws are written in a covariant form, by writing them as tensor equations.
They are then valid in an arbitrary, accelerated system. But the inertial field
(fictive force) in the accelerated frame is equivalent to a gravitational field. So,
starting with in a description referred to an inertial frame, we have obtained a
description valid in the presence of a gravitational field.
The tensor equations have in general a coordinate independent form. Yet,
such form-invariant, or covariant, equations need not fulfill the principle of rel-
ativity.
This is due to the following circumstances. A physical principle, for example
the principle of relativity, is concerned with observable relationships. Therefore,
when one is going to deduce the observable consequences of an equation, one
has to establish relations between the tensor-components of the equation and
observable physical quantities. Such relations have to be defined; they are not
determined by the covariance principle.
From the tensor equations, that are covariant, and the defined relations
between the tensor components and the observable physical quantities, one can
deduce equations between physical quantities. The special principle of relativity,
for example, demands that the laws which these equations express must be the
same with reference to every inertial frame
The relationships between physical quantities and tensors (vectors) are the-
ory dependent. The relative velocity between two bodies, for example, is a
vector within Newtonian kinematics. However, in the relativistic kinematics of
four-dimensional space-time, an ordinary velocity, which has only three com-
ponents, is not a vector. Vectors in space-time, so called 4-vectors, have four
components. Equations between physical quantities are not covariant in general.
For example, Maxwells equations in three-vector-form are not invariant un-
der a Galilei transformation. However, if these equations are rewritten in tensor-
form, then neither a Galilei transformation nor any other transformation will
change the form of the equations.
If all equations of a theory are tensor equations, the theory is said to be given
a manifestly covariant form. A theory that is written in a manifestly covariant
form, will automatically fulfill the covariance principle, but it need not fulfill
the principle of relativity.
12 Chapter 1. Newtons law of universal gravitation

1.7 Machs principle


Einstein gave up Newtons idea of an absolute space. According to Einstein all
motion is relative. This may sound simple, but it leads to some highly non-trivial
and fundamental questions.
Imagine that there are only two particles connected by a spring, in the
universe. What will happen if the two particles rotate about each other? Will
the spring be stretched due to centrifugal forces? Newton would have confirmed
that this is indeed what will happen. However, when there is no longer any
absolute space that the particles can rotate relatively to, the answer is not so
obvious. If we, as observers, rotate around the particles, and they are at rest,
we would not observe any stretching of the spring. But this situation is now
kinematically equivalent to the one with rotating particles and observers at rest,
which leads to stretching.
Such problems led Mach to the view that all motion is relative. The motion
of a particle in an empty universe is not defined. All motion is motion relatively
to something else, i.e. relatively to other masses. According to Mach this implies
that inertial forces must be due to a particles acceleration relatively to the great
masses of the universe. If there were no such cosmic masses, there would not
exist inertial forces, like the centrifugal force. In our example with two particles
connected by a string, there would not be any stretching of the spring, if there
were no cosmic masses that the particles could rotate relatively to.
Another example may be illustrated by means of a turnabout. If we stay
on this, while it rotates, we feel that the centrifugal forces lead us outwards.
At the same time we observe that the heavenly bodies rotate. According to
Mach identical centrifugal forces should appear if the turnabout is static and
the heavenly bodies rotate.
Einstein was strongly influenced by Machs arguments, which probably had
some influence, at least with regards to motivation, on Einsteins construction
of his general theory of relativity. Yet, it is clear that general relativity does not
fulfill all requirements set by Machs principle. For example there exist general
relativistic, rotating cosmological models, where free particles will tend to rotate
relative to the cosmic masses of the model.
However, some Machian effects have been shown to follow from the equations
of the general theory of relativity. For example, inside a rotating, massive
shell the inertial frames, i.e. the free particles, are dragged on and tend to
rotate in the same direction as the shell. This was discovered by Lense and
Thirring in 1918 and is therefore called the Lense-Thirring effect. More recent
investigations of this effect have, among others, lead to the following result (Brill
and Cohen 1966): A massive shell with radius equal to its Schwarzschild radius
has often been used as an idealized model of our universe. Our result shows
that in such models local inertial frames near the center cannot rotate relatively
to the mass of the universe. In this way our result gives an explanation in
accordance with Machs principle, of the fact that the fixed stars is at rest on
heaven as observed from an inertial reference frame.
Chapter 2

Vectors, Tensors and Forms

2.1 Vectors
An expression on the form a~e , where a , = 1, 2, ..., n are real numbers, is
known as a linear combination of the vectors ~e .
The vectors ~e1 , ..., ~en are said to be linearly independent if there does not
exist real numbers a 6= 0 such that a~e = 0.

Figure 2.1: Closed polygon (linearly dependent)

Geometrical interpretation: A set of vectors are linearly independent if it


is not possible to construct a closed polygon of the vectors (even by adjusting
their lengths).
A set of vectors ~e1 , . . . , ~en are said to be maximally linearly independent
if ~e1 , . . . , ~en , ~v are linearly dependent for all vectors ~v 6= ~e . We define the
dimension of a vector-space as the number of vectors in a maximally linearly
independent set of vectors of the space. The vectors ~e in such a set are known

13
14 Chapter 2. Vectors, Tensors and Forms

as the basis-vectors of the space.

~v + a~e = 0

~v = a~e (2.1)

The components of ~v are the numbers v defined by v = a ~v = v ~e .

2.1.1 4-vectors

4-vectors are vectors which exist in (4-dimensional) space-time. A 4-vector


equation represents 4 independent component equations.

L c

L c v

v t
2

Figure 2.2: Carriage at rest (top) and with velocity ~v (bottom)

Example 2.1.1 (Photon clock)


Carriage at rest:

2L
t0 =
c
2.1 Vectors 15

Carriage with velocity ~v :


q
2 (v t 2
2 ) +L
2
t =
c

c2 t2 = v 2 t2 + 4L2

2L 2L/c t0
t = =p =p (2.2)
2
c v 2 2
1 v /c 2 1 v 2 /c2

The proper time-interval is denoted by d (above it was denoted t 0 ). The


proper time-interval for a particle is measured with a standard clock which
follows the particle.

Definition 2.1.1 (4-velocity)

U~ = c dt ~et + dx ~ex + dy ~ey + dz ~ez , (2.3)


d d d d
where t is the coordinate time, measured with clocks at rest in the reference frame.

~ dx
U = U ~e = ~e , x = (ct, x, y, z), x0 ct
d
dt 1
= q (2.4)
d 1 v2
c2

~ = (c, ~v ), where ~v is the common 3-velocity of the particle.


U

Definition 2.1.2 (4-momentum)

P~ = m0 U
~, (2.5)
where m0 is the rest mass of the particle.
P~ = ( Ec , p~), where p~ = m0~v = m~v and E is the relativistic energy.

~
The 4-force or Minkowski-force F~ ddP and the common force f~ = d~p
dt .
Then
1
F~ = ( f~ ~v , f~) (2.6)
c
16 Chapter 2. Vectors, Tensors and Forms

ct

world line of a material particle

lightcone

tachyons, if they exist,


should have v > c

Figure 2.3: World-lines in a Minkowski diagram

Definition 2.1.3 (4-acceleration)

~
~ = dU
A (2.7)
d
The 4-velocity has the scalar value c so that
~ U
U ~ = c2 (2.8)

~ A
Definition of Eq. 2.8 gives U ~ = 0, which implies A ~ U
~ and that A~ is
space-like.
The line element for Minkowski space-time (flat space-time) with Cartesian
coordinates is
ds2 = c2 dt2 + dx2 + dy 2 + dz 2 (2.9)
In general relativity theory, gravitation is not considered a force. Gravitation
is instead described as motion in a curved space-time.
A particle in free fall, is in Newtonian gravitational theory said to be only
influenced by the gravitational force. According to general relativity theory the
particle is not influenced by any force.
Such a particle has no 4-acceleration. A ~ 6= 0 implies that the particle is not
in free fall. It is then influenced by non-gravitational forces.
One has to distinguish between observed acceleration, ie. common 3-acceleration,
and the absolute 4-acceleration.
2.1 Vectors 17

2.1.2 Tangent vector fields and coordinate vectors


In a curved space position vectors with finite length do not exist. (See figure
2.4).

N(North pole)

Figure 2.4: In curved space,vectors can only exist in tangent planes.The vectors

in the tangent plane of N,do not contain the vector N P (dashed line).

Different points in a curved space have different tangent planes. Finite vec-
tors do only exist in these tangent planes (See figure 2.5). However, infinitesimal
position vectors d~r do exist.

tangent plane of point P:

Figure 2.5: In curved space,vectors can only exist in tangent planes


18 Chapter 2. Vectors, Tensors and Forms

Definition 2.1.4 (Reference frame)


A reference frame is defined as a continuum of non-intersecting timelike world
lines in spacetime.

We can view a reference frame as a set of reference particles with a specified


motion. An inertial reference frame is a non-rotating set of free particles.

Definition 2.1.5 (Coordinate system)


A coordinate system is a continuum of 4-tuples giving a unique set of coordinates
for events in spacetime.

Definition 2.1.6 (Comoving coordinate system)


A comoving coordinate system in a frame is a coordinate system where the
particles in the reference frame have constant spatial coordinates.

Definition 2.1.7 (Orthonormal basis)


An orthonormal basis {~e } in spacetime is defined by

~et ~et = 1(c = 1)


(2.10)
~ei ~ej = ij

where i and j are space indices.

Definition 2.1.8 (Coordinate basis vectors.)


Temporary definition of coordinate basis vector:
Assume any coordinate system {x }.

~r
~e (2.11)
x

A vector field is a continuum of vectors in a space, where the components are


continuous and differentiable functions of the coordinates. Let ~v be a tangent
vector to the curve ~r():

d~r
~v = where ~r = ~r[x ()] (2.12)
d
2.1 Vectors 19

The chain rule for differentiation yields:


d~r ~r dx dx
~v = = = e~ = v ~e (2.13)
d x d d
Thus, the components of the tangent vector field along a curve, parameterised
by , is given by:
dx
v = (2.14)
d
In the theory of relativity, the invariant parameter is often chosen to be the
proper time. Tangent vector to the world line of a material particle:
dx
u = (2.15)
d
These are the components of the 4-velocity of the particle!

Digression 2.1.1 (Proper time of the photon.)


Minkowski-space:

ds2 = c2 dt2 + dx2


1 dx 2 
= c2 dt2 1 2 (2.16)
c dt
v2  2 2
= 1 2 c dt
c
For a photon,v = c so:
lim ds2 = 0 (2.17)
vc

Thus, the spacetime interval between two points on the world line of a photon, is
zero! This also means that the proper time for the photon is zero!! (See example
2.1.2).

Digression 2.1.2 (Relationships between spacetime intervals, time and proper time.)
Physical interpretation of the spacetime interval for a timelike interval:

ds2 = c2 d 2 (2.18)
where d is the proper time interval between two events, measured on a clock
moving in a way, such that it is present on both events (figure 2.6).
v2 
c2 d 2 = c2 1 2 dt2
r c
v 2
d = 1 2 dt (2.19)
c
20 Chapter 2. Vectors, Tensors and Forms

ct

d P2

P1

Figure 2.6: P1 and P2 are two events in spacetime, separated by a proper time
interval d .

The time interval between to events in the laboratory, is smaller measured on a


moving clock than measured on a stationary one, because the moving clock is
ticking slower!

2.1.3 Coordinate transformations


Given two coordinate systems {x } and {x0 }.

~r
~e0 = (2.20)
x0

Suppose there exists a coordinate transformation, such that the primed coor-
dinates are functions of the unprimed, and vice versa. Then we can apply the
chain rule:

~r ~r x x
~e0 = 0 = 0 = ~
e 0
(2.21)
x x x x

This is the transformation equation for the basis vectors. x
x
0
are elements
of the transformation matrix. Indices that are not sum-indices are called free
indices.

Rule: In all terms on each side in an equation, the free indices should
behave identically (high or low), and there should be exactly the same
indices in all terms!
2.1 Vectors 21

Applying this rule, we can now find the inverse transformation


0
x
~e = ~e0
x
0 x 0
~v = v ~e0 = v ~e = v ~e
x0
So, the transformation rules for the components of a vector becomes

x 0
0 0 x
v =v ; v =v (2.22)
x0 x
The directional derivative along a curve, parametrised by :
d dx
=
= v (2.23)
d x d x

where v = dxd are the components of the tangent vector of the curve. Direc-
tional derivative along a coordinate curve:
x
= x = = (2.24)
x x x x
In the primed system:
x
0 = (2.25)
x x0 x

Definition 2.1.9 (Coordinate basis vectors.)


We define the coordinate basis vectors as:


~e = (2.26)
x

This definition is not based upon the existence of finite position vectors. It applies
in curved spaces as well as in flat spaces.

Example 2.1.2 (Coordinate transformation)


From Figure 2.7 we see that

x = r cos , y = r sin (2.27)

Coordinate basis vectors were defined by



e~ (2.28)
x
22 Chapter 2. Vectors, Tensors and Forms

y
e

er
y

r
ey


x
ex x

Figure 2.7: Coordinate transformation, flat space.

This means that we have



e~x = , e~y = , e~r = , e~ =
x y r
(2.29)
x y
e~r = = +
r r x r y

Using the chain rule and Equations (2.27) and (2.29) we get

e~r = cos e~x + sin e~y


x y
e~ = = + (2.30)
x y
= r sin e~x + r cos e~y

But are the vectors in (2.30) also unit vectors?

e~r e~r = cos2 + sin2 = 1 (2.31)

So e~r is a unit vector, |e~r | = 1.

e~ e~ = r 2 (cos2 + sin2 ) = r 2 (2.32)

and we see that e~ is not a unit vector, |e~ | = r. But we have that e~r e~ = 0
e~r e~ . Coordinate basis vectors are not generally unit vectors.
2.1 Vectors 23

Definition 2.1.10 (Orthonormal basis)


An orthonormal basis is a vector basis consisting of unit vectors that are normal to
each other. To show that we are using an orthonormal basis we will use hats over
the indices, {~e }.

Orthonormal basis associated with planar polar coordinates:


1
~er = ~er , ~e = ~e (2.33)
r

Example 2.1.3 (Relativistic Doppler Effect)


The Lorentz transformation is known from special relativity and relates the reference
frames of two systems where one is moving with a constant velocity v with regard
to the other,

x0 = (x vt)
vx
t0 = (t 2 )
c
According to the vector component transformation (2.22), the 4-momentum for a
particle moving in the x-direction, P = ( Ec , p, 0, 0) transforms as
0
0 x
P = P ,
x
E 0 = (E vp).

Using the fact that a photon has energy E = h and momemtum p = h c , where
h is Plancks constant and is the photons frequency, we get the equation for the
frequency shift known as the relativistic Doppler effect,

0 v 1 vc
= ( ) = q  
c 1 v 1+ v c c

r
0 cv
= (2.34)
c+v

2.1.4 Structure coefficients


Definition 2.1.11 (Commutators between vectors)
The commutator between two vectors, ~u and ~v , is defined as

[~u , ~v ] ~u~v ~v ~u (2.35)


24 Chapter 2. Vectors, Tensors and Forms

where ~u~v is defined as



~u~v u e~ (v e~ ) = u
(v ) (2.36)
x x

We can think of a vector as a linear combination of partial derivatives. We get:


v 2
~u~v = u + u
v
x x x x (2.37)
v 2
= u e~ + u v
x x x
Due to the last term, ~u~v is not a vector.

~v~u = v (u )
x x
u 2
= v e~ + v u
x x x

(2.38)
v u
~u~v ~v~u = u
e~ v e~
x | x{z }

v u e~
x

v u
= (u v )e~
x x
Here we have used that
2 2

= (2.39)
x x x x
The Einstein comma notation
~u~v ~v~u = (u v , v u, )e~ (2.40)
As we can see, the commutator between two vectors is itself a vector.

Definition 2.1.12 (Structure coefficients c )


The structure coefficients c in an arbitrary basis {e~ } are defined by:
[e~ , e~ ] c e~ (2.41)

Structure coefficients in a coordinate basis:



[e~ , e~ ] = [ , ]
x x

=
( ) ( ) (2.42)
x x x x
2 2
= =0
x x x x
2.2 Tensors 25

The commutator between two coordinate basis vectors is zero, so the structure
coefficients are zero in coordinate basis.

Example 2.1.4 (Structure coefficients in planar polar coordinates)


We will find the structure coefficients of an orthonormal basis in planar polar coor-
dinates. In (2.33) we found that
1
~er = ~er , ~e = ~e (2.43)
r
We will now use this to find the structure coefficients.
1
[~er , ~e ] = [ , ]
r r
1 1
= ( ) ( )
r r r r (2.44)
1 1 2 1 2
= 2 +
r r r r r
1 1
= 2 ~e = ~e
r r
To find the structure coefficients in coordinate basis we must use [~e r , ~e ] = 1r ~e .

[~e , ~e ] = c ~e (2.45)

Using (2.44) and (2.45) we get


1
cr = (2.46)
r
From the definition of c ([~u , ~v ] = [~v , ~u]) we see that the structure coefficients
are anti symmetric in their lower indices:

c = c (2.47)

1
cr = = cr (2.48)
r

2.2 Tensors
A 1-form-basis 1 , . . ., n is defined by:

(e~ ) = (2.49)

An arbitrary 1-form can be expressed, in terms of its components, as a linear


combination of the basis forms:

= (2.50)
26 Chapter 2. Vectors, Tensors and Forms

where are the components of in the given basis.


Using eqs.(2.49) and (2.50), we find:

(~e ) = (~e ) = =
(2.51)
(~v ) = (v ~e ) = v (~e ) = v = v 1 1 + v 2 2 + . . .

We will now look at functions of multiple variables.

Definition 2.2.1 (Multilinear function, tensors)


A multilinear function is a function that is linear in all its arguments and maps
one-forms and vectors into real numbers.

A covariant tensor only maps vectors.

A contravariant tensor only maps forms.

A mixed tensor maps both vectors and forms into R.



A tensor of rank NN0 maps N one-forms and N 0 vectors into R. It is usual to
say that a tensor is of rank (N + N 0 ). A one-form, for example, is a covariant
tensor of rank 1:

(~v ) = v (2.52)

Definition 2.2.2 (Tensor product)


The basis of a tensor R of rank q contains a tensor product, . If T and S are
two tensors of rank m and n, the tensor product is defined by:

T S(u~1 ,..., u~m , v~1 ,..., v~n ) T (u~1 ,..., u~m )S(v~1 ,..., v~n ) (2.53)

where T and S are tensors of rank m and n, respectively. T S is a tensor of rank


(m + n).

Let R = T S. We then have

R = R1 ,...,q 1 2 q (2.54)

Notice that S T 6= T S. We get the components of a tensor (R) by using


the tensor on the basis vectors:

R1 ,...,q = R(e~1 ,..., e~q ) (2.55)

The indices of the components of a contravariant tensor are written as upper


indices, and the indices of a covariant tensor as lower indices.
2.2 Tensors 27

Example 2.2.1 (Example of a tensor)


Let ~u and ~v be two vectors and and two 1-forms.

~u = u~e ; ~v = v ~e ; = ; = (2.56)

From these we can construct tensors of rank 2 through the relation R = ~u ~v as


follows: The components of R are

R1 2 = R( 1 , 2 )
= ~u ~v ( 1 , 2 )
= ~u( 1 )~v ( 2 )
(2.57)
= u~e ( 1 )v ~e ( 2 )
= u 1 v 2
= u 1 v 2

2.2.1 Transformation of tensor components


We shall not limit our discussion to coordinate transformations. Instead,
we
will consider arbitrary transformations between bases, {~e } ~e0 . The
elements of transformation matrices are denoted by M 0 such that
0
~e0 = ~e M 0 and ~e = ~e0 M (2.58)
0
where M are elements of the inverse transformation matrix. Thus, it follows
that
0
M 0 M = (2.59)

If the transformation is a coordinate transformation, the elements of the matrix


become
0
0 x
M = (2.60)
x

2.2.2 Transformation of basis 1-forms


0 0
= M
0 (2.61)
= M 0

The components of a tensor of higher rank transform such that every con-
travariant index (upper) transforms as a basis 1-form and every covariant index
(lower) as a basis vector. Also, all elements of the transformation matrix are
multiplied with one another.
28 Chapter 2. Vectors, Tensors and Forms

Example 2.2.2 (A mixed tensor of rank 3)

0 0
T 0 0 = M M 0 M 0 T (2.62)

The components in the primed basis are linear combinations of the components
in the unprimed basis.

Tensor transformation of components means that tensors have a basis in-


dependent existence. That is, if a tensor has non-vanishing components in a
given basis then it has non-vanishing components in all bases. This means
that tensor equations have a basis independent form. Tensor equations are
invariant. A basis transformation might result in the vanishing of one or more
tensor components. Equations in component form may differ from one basis to
another. But an equation expressed in tensor components can be transformed
from one basis to another using the tensor component transformation rules. An
equation that is expressed only in terms of tensor components is said to be
covariant.

2.2.3 The metric tensor


Definition 2.2.3 (The metric tensor)
The scalar product of two vectors ~u and ~v is denoted by g(~u,~v ) and is defined as
a symmetric linear mapping which for each pair of vectors gives a scalar g(~v ,~u) =
g(~u,~v ).

The value of the scalar product g(~u,~v ) is given by specifying the scalar
products of each pair of basis-vectors in a basis.
g is a symmetric covariant tensor of rank 2. This tensor is known as the
metric tensor. The components of this tensor are

g(~e , ~e ) = g (2.63)

~u ~v = g(~u , ~v ) = g(u~e , v ~e ) = u v g(~e , ~e ) = u u g (2.64)

Usual notation:

~u ~v = g u v (2.65)

The absolute value of a vector:


p q
|~v | = g(~v , ~v ) = |g v v | (2.66)
2.2 Tensors 29

e2


e1

Figure 2.8: Basis-vectors ~e1 and ~e2

Example 2.2.3 (Cartesian coordinates in a plane)

~ex ~ex = 1, ~ey ~ey = 1, ~ex ~ey = ~ey ~ex = 0


g xx = g yy = 1, g xy = g yx = 0 (2.67)
 
1 0
g =
0 1

Example 2.2.4 (Basis-vectors in plane polar-coordinates)

~er ~er = 1, ~e ~e = r 2 , ~er ~e = 0, (2.68)


The metric tensor in plane polar-coordinates:
 
1 0
g = (2.69)
0 r2

Example 2.2.5 (Non-diagonal basis-vectors)

~e1 ~e1 = 1, ~e2 ~e2 = 1, ~e1 ~e2 = cos = ~e2 ~e1


   
1 cos g 11 g 12 (2.70)
g = =
cos 1 g 21 g 22

Definition 2.2.4 (Contravariant components)


The contravariant components g of the metric tensor are defined as:
g g g = w
~ w
~, (2.71)
~ is defined by
where w
~ w
w ~ . (2.72)
g is the inverse matrix of g .
30 Chapter 2. Vectors, Tensors and Forms

A22
A2
x2 = constant
e2

A2 e 2 A x1 = constant

e1
A1 e 1

A1

A1 1
1

Figure 2.9: The covariant- and contravariant components of a vector

It is possible to define a mapping between tensors of different type (eg.


covariant on contravariant) using the metric tensor.
We can for instance map a vector on a 1-form:

v = g(~v , ~e ) = g(v ~e , ~e ) = v g(~e , ~e ) = v g (2.73)

This is known as lowering of an index. Raising of an index becomes :

v = g v (2.74)

The mixed components of the metric tensor becomes:

g = g g = (2.75)

We now define distance along a curve. Let the curve be parameterized by


(proper-time for time-like curves). Let ~v be the tangent vector-field of the
curve.
The squared distance ds2 between the points along the curve is defined as:

ds2 g(~v , ~v )d2 (2.76)

gives

ds2 = g v v d2 . (2.77)
2.3 Forms 31

dx
The tangent vector has components v = d , which gives:

ds2 = g dx dx (2.78)

The expression ds2 is known as the line-element.

Example 2.2.6 (Cartesian coordinates in a plane)

g xx = g yy = 1, g xy = g yx = 0
(2.79)
ds2 = dx2 + dy 2

Example 2.2.7 (Plane polar coordinates)

g rr = 1, g = r 2
(2.80)
ds2 = dr 2 + r 2 d 2

Cartesian coordinates in the (flat) Minkowski space-time :

ds2 = c2 dt2 + dx2 + dy 2 + dz 2 (2.81)

In an arbitrary curved space, an orthonormal basis can be adopted in any


point. If ~et is tangent vector to the world line of an observer, then ~e t = ~u
where ~u is the 4-velocity of the observer. In this case, we are using what we call
the comoving orthonormal basis of the observer. In a such basis, we have the
Minkowski-metric:
ds2 = dx dx (2.82)

2.3 Forms
An antisymmetric tensor is a tensor whose sign changes under an arbitrary
exchange of two arguments.

A( , ~u, , ~v , ) = A( , ~v , , ~u, ) (2.83)

The components of an antisymmetric tensor change sign under exchange of


two indices.

A = A (2.84)
32 Chapter 2. Vectors, Tensors and Forms

Definition 2.3.1 (p-form)


A p-form is defined to be an antisymmetric, covariant tensor of rank p.
An antisymmetric tensor product is defined by:
(p + q)! [1
[1 p ] [1 q ] q ] (2.85)
p!q!

where [ ] denotes antisymmetric combinations defined by:


1
[1 p ] (the sum of terms with
p!
all possible permutations (2.86)
of indices with, + for even
and - for odd permutations)

Example 2.3.1 (antisymmetric combinations)

1
[1 2 ] = ( 1 2 2 1 ) (2.87)
2

Example 2.3.2 (antisymmetric combinations)

[1 2 3 ] =
1 1
( 2 3 + 3 1 2 + 2 3 1
3!
2 1 3 3 2 1 1 3 2 )
1
= ijk ( i j k ) (2.88)
3!

Example 2.3.3 (A 2-form in 3-space)

= 12 1 2 +21 2 1 +13 1 3 +31 3 1 +23 2 3 +32 3 2


(2.89)

Now the antisymmetry of means that

+21 = 12 ; +31 = 13 ; +32 = 23 (2.90)


2.3 Forms 33

=12 ( 1 2 2 1 )
+ 13 ( 1 3 3 1 )
(2.91)
+ 23 ( 2 3 3 2 )
= || 2 [ ]

where || means summation only for < (see (Misner, Thorne and Wheeler
1973)). We now use the definition of with p = q = 1. This gives

= ||
is the
We can also write form basis.

1
=
2

A tensor of rank 2 can always be split up into a symmetric and an antisym-


metric part.
1 1
T = (T T ) + (T + T )
2 2 (2.92)
= A + S

We thus have:
1
S A = (T + T )(T T )
4
1 (2.93)
= (T T T T + T T T T )
4
=0

In general, summation over indices of a symmetric and an antisymmetric quan-


tity vanishes. In a summation T A where A is antisymmetric and T has
no symmetry, only the antisymmetric part of T contributes. So that, in

1
= (2.94)
2
only the antisymmetric elements = , contribute to the summation.
These antisymmetric elements are the form components
Forms are antisymmetric covariant tensors. Because of this antisymmetry
a form with two identical components must be a null form (= zero). e.g.
131 = 131 131 = 0
In an n-dimensional space all p-forms with p > n are null forms.
Chapter 3

Accelerated Reference Frames

3.1 Rotating reference frames


3.1.1 Space geometry
Let ~e0 be the 4-velocity field (x0 = ct, c = 1, x0 = t) of the reference particles in
a reference frame R. A set of simultanous events in R, defines a 3 dimensional
space called 3-space in R. This space is orthogonal to ~e 0 . We are going to
find the metric tensor ij in this space, expressed by the metric tensor g of
spacetime.
In an arbitrary coordinate basis {~e }, {~ei } is not necessarily orthogonal to
~e0 . We choose ~e0 k~e0 . Let ~ei be the projection of ~ei orthogonal to ~e0 , that
is:~ei ~e0 = 0. The metric tensor of space is defined by:

ij = ~ei ~ej , i0 = 0, 00 = 0

~ei = ~ei ~eki


~ei ~e0 gi0
~eki = ~e0 = ~e0
~e0 ~eo g00
ij = (~ei ~eki ) (~ej ~ekj )
gi0 gj0
= (~ei ~e0 ) (~ej ~e0 )
g00 g00
gj0 gi0 gi0 gj0
= ~ei ~ej ~e0 ~ei ~e0 ~ej + 2 ~e0 ~e0
g00 g00 g00
gi0 gj0 gi0 gj0 gi0 gj0
= gij +
g00 g00 g00

gi0 gj0
ij = gij (3.1)
g00

(Note:gij = gji ij = ji )
The line element in space:
gi0 gj0  i j
dl2 = ij dxi dxj = gij dx dx (3.2)
g00

34
3.1 Rotating reference frames 35

gives the geometry of a simultaneity space in a reference frame where the


metric tensor of spacetime in a comoving coordinate system is g .
The line element for spacetime can be expressed as:

ds2 = dt2 + dl2 (3.3)

It follows that dt = 0 represents the simultaneity defining the 3-space with


metric ij .

dt2 = dl2 ds2 = ( g )dx dx


= (ij gij )dxi dxj + 2(i0 gi0 )dxi dx0 + (00 g00 )dx0 dx0
gi0 gj0
= (gij gij )dxi dxj 2gi0 dxi dx0 g00 (dx0 )2
g00
 
gi0 0 i gi0 gj0 i j
= g00 (dx0 )2 + 2 dx dx + 2 dx dx
g00 g00
 2
gi0 i
= (g00 )1/2 (dx0 + dx )
g00

So finally we get
gi0 i
dt = (g00 )1/2 (dx0 + dx ) (3.4)
g00

The 3-space orthogonal to the world lines of the reference particles in R, d t = 0,


corresponds to a coordinate time interval dt = gg00 i0
dxi . This is not an exact
differential , that is , the line integral of dt around a closed curve is in general not
equal to 0. Hence you can not in general define simultaneity (given by d t = 0)
around closed curves. This can only be done if the spacetime metric is diagonal,
gi0 = 0. The condition dt = 0 means simultaneity on Einstein synchronized
clocks . Conclusion:It is in general (g i0 6= 0) not possible to Einstein
synchronize clocks around closed curves.
In particular, it is not possible to Einstein-synchronize clocks around a closed
curve in a rotating reference frame. If this is attempted, contradictory bound-
ary conditions in the non-rotating lab frame will arise, due to the relativity of
simultaneity. (See figure 3.1)
The distance in the laboratory frame between two points is:

2r
L0 = (3.5)
n
Lorentz transformation from the instantaneous rest frame (x 0 , t0 ) to the labo-
ratory system (x, t):

v 1
t = (t0 + x0 ), =q
c2 2 2
1 r c2 (3.6)
x = (x0 + vt0 )
36 Chapter 3. Accelerated Reference Frames

t+2 t
S2
3
t+ t
2
S1

t & t+n t (discontinuity)

r
Sn

n
t+(n-1) t

Sn-1
n-1

Figure 3.1: Events simultanous in the rotating reference frame. 1 comes before
2, before 3, etc. . . Note the discontinuity at t.

L0 v = r

Figure 3.2: The distance between two points on the circumference is L 0 .


3.1 Rotating reference frames 37

 
  


Figure 3.3: Discontinuity in simultaneity.

Since we for simultaneous events in the rotating reference frame have t 0 = 0,


and proper distance x0 = L0 , we get in the laboratory frame
r r 2r
t = 2 2
L0 = 2 2 (3.7)
c c n
The fact that t0 = 0 and t 6= 0 is an expression of the relativity of simul-
taneity. Around the circumference this is accumulated to

2r 2
nt = 2 (3.8)
c2
and we get a discontinuity in simultaneity, as shown in figure 3.3. Let IF be an
inertial frame with cylinder coordinates (T, R, , Z). The line element is then
given by

ds2 = dT 2 + dR2 + R2 d2 + dZ 2 (c = 1) (3.9)

In a rotating reference frame, RF, we have cylinder coordinates (t, r, , z). We


then have the following coordinate transformation :

t = T, r = R, = T, z=Z (3.10)

The line element in the co-moving coordinate system in RF is then

ds2 = dt2 + dr 2 + r 2 (d + dt)2 + dz 2


(3.11)
= (1 r 2 2 )dt2 + dr 2 + r 2 d 2 + dz 2 + 2r 2 ddt (c = 1)
38 Chapter 3. Accelerated Reference Frames

The metric tensor have the following components:

gtt = (1 r 2 2 ), grr = 1, g = r 2 , gzz = 1


(3.12)
gt = gt = r 2

dt = 0 gives

ds2 = dr 2 + r 2 d 2 + dz 2 (3.13)

This represents the Euclidean geometry of the 3-space (simultaneity space, t =


T ) in IF.
The spatial geometry in the rotating system is given by the spatial line
element:
gi0 gj0
dl2 = (gij )dxi dxj
g00
rr = grr = 1, zz = gzz = 1,
2
g0
= g
g00
(r 2 )2 r2
= r2 =
(1 r 2 2 ) 1 r2 2

r 2 d 2
dl2 = dr 2 + + dz 2 (3.14)
1 r2 2
So we have a non Euclidean spatial geometry in RF. The circumference of a
circle with radius r is
2r
l = > 2r (3.15)
1 r2 2
We see that the quotient between circumference and radius > 2 which means
that the spatial geometry is hyperbolic. (For spherical geometry we have l <
2r.)

3.1.2 Angular acceleration in the rotating frame


We will now investigate what happens when we give RF an angular acceleration.
Then we use a rotating circle made of standard measuring rods, as shown in
Figure 3.4. All points on a circle are accelerated simultaneously in IF (the
laboratory system). We let the angular velocity increase from to + d,
measured in IF. Lorentz transformation to an instantaneous rest frame for a
point on the circumference then gives an increase in velocity in this system:

rd
rd 0 = , (3.16)
1 r2 2
where we have used that the initial velocity in this frame is zero.
3.1 Rotating reference frames 39

"nail"

Standard measuring rod

Figure 3.4: The standard measuring rods are fastened with nails in one end.
We will see what then happens when we have an angular acceleration.

The time difference for the accelerations of the front and back ends of the
rods (the front end is accelerated first) in the instantaneous rest frame is:

rL0
t0 = (3.17)
1 r2 2

where L0 is the distance between points on the circumference when at rest (= the
length of the rods when at rest), L0 = 2r
n . In IF all points on the circumference
are accelerated simultaneously. In RF, however, this is not the case. Here the
distance between points on the circumference will increase, see Figure 3.5. The
rest distance increases by

r 2 L0 d
dL0 = rd 0 t0 = . (3.18)
(1 r 2 2 )3/2

The increase of the distance during the acceleration (in an instantaneous

t+ t t
dv
Figure 3.5: In RF two points on the circumference are accelerated at different
times. Thus the distance between them is increased.
40 Chapter 3. Accelerated Reference Frames

rest frame) is
Z
0 2 d 1
L = r L0 2 2 3/2
= ( 1)L0 . (3.19)
0 (1 r ) 1 r2 2
Hence, after the acceleration there is a proper distance L 0 between the rods. In
the laboratory system (IF) the distance between the rods is
p p 1 p
L= 1 r 2 2 L0 = 1 r 2 2 ( 1)L0 = L0 L0 1 r 2 2 ,
1 r2 2
(3.20)

where L0 is the rest length of the rods and L0 1 r 2 2 is their Lorentz con-
tracted length. We now have the situation shown in Figure 3.6.

Standard measuring rod,


Lorenz contracted

Figure 3.6: The standard measuring rods have been Lorentz contracted.

Thus, there is room for more standard rods around the periphery the faster
the disk rotates. This means that the measured length of the periphery (number
of standard rods) gets larger with increasing angular velocity.

3.1.3 Gravitational time dilation

r2 2 2 2
ds2 = (1 )c dt + dr 2 + r 2 d 2 + dz 2 + 2r 2 ddt (3.21)
c2
We now look at standard clocks with constant r and z.
r2 2 r 2 d 2 r 2 d
ds2 = c2 dt2 [(1 ) + ( ) + 2 ] (3.22)
c2 c2 dt c2 dt
3.1 Rotating reference frames 41

Let d
dt be the angular velocity of the clock in RF. The proper time interval
measured by the clock is then

ds2 = c2 d 2 (3.23)

From this we see that


s
r 2 2 r 2 2 r 2
d = dt 1 2 (3.24)
c2 c2 c2

A non-moving standard clock in RF: = 0


r
r2 2
d = dt 1 2 (3.25)
c
Seen from IF, the non-rotating laboratory system, (3.25) represents the velocity
dependent time dilation from the special theory of relativity.
But how is (3.25) interpreted in RF? The clock does not move relative to
an observer in this system, hence what happens can not bee interpreted as a
velocity dependent phenomenon. According to Einstein, the fact that standard
clocks slow down the farther away from the axis of rotation they are, is due to
a gravitational effect.
We will now find the gravitational potential at a distance r from the axis.
The sentripetal acceleration is v 2 /r, v = r so:
Z r Z r
1
= g(r)dr = r 2 dr = r 2 2
0 0 2

We then get:
r r
r2 2 2
d = dt 1 2 = dt 1+ (3.26)
c c2
In RF the position dependent time dilation is interpreted as a gravitational
time dilation: Time flows slower further down in a gravitational field.

3.1.4 Path of photons emitted from axes in the rotating refer-


ence frame (RF)
We start with description in the inertial frame (IF). In IF photon paths are
radial. Consider a photon path with = 0, R = T with light source at R = 0.
Transforming to RF:

t = T, r = R, = T
(3.27)
r = t, = t

The orbit equation is thus = r which is the equation for an Archimedean


spiral. The time used by a photon out to distance r from axis is t = rc .
42 Chapter 3. Accelerated Reference Frames

3.1.5 The Sagnac effect


IF description:
Here the velocity of light is isotropic, but the emitter/receiver moves due to the
discs rotation as shown in Figure 3.7. Photons are emitted/received in/from
opposite directions. Let t1 be the travel time of photons which move with the
rotation.
Emitter/Receiver
X

r
+

Figure 3.7: The Sagnac effect demonstrates the anisotropy of the speed of light
when measured in a rotating reference frame.

Then
2r + rt1 = ct1
2r (3.28)
t1 =
c r
Let t2 be the travel time for photons moving against the rotation of the disc.
A is the area The difference in travel time is
enclosed by the  
1 1
photon path or t = t1 t2 = 2r
orbit. c r c + r
2r2r (3.29)
= 2
c r2 2
4A
= 2 2
c

RF description:
ds2 = 0 along the
world line of a  
2 r2 2 2 2
photon ds = 1 2 c dt + r 2 d 2 + 2r 2 ddt
c
d
let =
dt
r + 2r 2 (c2 r 2 2 ) = 0
2 2
p
r 2 (r 4 2 + r 2 c2 r 4 2 )
=
r2
3.2 Hyperbolically accelerated reference frames 43

rc
=
r2 (3.30)
c
=
r

The speed of light: v = r = r c. We see that in the rotating frame RF,


the measured (coordinate) velocity of light is NOT isotropic. The difference in
the travel time of the two beams is

2r 2r
t =
c r c + r (3.31)
4A
= 2 2
c

(See Phil. Mag. series 6, vol. 8 (1904) for Michelsons article)

3.2 Hyperbolically accelerated reference frames

Consider a particle moving along a straight line with velocity u and acceleration
du
a = dT . Rest acceleration is a.

3/2
a = 1 u2 /c2 a. (3.32)

Assume that the particle has constant rest acceleration a = g. That is

du 3/2
= 1 u2 /c2 g. (3.33)
dT

Which on integration with u(0) = 0 gives

gT dX
u=   1/2
=
g2 T 2 dT
1 + c2
 1/2
c2 g2 T 2
X= 1+ 2 +k
g c

c4
= (X k)2 c2 T 2 (3.34)
g2

In its final form the above equation describes a hyperbola in the Minkowski
diagram as shown in figure(3.8).
44 Chapter 3. Accelerated Reference Frames

cT

Figure 3.8: Hyperbolically accelerated reference frames are so called because


the loci of particle trajectories in space-time are hyperbolae.

The proper time interval as measured by a clock which follows the particle:
 1/2
u2
d = 1 2 dT (3.35)
c
Substitution for u(T ) and integration with (0) = 0 gives
 
c gT
= arcsinh
g c
c  g 
or T = sinh (3.36)
g c
c2  g 
and X = cosh +k
g c
We now use this particle as the origin of space in an hyperbolically accelerated
reference frame.

Definition 3.2.1 (Born-stiff motion)


Born-stiff motion of a system is motion such that every element of the system has
constant rest length. We demand that our accelerated reference frame is Born-stiff.

Let the inertial frame have coordinates (T, X, Y, Z) and the accelerated
frame have coordinates (t, x, y, z). We now denote the X-coordinate of the
origin particle by X0 .
gX0 g0
1+ 2
= cosh (3.37)
c c
3.2 Hyperbolically accelerated reference frames 45

c2
where 0 is the proper time for this particle and k is set to g . (These are
Mller coordinates. Setting k = 0 gives Rindler coordinates).
Let us denote the accelerated frame by . The coordinate time at an ar-
bitrary point in is defined by t = 0 . That is coordinate clocks in run
~ 0 be the posi-
identically with the standard clock at the origin particle. Let X
tion 4-vector of the origin particle. Decomposed in the laboratory frame, this
becomes

   
~0 = c2 gt c2 gt
X sinh , cosh 1 , 0, 0 (3.38)
g c g c

P is chosen such that P and P0 are simultaneous in the accelerated frame . The
distance (see figure(3.9)) vector from P 0 to P , decomposed into an orthonormal
comoving basis of the origin particle is X ~ = (0, x, y, z) where x, y and z are
physical distances measured simultaneously in . The space coordinates in
are defined by

x x, y y, z z. (3.39)

~ =X
The position vector of P is X ~ The relationship between basis vectors
~ 0 + X.
in IF and the comoving orthonormal basis is given by a Lorentz transformation
in the x-direction.

x
~e = ~e
x
cosh sinh 0 0
sinh cosh 0 0 (3.40)
= (~eT , ~eX , ~eY , ~eZ , )
0

0 1 0
0 0 0 1

where is the rapidity defined by

U0
tanh (3.41)
c

U0 being the velocity of the origin particle.

dX0 gt
U0 = = c tanh
dT0 c (3.42)
gt
=
c
46 Chapter 3. Accelerated Reference Frames

b
~ P
X

cT ~et
P0
~eX

~
X
~0
X

X
Figure 3.9: Simultaneity in hyperbolically accelerated reference frames. The
b
~ lies along the simultaneity line which makes the same angle with the
vector X
X-axis as does ~et with the cT-axis.

So the basis vectors can be written as follows

gt gt
~et = ~eT cosh + ~eX sinh
c c
gt gt
~ex = ~eT sinh + ~eX cosh (3.43)
c c
~ey = ~eY
~ez = ~eZ

~ =X
The equation X ~ can now be decomposed in IF:
~0 + X

T~eT + X~eX + Y ~eY + Z~eZ =


 
c gt c2 gt x gt gt
sinh ~eT + cosh 1 ~eX + sinh ~eT + x cosh ~eX + y~eY + z~eZ
g c g c c c c
(3.44)
3.2 Hyperbolically accelerated reference frames 47

This then, gives the coordinate transformations


c gt x gt
T = sinh + sinh
g c c c
2
 
c gt gt
X= cosh 1 + x cosh
g c c
Y =y
Z=z
gT  gx  gt
= 1 + 2 sinh
c c c
gX  gx  gt
1 + 2 = 1 + 2 cosh
c c c
Now dividing the last two of the above equations we get
 
gT gX gt
= 1 + 2 tanh (3.45)
c c c

showing that the coordinate curves t = constant are straight lines in the T,X-
2
frame passing through the point T = 0, X = cg . Using the identity cosh2
sinh2 = 1 we get
 2  2 
gX gT gx 2
1+ 2 = 1+ 2 (3.46)
c c c

showing that the coordinate curves x = constant are hypebolae in the T,X-
diagram.
The line element (the metric) gives : ds 2 is an
invariant
ds2 = c2 dT 2 + dX 2 + dY 2 + dZ 2 quantity
gx
= (1 + 2 )2 c2 dt2 + dx2 + dy 2 + dz 2 (3.47)
c
Note: When the metric is diagonal the unit vectors are orthogonal.
Clocks as rest in the accelerated system:

dx = dy = dz = 0, ds2 = c2 d 2


gx 2 2 2
c2 d 2 = (1 + ) c dt
c2

gx
d = (1 + )dt (3.48)
c2
Here d is the proper time and dt the coordinate time.
An observer in the accelerated system experiences a gravitational field in
the negative x-direction. When x < 0 then d < dt. The coordinate clocks
48 Chapter 3. Accelerated Reference Frames

cT

horizon

light t=constant

x=constant

X
-c 2
g

Figure 3.10: The hyperbolically accelerated reference system

tick equally fast independently of their position. This implies that time passes
slower further down in a gravitational field.
Consider a standard clock moving in the x-direction with velocity v = dx/dt.
Then
 gx 2
c2 d 2 = 1 + 2 c2 dt2 + dx2
 c  
gx 2 v 2 2 2
= 1+ 2 2 c dt (3.49)
c c

Hence r

gx 2 v 2
d = 1+ 2 dt (3.50)
c2 c
This expresses the combined effect of the gravitational- and the kinematic time
dilation.
Chapter 4

Covariant Differentiation

4.1 Differentiation of forms


We must have a method of differentiation that maintains the anti symmetry,
thus making sure that what we end up with after differentiation is still a form.

4.1.1 Exterior differentiation


The exterior derivative of a 0-form, i.e. a scalar function, f , is given by:
f
df = = f, (4.1)
x
where are coordinate basis forms:

( ) = (4.2)
x
We then (in general) get:
x
= = = dx (4.3)
x
In coordinate basis we can always write the basis forms as exterior derivatives
of the coordinates. The differential dx is given by

dx (d~r) = dx (4.4)

where d~r is an infinitesimal position vector. dx are not infinitesimal quantities.


In coordinate basis the exterior derivative of a p-form will have the following
component form:
1
d= p ,0 dx0 dx1 dxp (4.5)
p! 1
where , 0 x0 . The exterior derivative of a p-form is a (p + 1)-form.
Consider the exterior derivative of a p-form .
1
d = , dx0 dxp . (4.6)
p! 1 p 0

49
50 Chapter 4. Covariant Differentiation

Let (d)0 p be the form components of d. They must, by definition, be


antisymmetric under an arbitrary interchange of indices.

1
d = (d)0 p dx0 dxp
(p + 1)!
1
which, by (4.6) = [1 p ,0 ] dx0 dxp
p!

(d)0 p = (p + 1)[1 p ,0 ] (4.7)

The form equation d = 0 in component form is

[1 p ,0 ] = 0 (4.8)

Example 4.1.1 (Outer product of 1-forms in 3-space)

= i dxi xi = (x, y, z)
(4.9)
d = i,j dxj dxi

Also, assume that d = 0. The corresponding component equation is

[i,j] = 0 i,j j,i = 0


x y x z y z (4.10)
= 0, = 0, =0
y x z x z y

which corresponds to

~
=0 (4.11)

The outer product of an outer product!

d2 d(d)
1 (4.12)
d2 = , dx2 dx1 dxp
p! 1 p 1 2

2
,1 2 (4.13)
x1 x2

Since
2 2
,1 2 = ,2 1 (4.14)
x1 x2 x2 x1
4.1 Differentiation of forms 51

summation over 1 and 2 which are symmetric in 1 p ,1 2 and antisymmetric


in the basis we get Poincars lemma (valid only for scalar fields)

d2 = 0 (4.15)

This corresponds to the vector equation

~ =0
( A) (4.16)

Let be a p-form and be a q-form. Then

d( ) = d + (1)p d (4.17)

4.1.2 Covariant derivative


The general theory of relativity contains a covariance principle which states
that all equations expressing laws of nature must have the same form irrespective
of the coordinate system in which they are derived. This is achieved by writing
all equations in terms of tensors. Let us see if the partial derivative of vector
components transform as tensor components. Given a vector A ~ = A~e =
0
A ~e0 with the transformation of basis given by

x
= (4.18)
x0 x0 x
So that
0  0 
A, 0 A
x0
x  0 
= A
x0 x  
x x0
= A
x0 x x
0
x x0 x 2 x
= A + A (4.19)
x0 x , x0 x x
The first term corresponds to a tensorial transformation. The existence of the
last term shows that A, does not, in general, transform as the components of
a tensor. Note that A, will transform as a tensor under linear transformations
such as the Lorentz transformations.
The partial derivative must be generalized such as to ensure that when it is
applied to tensor components it produces tensor components.
52 Chapter 4. Covariant Differentiation

Example 4.1.2 (The derivative of a vector field with rotation)


We have a vector field:

~ = kr~e
A

The chain rule for derivation gives:

d dx
=
= u
d x d x

~
dA
= u (A~e ),
d 
= u A, ~e + A~e,

The change of the vector field with a displacement along a coordinate-curve is


expressed by:

A~
~ , = A ~e + A~e,
=A ,
x
~ with the displacement in the -direction is:
The change in A

~
A
= A,~e + A~e,

For our vector field, with Ar = 0, we get

~
A
= A, ~e + A ~e,
|{z}
=0

and since A, = 0 because A = kr we end up with

~
A
= A ~e, = kr~e,

4.2 The Christoffel Symbols 53

We now need to calculate the derivative of ~e . We have:

x = r cos y = r sin

Using ~e = x we can write:

x y
~e = = +
x y
= r sin ~ex + r cos ~ey


~er = = cos ~ex + sin ~ey
r
Gives:
~e, = r cos ~ex r sin ~ey
= r(cos ~ex + sin ~ey ) = r~er

This gives us finally:


~
A
= kr 2~er

~
A ~ = A~e and A = 0.
Thus 6= 0 even if A ,

4.2 The Christoffel Symbols


The covariant derivative was introduced by Christoffel to be able to differenti-
ate tensor fields. It is defined in coordinate basis by generalizing the partially
derivative A, to a derivative written as A; and which transforms tensorially,
0
0 x x
A; 0 A . (4.20)
x x 0 ;
The covariant derivative of the contravariant vector components are written as:

A; A, + A (4.21)

This equation defines the Christoffel symbols , which are also called the
connection coefficients in coordinate basis. From the transformation formulae
for the two first terms follows that the Christoffel symbols transform as:
0 0
0 x x x x 2 x
0 0 = + (4.22)
x x x
0 0
x x0 x 0
The Christoffel symbols do not transform as tensor components. It is possible to
cancel all Christoffel symbols by transforming into a locally Cartesian coordinate
54 Chapter 4. Covariant Differentiation

system which is co-moving in a locally non-rotating reference frame in free fall.


Such coordinates are known as Gaussian coordinates.
In general relativity theory an inertial frame is defined as a non-rotating
frame in free fall. The Christoffel symbols are 0 (zero) in a locally Cartesian
coordinate system which is co-moving in a local inertial frame. Local Gaussian
coordinates are indicated with a bar over the indices, giving

= 0 (4.23)

A transformation from local Gaussian coordinates to any coordinates leads to:


0
0 x 2 x
0 0 = (4.24)
x x0 x 0
This equation shows that the Christoffel symbols are symmetric in the two lower
indices, ie.
0 0
0 0 = 0 0 (4.25)

Example 4.2.1 (The Christoffel symbols in plane polar coordinates)

x = r cos , y = r sin
p y
r = x2 + y 2 , = arctan
x

x x r x r
= cos , = r sin = = cos , = sin
r x r y
y y sin cos
= sin , = r cos = , =
r x r y r

r 2 x r 2 y
r = +
x 2 y 2
= cos (r cos ) + sin (r sin )
= r(cos 2 + sin 2 ) = r

2 x 2 y
r = r = +
x r y r
sin cos
= ( sin ) + (cos )
r r
1
=
r
4.2 The Christoffel Symbols 55

The geometrical interpretation of the covariant derivative was given by Levi-


Civita.
Consider a curve S in any (eg. curved) space. It is parameterized by , ie.
x = x (). is invariant and chosen to be the curve length.
The tangent vector field of the curve is ~u = (dx /d)~e . The curve passes
~ The covariant directional derivative of the vector field
through a vector field A.
along the curve is defined as:

~
~ = dA A; dx ~e = A; u ~e
u~ A (4.26)
d d

The vectors in the vector field are said to be


connected by parallel transport along the curve
if
A; u = 0

A( +)
B

+
A ( + )
Q
A( )
u

~ = A; u ~e
Figure 4.1: Parallel transport from P to Q. The vector B

dx
~u = ~e (4.27)
d
According to the geometrical interpretation of Levi-Civita, the covariant direc-
tional derivative is:

~ k ( + ) A()
A ~
~ = A u ~e = lim
u~ A (4.28)
;
0

~ k ( + ) means the vector A


where A ~ parallel transported from Q to P .
56 Chapter 4. Covariant Differentiation

4.3 Geodesic curves


Definition 4.3.1 (Geodesic curves)
A geodesic curve is defined in such a way that,the vectors of the tangent vector field
of the curve is connected by parallell transport.

This definition says that geodesic curves are as straight as possible.


~
If vectors in a vector field A() are connected by parallell transport by a dis-
placement along a vector ~u , we have A ; u = 0. For geodesic curves, we then
have:
u; u = 0 (4.29)

which is the geodesic equation.

(u, + u )u = 0 (4.30)

d dx
Then we are using that d d x = u x :

du u
= u = u u, (4.31)
d x

The geodesic equation can also be written as:

du
+ u u = 0 (4.32)
d
d
Usual notation: = d
dx
u = = x (4.33)
d

x + x x = 0 (4.34)

By comparing eq.4.34 with the equation of motion(4.52) for a free particle (which
we deduced from the Lagrangian equations) , we find the equations to be iden-
tical. Conclusion:Free particles follow geodesic curves in spacetime.

Example 4.3.1 (vertical motion of free particle in hyperb. acc. ref. frame)
Inserting the Christoffel symbols xtt = (1 + gx c2
)g from example 4.5.3 into the
geodesic equation for a vertical geodesic curve in a hyperbolically accelerated refer-
ence frame, we get:
gx
x + (1 + 2 )g t2 = 0
c
4.4 The covariant Euler-Lagrange equations 57

4.4 The covariant Euler-Lagrange equations


Geodesic curves can also be defined as curves with an extremal distance between
two points. Let a particle have a world-line (in space-time) between two points
(events) P1 and P2 . Let the curves be described by an invariant parameter
(proper time is used for particles with a rest mass).
The Lagrange-function is a function of coordinates and their derivatives,
dx
L = L(x , x ), x . (4.35)
d
(Note: if = then x are theR 4-velocity components)
The action-integral is S = L(x , x )d. The principle of extremal action
(Hamiltons-principle): The world-line of a particle is determined by the condi-
tion that S shall be extremal for all infinitesimal variations of curves which keep
P1 and P2 rigid, ie.
Z 2
L(x , x )d = 0, (4.36)
1

where 1 and 2 are the parameter-values at P1 and P2 . For all the variations

ct

P2

P1

x
Figure 4.2: Different world-lines connecting P 1 and P2 in a Minkowski diagram

the following condition applies:

x (1 ) = x (2 ) = 0 (4.37)

We write Eq. (4.36) as


Z 2 Z 2  
L L
Ld = x + x d (4.38)
1 1 x x
58 Chapter 4. Covariant Differentiation

Partial integration of the last term


Z 2   Z 2  
L L 2 d L

x d = x x d (4.39)
1 x x 1 1 d x

Due to the conditions x (1 ) = x (2 ) = 0 the first term becomes zero. Then


we have : Z 2   
L d L
S =

x d (4.40)
1 x d x
The world-line the particle follows is determined by the condition S = 0 for
any variation x . Hence, the world-line of the particle must be given by
 
L d L

=0 (4.41)
x d x

These are the covariant Euler-Lagrange equations.


The canonical momentum p conjugated to a coordinate x is defined as

L
p (4.42)
x
The Lagrange-equations can now be written as

dp L L
= or p = . (4.43)
d x x

A coordinate which the Lagrange-function does not depend on is known as a


L
cyclic coordinate. Hence, x = 0 for a cyclic coordinate. From this follows:

The canonical momentum conjugated to a cyclic


coordinate is a constant of motion

ie. p = C (constant) if x is cyclic.


A free particle in space-time (curved space-time includes gravitation) has
the Lagrange function
1 1 1
L = ~u ~u = x x = g x x (4.44)
2 2 2
An integral of the Lagrange-equations is obtained readily from the 4-velocity
identity: (
x x = c2 for a particle with rest-mass
(4.45)
x x = 0 for light
The line-element is:

ds2 = g dx dx = g x x d2 = 2Ld2 . (4.46)

Thus the Lagrange function of a free particle is obtained from the line-element.
4.5 Application of the Lagrangian formalism to free particles 59

flat surface:
Q

Figure 4.3: On a flat surface, the geodesic curve is the minimal distance between
P and Q

sphere:

Figure 4.4: On a sphere, the geodesic curves are great circles.

4.5 Application of the Lagrangian formalism to free


particles
To describe the motion of a free particle, we start by setting up the line element
of the space-time in the chosen coordinate system. There are coordinates on
which the metric does not depend. For example, given axial symmetry we may
choose the angle which is a cyclic coordinate here and the conjugate (covariant)
impulse P is a constant of the motion (the orbital spin of the particle). If, in
addition, the metric is time independent (stationary metric) then t is also
cyclic and pt is a constant of the motion (the mechanical energy of the particle).
A static metric is time-independent and unchanged under time reversal
(i.e. t t). A stationary metric changed under time reversal. Examples
of static metrics are Minkowski and hyperbolically accelerated frames. The
rotating cylindrical coordinate system is stationary.
60 Chapter 4. Covariant Differentiation

4.5.1 Equation of motion from Lagranges equation


The Lagrange function for a free particle is:

1
L = g x x
2 (4.47)

where g = g (x ). And the Lagrange equations are


 
L d L
= 0,
x d x
L (4.48)
= g x ,
x
L 1
= g , x x .
x 2

   
d L L
= g x + g x
d x x (4.49)

= g , x x + g x .

Now, (4.49) and (4.48) together give:

1
g x x g , x x g x = 0. (4.50)
2 ,

The second term on the left hand side of (4.50) may be rewritten making use of
the fact that x x is symmetric in , as as follows

1
g , x x = (g + g , )x x
2 , (4.51)
1
g x + (g , + g , g , )x x = 0.
2

Finally, since we are free to multiply (4.51) through by g , we can isolate x


to get the equation of motion in a particularly elegant and simple form:

x + x x = 0 (4.52)

where the Christoffel symbols in (4.52) are defined by

1
g (g , + g , g , ). (4.53)
2

Equation(4.52) describes a geodesic curve .


4.5 Application of the Lagrangian formalism to free particles 61

4.5.2 Geodesic world lines in spacetime


Consider two timelike curves between two events in spacetime. In fig.4.5 they
are drawn in a Minkowski diagram which refers to an inertial reference frame.

cT

t1
P

non-geodetic curve between O and P

geodetic curve
v(t)

O
t X
0

Figure 4.5: Timelike curves in spacetime.

The spacetime distance between O and P (See figure 4.5) equals the proper
time interval between two events O and P measured on a clock moving in a such
way, that it is present both at O and P.
ds2 = c2 d 2
Z T1 r (4.54)
v 2 (T )
01 = 1 dT
T0 c2
We can see that 01 is maximal along the geodesic curve with v(T ) = 0. Time-
like geodesic curves in spacetime have maximal distance between two points.

Example 4.5.1 (How geodesics in spacetime can give parabolas in space)


A geodesic curve between two events O and P has maximal proper time. Consider
the last expression in Section 3.2 of the propertime interval of a particle with position
x and velocity v in a gravitational field with acceleration of gravity g.
r
gx 2 v 2
d = dt 1 + 2 2
c c
This expression shows that the proper time of the particle proceeds faster the higher
up in the field the particle is, and it proceeds slower the faster the particle moves.
62 Chapter 4. Covariant Differentiation

Consider figure 4.6. The path a free particle follows between the events O and P
is a compromize between moving as slowly as possible in space, in order to keep
the velocity dependent time dilation small, and moving through regions high up in
the gravitational field, in order to prevent the slow proceeding of proper time far
down. However if the particle moves too high up, its velocity becomes so large that
it proceeds slower again. The compromise between kinematic and gravitational time
dilation which gives maximal proper time between O and P is obtained for the thick
curve in fig. 4.6. This is the curve followed by a free particle between the events O
and P.
We shall now deduce the mathematical expression of what has been said above.
Timelike geodesic curves are curves with maximal proper time, i.e.
Z 1
p
= g x x d
0

is maximal for a geodesic curve. However the action


Z 1 Z 1
J = 2 Ld = g x x d
0 0

is maximal for the same curves and this gives an easier calculation.
In the case of a vertical curve in a hyperbolically accelerated reference frame the
Lagrangian is   
1 gx 2 2 x2
L= 1 + 2 t + 2 (4.55)
2 c c
Using the Euler-Lagrange equations now gives
gx 2
x + (1 + )g t = 0
c2
which is the equation of the geodesic curve in example 4.3.1.
Since spacetime is flat, the equation represents straight lines in spacetime. The
projection of such curves into the three space of arbitrary inertial frames gives
straight paths in 3-space, in accordance with Newtons 1st law. However projecting
it into an accelerated frame where the particle also has a horizontal motion, and
taking the Newtonian limit, one finds the parabolic path of projectile motion.

Example 4.5.2 (Spatial geodesics described in the reference frame of a rotating disc.)
In Figure 4.7, we see a rotating disc. We can see two geodesic curves between P 1
and P2 . The dashed line is the geodesic for the non-rotating disc. The other curve
is a geodesic for the 3-space of a rotating reference frame. We can see that the
geodesic is curved inward when the disc is rotating. The curve has to curve inward
since the measuring rods are longer there (because of Lorentz-contraction). Thus,
the minimum distance between P1 and P2 will be achieved by an inwardly bent
curve.
We will show this mathematically, using the Lagrangian equations. The line
element for the space dt = dz = 0 of the rotating reference frame is
4.5 Application of the Lagrangian formalism to free particles 63

The path of the particle


g

O P
Figure 4.6: The particle moves between two events O and P at fixed points in
time. The path chosen by the particle between O and P is such that the proper
time taken by the particle betweem these two events is as large as possible.
Thus the goal of the particle is to follow a path such that its comoving standard
clocks goes as fast as possible. If the particle follows the horizontal line between
O and P it goes as slowly as possible and the kinematical time dilation is as
small as possible. Then the slowing down of its comoving standard clocks due to
the kinematical time dilation is as small as possible, but the particle is far down
in the gravitational field and its proper time goes slowly for that reason. Paths
futher up leads to a greater rate of proper time. But above the curve drawn as
a thick line, the kinematical time dilation will dominate, and the proper time
proceeds more slowly.
64 Chapter 4. Covariant Differentiation

P
1

P
2

Figure 4.7: Geodesic curves on a non-rotating (dashed line) and rotating (solid line)
disc.

r 2 d 2
dl2 = dr 2 + r2 2
1 c2
Lagrangian function:
1 2 1 r 2 2
L= r +
2 2 1 r2 2 2
c
We will also use the identity:

r 2 2
r 2 + r2 2
=1 (4.56)
1 c2

(We got this from using ~u ~u = 1) We see that is cyclic ( L


= 0), implying:

L r 2
p = = r2 2
= constant
1 c2

This gives:
r 2 2  p p 2 p
= 1 = (4.57)
c2 r 2 r2 c2
Inserting 4.57 into 4.56:
2 p2 p2
r 2 = 1 +
2
2 (4.58)
c r
This gives us the equation of the geodesic curve between P 1 and P2 :
q
2 2 p2 p2
r dr r 1 + c2 r2
= = 2 2 (4.59)
d p 1 r c2

Boundary conditions:
4.5 Application of the Lagrangian formalism to free particles 65

P
2

r
0

geodesic

P1

Figure 4.8: Geodesic curves on a rotating disc,coordinates

r = 0, r = r0 , for = 0

Inserting this into 4.58 gives:


s
p p2 2
= 1+ (4.60)
r0 c2

Rearranging 4.59,using 4.60 gives:

dr 2 rdr d
p 2
p =
2 2
r r r0 c 2 2
r r0 r0

Integrating this yields:


q
r0 2 r0
= r 2 r02 arccos
c2 r

Example 4.5.3 (Christoffel symbols in a hyperbolically accelerated reference frame)


The Christoffel symbols were defined in Equation (4.52).

1
g (g , + g , g , ).
2

In this example
 gx 2
gtt = 1 + 2 c2 , gxx = gyy = gzz = 1
c
gtt
and only the term x contributes to . Thus the only non-vanishing Christoffel
66 Chapter 4. Covariant Differentiation

ct

LOW
HIGH

g
P

non-geodetic curve between O and P


geodetic curve
2
1

O
x
Figure 4.9: Vertical throw in the accelerated referenceframe.

symbols are

 
1 tt gtt
txt = ttx = g
2 x
1 gtt
=
2gtt x

2 1 + gx c2
g
=  2
2 1 + gx c2
c2
1 g
=  2
1 + gx
c2
c
 
1 gtt
xtt = g xx
2 x
1 n  gx  g 2 o
= 2 1 + 2 c
 2 gx  c c2
= 1+ 2 g
c
4.5 Application of the Lagrangian formalism to free particles 67

Example 4.5.4 (Vertical projectile motion in a hyperbolically accelerated reference frame)

 gx 2
ds2 = 1 + 2 c2 dt2 + dx2 + dy 2 + dz 2 (4.61)
c
Vertical motion implies that dy = dz = 0 and the Lagrange function becomes
1
L= g x x
2
1 gx 2 1
= 1 + 2 c2 t2 + x2
2 c 2
where the dots imply differentiation w.r.t the particles proper time, . And the
initial conditions are:
x(0) = 0, x(0) = (u0 , ux , 0, 0)
= (c, v, 0, 0),

2 1/2
where, = 1 v 2 /c .
What is the maximum height, h reached by the particle?
v2
Newtonian description: 12 mv 2 = mgh h = 2g .
Relativistic description: t is a cyclic coordinate x 0 = ct is cyclic and p0 =
constant.

L 1 L  gx 2
p0 = = = c 1 + t (4.62)
x0 c t c2
Now the 4-velocity identity is

~u ~u = g x x = c2 (4.63)

so
1 gx 2 1 1
1 + 2 c2 t2 + x2 = c2 (4.64)
2 c 2 2
and given that the maximum height h is reached when x = 0 we get
 
gh 2 2
1+ 2 tx=h = 1. (4.65)
c
Now, since p0 is a constant of the motion, it preserves its initial value throughout
the flight (i.e. p0 = ct(0) = c) and particularly at x = h,
 
gh 2
(4.62) p0 = c = c 1 + 2 tx=h (4.66)
c
Finally, dividing equation (4.65) by equation (4.66) and substituting back in equation
(4.65) gives
c2
h= ( 1) (4.67)
g
68 Chapter 4. Covariant Differentiation

In the Newtonian limit (4.67) becomes


   
c2 1 c2 1 v2 v2
h= 1 u 1+ 1 h u
g (1 v 2 /c2 )1/2 g 2 c2 2g

Example 4.5.5 (The twin paradox)


Eva travels to Alpha Centauri, 4 light years from the Earth, with a velocity v = 0.8c
( = 1/0.6). The trip takes 5 years out and 5 years back. This means that Eli, who
remains at Earth is 10 years older when she meets Eva at the end of her journey.
Eva, on the other hand, is 10(1 v 2 /c2 )1/2 = 10(0.6) = 6 years older.

Eli Eva

X
Figure 4.10: The twins Eli and Eva each travel between two fixed events in space-
time

According to the theory of relativity, Eva can consider herself as being stationary
and Eli as the one whom undertakes the long journey. In this picture it seems that
Eva and Eli must be 10 and 6 years older respectively upon their return.
Let us accept the principle of general relativity as applied to accelerated reference
frames and review the twin paradox in this light.
Evas description of the trip when she sees herself as stationary is as follows.
Eva perceives a Lorentz contracted distance between the Earth and Alpha Cen-
tauri, namely, 4 light years 1/ = 2.4light years. The Earth and Eli travel with
years
v = 0.8c. Her travel time in one direction is then 2.4light 0.8c = 3 years. So the
round trip takes 6 years according to Eva. That is Eva is 6 years older when they
meet again. This is in accordance with the result arrived at by Eli. According to
4.5 Application of the Lagrangian formalism to free particles 69

Eva, Eli ages by only 6 years 1/ = 3.6 years during the round trip, not 10 years
as Eli found.
On turning about Eva experiences a force which reduces her velocity and acceler-
ates her towards the Earth and Eli. This means that she experiences a gravitational
force directed away from the Earth. Eli is higher up in this gravitational field and
ages faster than Eva, because of the gravitational time dilation. We assume that
Eva has constant proper acceleration and is stationary in a hyperbolically accelerated
frame as she turns about.
The canonical momentum pt for Eli is then(see Equation (4.62))
 gx 2
pt = 1 + 2 ct
c
Inserting this into the 4-velocity identity gives
 gx 2  gx 2
p2t c2 1 + 2 = 1 + 2 x2 , (4.68)
c c
or

1 + gx
c2
d = q dx
gx 2
p2t c2 1 + c2

Now, since x = 0 for x = x2 (x2 is Elis turning point according to Eva), we


have that
 gx2 
pt = c 1 + 2
c
Let x1 be Elis position according to Eva just as Eva begins to notice the gravitational
field. That is when Eli begins to slow down in Evas frame.
Integration from x1 to x2 and inserting the value of pt gives
r
c gx2 2  gx1 2
12 = 1+ 2 1+ 2
g c c
q
1
lim 12 = x22 x21 .
g c

Now setting x2 = 4 and x1 = 2.4 light years respectively we get

lim 12 = 3.2 years


g

Elis aging as she turns about is, according to Eva,

Eli = 2 lim 12 = 6.4 years.


g

So Elis has aged by a total of Eli = 3.6 + 6.4 = 10 years, according to Eva, which
is just what Eli herself found.
70 Chapter 4. Covariant Differentiation

4.5.3 Gravitational Doppler effect


This concerns the frequency shift of light traversing up or down in a gravitational
field. The 4-momentum of a particle with relativistic energy E and spatial
velocity w
~ (3-velocity) is given by:

P~ = E(1, w)
~ (c = 1) (4.69)
~ be the 4-velocity of an observer. In a co-moving orthonormal basis of the
Let U
observer we have U~ = (1, 0, 0, 0). This gives

~ P~ = E
U (4.70)

The energy of a particle with 4-momentum P~ measured by an observer with


4-velocity U~ is
E = U ~ P~ (4.71)
Let ES = (U ~ P~ )S and Ea = (U ~ P~ )a be the energy of a photon, measured
locally by observers in rest in the transmitter and receiver positions, respectively.
This gives1
ES Ea
= (4.72)
~ ~
(U P ) S (U P~ )a
~
Let the angular frequency of the light, measured by the transmitter and receiver,
be ws and wa , respectively. We then have
ES Ea
ws = , wa = , (4.73)
h h
which gives:
~ P~ )a
(U
wa = w (4.74)
~ P~ )s s
(U
For an observer in rest in a time-independent orthogonal metric we have

~ P~ = U t Pt = dt Pt
U (4.75)
d
where Pt is a constant of motion (since t is a cyclic coordinate) for photons and
hence has the same value in transmitter and receiver positions.

ds2 = gtt dt2 + gii (dxi )2 ds2 = d 2



d 2 = gtt dt2 d = gtt dt
dt 1
= , (4.76)
d gtt
which gives
~ P~ = 1 Pt .
U (4.77)
gtt
1~ ~ = A0 B 0 + A1 B 1 + ... = g00 A0 B 0 + g11 A1 B 1 + ..., an orthonormal basis gives
AB
~ ~
A B = A0 B 0 + A1 B 1 + ...
4.6 The Koszul connection 71

Inserting this into the expression for angular frequency (4.74) gives
s
(gtt )s
wa = ws
(gtt )a

Note: we have assumed an orthogonal and time-independent metric, i.e. P t1 =


Pt2 . Inserting the metric of a hyperbolically accelerated reference system with
gx 2
gtt = (1 + ) (4.78)
c2
gives
gxs
1+ c2
wa = gxa ws , (4.79)
1+ c2
or gxs g
wa w s 1+ c2 c2
(xs xa ) g
= gxa 1 = gxa H, (4.80)
ws 1+ c2 1+ c2 c2
where H = xs xa is the difference in height between transmitter and receiver.

Example 4.5.6 (Measurements of gravitational Doppler effects (Pound and Rebka 1960))

H 20m, g = 10m/s2
gives
w 200
= = 2.2 1015 .
w 9 1016
This effect was measured by Pound and Rebka in 1960.

4.6 The Koszul connection


The covariant directional derivative of a scalar field f in the direction of a vector
~u is defined as:
u~ f ~u(f ) (4.81)
Here the vector ~u should be taken as a differensial operator. (In coordinate
basis, ~u = u x )
The directional derivative along a basis vector ~e is written as:

~e (4.82)

Hence ( ) = ~e ( ) = ~e ( )
72 Chapter 4. Covariant Differentiation

Definition 4.6.1 (Koszuls connection coeffecients in an arbitrary basis)


In an aribitrary basis the Koszul connection coefficients are defined by

~e ~e (4.83)

which may also be written ~e (~e ) = ~e . In coordinate basis , is reduced


to Christoffel symbols and one often writes ~e , = ~e . In an arbitrary basis
, has no symmetry.

Example 4.6.1 (The connection coefficients in a rotating reference frame.)


Coordinate transformation: (T, R, are coordinates in the non-rotating reference
frame, t, r, in the rotating.) Corresponding Cartesian coordinates:X, Y and x, y.

t = T, r = R, = T
X = R cos , Y = R sin
X = r cos( + t), Y = r sin( + t)

P (x,y) x
(X,Y)

+t t
X

Figure 4.11: The non-rotating coordinate system (X,Y) and the rotating system
(x,y),rotating with angular velocity

X Y T
~et = = + +
t t X t Y t T
4.6 The Koszul connection 73

gives:
~et = r sin( + t)~eX + r cos( + t)~eY + ~eT
X Y
~er = +
r X r Y
= cos( + t)~eX + sin( + t)~eY
X Y
~e = +
X Y
= r sin( + t)~eX + r cos( + t)~eY
We are going to find the Christoffel symbols, which involves differentiation of ba-
sis vectors. This coordinate transformation makes this easy, since ~e X , ~eY , ~eT are
constant. Differentiation gives
t~et = r 2 cos( + t)~eX r 2 sin( + t)~eY (4.84)
The connection coefficients are (see eq. 4.83)
~e ~e (4.85)
So, to calculate , the right hand side of eq.4.84 has to by expressed by the
basis that we are differentiating.
By inspection, the right hand side is r 2~er .
That is t~et = r 2~er giving rtt = r 2 .
The other nonzero Christoffel symbols are
1
rt = tr = , r = r =
r r
rt = rt = r, r = r

Example 4.6.2 (Acceleration in a non-rotating reference frame (Newton))

~r = ~v = (v i + ijk v j v k )~ei ,
d
where dt . i, j, and k are space indices. Inserting the Christoffel symbols for
plane polar coordinates (see example 4.2.1), gives:
2
~ainert = (r r 2 )~er + ( + r )~e
r

Example 4.6.3 (The acceleration of a particle, relative to the rotating reference frame)
Inserting the Christoffel symbols from example 4.6.1:
2
~arot = (r r 2 r 2 + rt t + rt t)~er + ( + r + rt r t + tr tr)~e
r
= (r r 2 r 2 2r )~er + (r + 2r + 2r)~e
= ~ainert (r 2 2r )~er + 2r~e
74 Chapter 4. Covariant Differentiation

The angular velocity of the reference frame, is


~ = ~e z . We also introduce ~r = r~er .
The velocity relative to the rotating reference frame is then:

~r = r~er + r~e r

Furthermore
d~er ~er dxi
~e r = = = v i~er,i
dt xi dt
Using definition 4.6.1 in a coordinate basis, this may be written

~e r = v i jri~ej

Using the expressions of the Christoffel symbols in example 4.6.1, we get


1
~e r = v r~e = ~e = ~e
r
Hence
~v = ~r = r~er + r ~e
Inserting this into the expression for the acceleration, gives:

~rrot = ~rinert +
~ (~ ~r) + 2~ ~v

We can see that the centrifugal acceleration (the term in the middle) and the coriolis
acceleration (last term) is contained in the expression for the covariant derivative.

4.7 Connection coefficients and structure coeffi-


cients c in a Riemannian (torsion free) space
The commutator of two vectors, ~u and ~v , expressed by covariant directional
derivatives is given by:
[~u, ~v ] = u~ ~v ~v ~u (4.86)
Let ~u = e~ and ~v = e~ . We then have:

[e~ , e~ ] = e~ e~ . (4.87)

Using the definitions of the connection and structure coefficients we get:

c e~ = ( )e~ (4.88)

Thus in a torsion free space

c = (4.89)

In coordinate basis we have



e~ = , e~ = (4.90)
x x
4.8 Covariant differentiation of vectors, forms and tensors 75

And therefore:

[e~ , e~ ] = [

, ]
x x

=
( ) ( ) (4.91)
x x x x
2 2
= =0
x x x x
Equation (4.91) shows that c = 0, and that the connection coefficients in
Equation (4.89) therefore are symmetrical in a coordinate basis:
= (4.92)

4.8 Covariant differentiation of vectors, forms and ten-


sors
4.8.1 Covariant differentiation of a vector in an arbitrary basis
~ = (A e~ )
A
(4.93)
= A e~ + A e~

A = e~ (A ) , e~ = M , (4.94)
x
where M are the elements of a transformation matrix between a coordinate
basis { x } and an arbitrary basis {e~ }. ( If e~ had been a coordinate basis
vector, we would have gotten e~ (A ) = x (A ) = A , ).
~ = [e~ (A ) + A ]e~
A (4.95)

Definition 4.8.1 (Covariant derivative of a vector)


The covariant derivative of a vector in an arbitrary basis is defined by:
~ A ; ~e
A (4.96)
So:
A; = ~e (A ) + A
(4.97)
where ~e ~e

4.8.2 Covariant differentiation of forms


Definition 4.8.2 (Covariant directional derivative of a one-form field)
~ and a one-form field , the covariant directional derivative of
Given a vector field A
in the direction of the vector ~u is defined by:
(u~ )(A~ ) u~ [(A~ )] (u~ A)
~ (4.98)
| {z }
A
76 Chapter 4. Covariant Differentiation

Let = (basis form), (e~ ) and let A ~ = e~ and ~u = e~ . We then


have:
( )(e~ ) = [ (e~ )] ( e~ ) (4.99)
| {z }

The covariant directional derivative of a constant scalar field is zero, =
0. We therefore get:
( )(e~ ) = ( e~ )
= ( e~ )
= (e~ ) (4.100)
=
=
The contraction between a one-form and a basis vector gives the components
of the one-form, (e~ ) = . Equation (4.100) tells us that the -component of
is equal to , and that we therefore have

= (4.101)

Equation (4.101) gives the directional derivatives of the basis forms. Using the
product of differentiation gives
= ( )
= ( ) + (4.102)

= e~ ( )

Definition 4.8.3 (Covariant derivative of a one-form)


The covariant derivative of a one-form = is therefore given by Equation
(4.103) below, when we let in the first term on the right hand side in (4.102):

= [e~ ( ) ] (4.103)

The covariant derivative of the one-form components are denoted by ; and


are defined by
; (4.104)
It then follows that
; = e~ ( ) (4.105)

It is worth to note that in Equation (4.105) are not Christoffel symbols. In


coordinate basis we get:
; = , (4.106)
where = are Christoffel symbols.
4.9 The Cartan connection 77

4.8.3 Generalization for tensors of higher rank


Definition 4.8.4 (Covariant derivative of a tensor)
Let A and B be two tensors of arbitrary rank. The covariant directional derivative
along a basis vector e~ of a tensor A B of arbitrary rank is defined by:

(A B) ( A) B + A ( B) (4.107)

We will use (4.107) to find the formula for the covariant derivative of the com-
ponents of a tensor of rank 2:

S = (S )
= ( S ) + S ( ) + S ( ) (4.108)
= (S, S S )

where S, = ~e (S ). Defining the covariant derivative S ; by

S = S; (4.109)

we get

S; = S, S S (4.110)

For the metric tensor we get

g; = g, g g (4.111)

From

g = e~ e~ (4.112)

we get:

g, = ( e~ ) e~ + e~ ( e~ )
= e~ e~ + e~ e~ (4.113)
= g + g

This means that


g; = 0 (4.114)
So the metric tensor is a (covariant) constant tensor.

4.9 The Cartan connection


Definition 4.9.1 (Exterior derivative of a basis vector)

d~e ~e (4.115)
78 Chapter 4. Covariant Differentiation

Exterior derivative of a vector field:

~ = d(~e A ) = ~e dA + A d~e
dA (4.116)

In arbitrary basis:
dA = ~e (A ) (4.117)

(In coordinate basis, ~e (A ) = x
(A ) = A, )
giving:

~ = ~e [~e (A ) ] + A ~e
dA
(4.118)
= (~e (A ) + A )~e

~ = A ~e
dA (4.119)
;

Definition 4.9.2 (Connection forms )


The connection forms are 1-forms, defined by:

d~e ~e
(4.120)
~e = ~e = ~e

= (4.121)

The exterior derivatives of the components of the metric tensor:

dg = d(~e ~e ) = ~e d~e + ~e d~e (4.122)


where the meaning of the dot is defined as follows:

Definition 4.9.3 (Scalar product between vector and 1-form)


The scalar product between a vector ~u and a (vectorial) one form A = A ~e
is defined by:
~u A u A (~e ~e ) (4.123)

Using this definition, we get:

dg = (~e ~e ) + (~e ~e )
(4.124)
= g + g
4.9 The Cartan connection 79

Lowering an index gives

dg = + (4.125)

In an orthonormal basis field there is Minkowski-metric:

g = (4.126)

which is constant. Then we have :

dg = 0 = (4.127)

where we write = . It follows that = .


It also follows that

tij = tij = itj = itj


(4.128)
ij k = j
ik

Cartans 1st structure equation (without proof):


1
d = c
2
1
= ( ) (4.129)
2
=
=

d = and d = (4.130)

In coordinate basis, we have = dx .


Thus, d = d2 x = 0.
We also have c = 0 , and C1 is reduced to an identity.This formalism can
not be used in coordinate basis!

Example 4.9.1 (Cartan-connection in an orthonormal basis field in plane polar coord.)

ds2 = dr 2 + r 2 d 2
Introducing basis forms in an orthonormal basis field (where the metric is g rr =
g = 1):
ds2 = grr r r + g = r r +

r = dr, = rd
Exterior differentiation gives:
1
d r = d2 r = 0, d = dr d = r
r
80 Chapter 4. Covariant Differentiation

C1:

d =
= r r

We have that d r = 0 , which gives:

r = r (4.131)

since = 0. (rr =0 because of the antisymmetry = .)


We also have: d = 1r r . C1:

d = r r
|{z}
=0

r = r + rr r (4.132)

giving r = 1r .
We have: r = r . Using equations 4.131 and 4.132 we get:

rr = 0
1
r =
r
giving r = r = 1r .
Chapter 5

Curvature

5.1 The Riemann curvature tensor

~ + ) 
A(

~
u~ A1

M ~ QP ( + )
A
~
A() ~u
* Q
+

~
Figure 5.1: Parallel transport of A

~ along a vector ~u was


The covariant directional derivative of a vector field A
defined and interpreted geometrically in section 4.2, as follows

~
~ = dA = A; u ~e
u~ A
d (5.1)
~ QP ( + ) A()
A ~
= lim
0

~ QP be the parallel transported of A


Let A ~ from Q to P. Then to first order in
~ ~ ~
we have: AQP = AP + (u~ A)P and

81
82 Chapter 5. Curvature




3
~k
A
6

:
-
~
A

Figure 5.2: Parallel transport of a vector along a triangle of angles 90 is rotated


90

~P Q = A
A ~ Q (u~ A)
~ Q (5.2)

To second order in we have:

~ P Q = (1 u~ + 1 u~ u~ ()2 )A
A ~Q (5.3)
2
~ P Q is parallel transported further on to R we get
If A

~ P QR = (1 u~ + 1 u~ u~ ()2 )
A
2 (5.4)
1 ~R
(1 ~v + ~v ~v ()2 )A
2
~ around the whole polygon we get:
If we parallel transport A

~ P QRST P = (1 + u~ + 1 u~ u~ ()2 )
A
2
1
(1 + ~v + ~v ~v ()2 )
2 (5.5)
1
(1 [~u,~v ] ()2 ) (1 u~ + u~ u~ ()2 )
2
1 ~P
(1 ~v + ~v ~v ()2 )A
2
Calculating to 2. order in gives:
~ P QRST P = A
A ~ P + ([u~ , ~v ] [~u,~v ] )()2 A
~P (5.6)
5.1 The Riemann curvature tensor 83

1 S
6
]
(u~ ~v ~v ~u)
= [~u, ~v ]2
~v (T )
u~ ~v

T -
 ~vP T 

~u(P ) ~uP Q ~v ~u
j
3R
~
~v A 1A~
O Q
~I
A P
~ PI
A Q
~u(Q)

-
P ~v (P ) Q
+

Figure 5.3: Geometrically implied curvature from non-zero differences between


vectors along a curve (parameterized by ) and their parallel transported equiv-
alents

There is a variation of the vector under parallel transport around the closed
polygon:
~=A
A ~ P QRST P A
~ P = ([u~ , ~v ] [~u,~v] )A
~ P ()2 (5.7)

We now introduce the Riemanns curvature tensor as:


~ ~u, ~v ) ([u~ , ~v ] [~u,~v] )(A)
R( , A, ~ (5.8)

The components of the Riemann curvature tensor is defined by applying the


tensor on basis vectors,

R ~e ([ , ] [~e ,~e ] )(~e ) (5.9)

Anti-symmetry follows from the definition:

R = R (5.10)

~ under parallel transport around the poly-


The expression for the variation of A
gon, Eq. (5.7), can now be written as:
~ = R(
A ~ ~u, ~v )()2
, A,
= R( , A ~e , u~e , v ~e )()2
= ~e R A u v ()2 (5.11)
1
= ~e R A (u v u v )()2
2
84 Chapter 5. Curvature

~ = ~u ~v ()2
S
6
~u

S

-
~v

The area of the parallellogram defined by the vectors ~u and ~v is

~ = ~n ~v ()2 .
S

Using that
(~u ~v ) = u v u v .
we can write Eq. (5.11) as:

~= 1
A A R S ~e . (5.12)
2

The components of the Riemann tensor expressed by the connection- and structure-
coefficients are given below:

~e R = [ , ]~e [~e ,~e ]~e


= ( c )~e
= ~e ~e c ~e
(Kozul-connection) = ~e ~e c ~e
(5.13)
= ( )~e + ~e
( )~e ~e c ~e
= ~e ( )~e + ~e
~e ( )~e ~e c ~e .

This gives (in arbitrary basis):

R = ~e ( ) ~e ( )
(5.14)
+ c .

In coordinate basis eq. (5.14) is reduced to:

R = , , + , (5.15)

where = are the Christoffel symbols.


5.1 The Riemann curvature tensor 85

Due to the antisymmetry (5.10) we can define a matrix of curvature-forms

1
R = R (5.16)
2

Inserting the components of the Riemann tensor from eq. (5.14) gives

1
R = (~e ( ) + c ) (5.17)
2

The connection forms:

= (5.18)

Exterior derivatives of basis forms:

1
d = c (5.19)
2

Exterior derivatives of connection forms (C1: d = ) :

d = d + d
1 (5.20)
= ~e ( c
2

The curvature forms can now be written as:

R = d + (5.21)

This is Cartans 2nd structure equation.


86 Chapter 5. Curvature

5.2 Differential geometry of surfaces

Figure 5.4: The geometry of a surface. We see the normal vector and the unit
vectors of the tangent plane of a point on the surface.

Imagine an arbitrary surface embedded in an Euclidian 3 dimensional space.


(See figure 5.4). Coordinate vectors on the surface :


~eu = , ~ev = (5.22)
u v
where u and v are coordinates on the surface.
Line element on the surface:

ds2 = g dx dx (5.23)

with x1 = u and x2 = v.
(1st fundamental form)
The directional derivatives of the basis vectors are written

~ , = 1, 2
~e, = ~e + K N (5.24)

~ is a unit vector orthogonal


Greek indices run through the surface coordinates, N
to the surface.
5.2 Differential geometry of surfaces 87

The equation above is called Gauss equation. We have: K = ~e, N ~ . In


2 2
coordinate basis, we have ~e, = xx = xx = ~e, . It follows that

K = K (5.25)

Let ~u be the unit tangent vector to a curve on the surface, parametrised by .


Differentiating ~u along the curve:

d~u ~
= u; u ~e + K u u N (5.26)
d | {z }
2nd fundamental form

We define g and N by:


d~u ~
= g ~e + N N (5.27)
d
g is called geodesic curvature. N is called normal curvature (external curva-
ture). g = 0 for geodesic curves on the surface.

g ~e = u; u ~e = u~ ~u
N = K u u (5.28)
d~u ~
And :N = N
d
~ = 0 along the whole curve. Differentiation gives:
We also have that ~u N

d~u ~ ~
dN
N + ~u =0 (5.29)
d d
gives:
~
dN
N = ~u (5.30)
d
which is called Weingartens equation.
g and N together give a complete description of the geometry of a surface
in a flat 3 dimensional space. We are now going to consider geodesic curves
through a point on the surface. Tangent vector ~u = u ~e with ~u ~u = g u u =
1. Directions with maximum and minimum values for the normal curvatures are
found, by extremalizing N under the condition g u u = 1. We then solve the
variation problem F = 0 for arbitrary u , where F = K u u k(g u u 1).
Here k is the Lagrange multiplicator. Variation with respect to u gives:

F = 2(K kg )u u
F = 0 for arbitrary u demands:

(K kg )u = 0 (5.31)
For this system of equations to have nonzero solutions, we must have:

det(K kg ) = 0 (5.32)
88 Chapter 5. Curvature


K11 kg11 K12 kg12
(5.33)
K21 kg21 K22 kg22 = 0

This gives the following quadratic equation for k:

k 2 det(g ) (g11 K22 2g12 K12 + g22 K11 )k + det(K ) = 0 (5.34)


(K symmetric K12 = K21 )

The equation has two solutions, k1 and k2 . These are the extremal values of k.
To find the meaning of k, we multiply eq.5.31 by u :

0 = (K kg )u u
= K u u kg u u (5.35)
= N k k = N

The extremal values of N are called the principal curvatures of the surface.
Let the directions of the geodesics with extreme normal curvature be given by
the tangent vectors ~u and ~v .Eq.5.31 gives:

K u = kg u (5.36)

We then get:

K u v = k1 g u v
= k1 u v = k1 (~u ~v )
K v u = k2 g v u = k2 (~u ~v )

gives (k1 k2 )(~u ~v ) = K (u v v u )


(5.37)
= 2K u[ v ]

K is symmetric in and . So we get:(k1 k2 )(~u ~v ) = 0. For k1 6= k2 we


have to demand ~u ~v = 0. So the geodesics with extremal normal curvature, are
orthogonal to each other.
The Gaussian curvature (at a point) is defined as:

K = N 1 N 2 (5.38)

Since N 1 and N 2 are solutions of the quadratic equation above, we get:

det(K )
K= (5.39)
det(g )
5.2 Differential geometry of surfaces 89

5.2.1 Surface curvature, using the Cartan formalism


In each point on the surface we have an orthonormal set of basis vectors. Greek
indices run through the surface coordinates (two dimensional) and Latin indices
through the space coordinates (three dimensional):
~ ) , e~ = {e~1 , e~2 }
e~a = (e~1 , e~2 , N (5.40)
Using the exterior derivative and form formalism, we find how the unit vectors
on the surface change:
de~ = e~a a
(5.41)
~ 3 ,
= e~ + N
where = are the connection forms on the surface, i.e. the intrinsic
connection forms. The extrinsic connection forms are
3 = K , 3 = K (5.42)
We let the surface be embedded in an Euclidean (flat) 3-dimensional space. This
means that the curvature forms of the 3-dimensional space are zero:
Ra3b = 0 = d ab + ak kb (5.43)
which gives:
R3 = 0 = d + + 3 3
(5.44)
= R + 3 3 ,
where R are the curvature forms of the surface. We then have:
1
R = 3 3 (5.45)
2
Inserting the components of the extrinsic connection forms, we get: (when we
remember the anti symmetry of and in R )
R = K K K K (5.46)
We now lower the first index:
R = K K K K (5.47)
R are the components of a curvature tensor which only refer to the dimen-
sions of the surface. In particular we have:
R1212 = K11 K22 K12 K21 = det K (5.48)
We then have the following connection between this component of the Riemann
curvature tensor of the surface and the Gaussian curvature of the surface:
det K R1212
K = N 1 N 2 = = (5.49)
det g det g
Since the right hand side refers to the intrinsic curvature and the metric on the
surface, we have proved that the Gaussian curvature of a surface is an intrinsic
quantity. It can be measured by observers on the surface without embedding the
surface in a three-dimensional space. This is the contents of Gauss theorema
egregium.
90 Chapter 5. Curvature

5.3 The Ricci identity

e~ R A = ( [e~ ,e~ ] )(A)


~ (5.50)

In coordinate basis this is reduced to

e~ R A = (A; A; )e~ , (5.51)

where

A; (A; ); (5.52)

The Ricci identity on component form is:

A R = A; A; (5.53)

We can write this as:

~ = 1 R A e~
d2 A (5.54)
2
This shows us that the 2nd exterior derivative of a vector is equal to zero only
in a flat space. Equations (5.53) and (5.54) both represents the Ricci identity.

5.4 Bianchis 1st identity


Cartans 1st structure equation:

d = (5.55)

Cartans 2nd structure equation:

R = d + (5.56)

Exterior differentiation of (5.55) and use of P oincare 0 s lemma (4.15) gives:


(d2 = 0)

0 = d d (5.57)

Use of (5.55) gives:

d +
= 0 (5.58)

From this we see that

(d + ) = 0 (5.59)

We now get Bianchis 1st identity:

R = 0 (5.60)
5.5 Bianchis 2nd identity 91

On component form Bianchis 1st identity is


1
R = 0 (5.61)
2
| {z }
R

The component equation is: (remember the anti symmetry in and )

R[] = 0 (5.62)

or

R + R + R + = 0 (5.63)

where the anti symmetry R = R has been used. Without this anti
symmetry we would have gotten six, and not three, terms in this equation.

5.5 Bianchis 2nd identity


Exterior differentiation of (5.56)

d R = R R +
(5.64)
= R R

We now have Bianchis 2nd identity as a form equation:

d R + R R = 0 (5.65)

As a component equation Bianchis 2nd identity is given by

R[;] = 0 (5.66)

Definition 5.5.1 (Contraction)


Contraction is a tensor operation defined by

R R (5.67)

We must here have summation over . What we do, then, is constructing a new
tensor from another given tensor, with a rank 2 lower than the given one.

The tensor with components R is called the Ricci curvature tensor.


Another contraction gives the Ricci curvature scalar, R = R .
92 Chapter 5. Curvature

Riemann curvature tensor has four symmetries. The definition of the Rie-
mann tensor implies that R = R
Bianchis 1st identity: R [] = 0
From Cartans 2nd structure equation follows

R = d +
(5.68)
R = R

By choosing a locally Cartesian coordinate system in an inertial frame we get


the following expression for the components of the Riemann curvature tensor:
1
R = (g, g, + g, g, ) (5.69)
2
from which it follows that R = R . Contraction of and leads to:

R = R
(5.70)
R = R

i.e. the Ricci tensor is symmetric. In 4-D the Ricci tensor has 10 independent
components.
Chapter 6

Einsteins Field Equations

6.1 Energy-momentum conservation


6.1.1 Newtonian fluid
Energy-momentum conservation for a Newtonian fluid in terms of the divergence
of the energy momentum tensor can be shown as follows. The total derivative
of a velocity field is
D~v ~v ~ v
+ (~v )~ (6.1)
Dt t
~v
t is the local derivative which gives the change in ~v as a function of time
~ v is called the convective derivative of ~v . It
at a given point in space. (~v )~
represents the change of ~v for a moving fluid particle due to the inhomogeneity
of the fluid velocity field. In component notation the above become

Dv i v i v i
+ vj j (6.2)
Dt t x
The continuity equation

(v i )
+ (~v ) = 0 or + =0 (6.3)
t t xi
Eulers equation of motion (ignoring gravity)
 i i

D~v ~ v j v p
= p or +v j
= i (6.4)
Dt t x x
The energy momentum tensor is a symmetric tensor of rank 2 that
describes material characteristics.
00
T T 01 T 02 T 03
T 10 T 11 T 12 T 13
T = T 20 T 21
(6.5)
T 22 T 23
T 30 T 31 T 32 T 33
c1

93
94 Chapter 6. Einsteins Field Equations

T 00 represents energy density.

T i0 represents momentum density.

T ii represents pressure (T ii > 0).

T ii represents stress (T ii < 0).

T ij represents shear forces (i 6= j).

Example 6.1.1 (Energy momentum tensor for a Newtonian fluid)

T 00 = T i0 = v i
(6.6)
T ij = v i v j + p ij

where p is pressure, assumed isotropic here. We choose a locally Cartesian coordinate


system in an inertial frame such that the covariant derivatives are reduced to partial
derivatives. The divergence of the momentum energy tensor, T ; has 4 components,
one for each value of .
The zeroth component is

T 0 0 00 0i
; = T , = T ,0 + T ,i
(v i ) (6.7)
= +
t xi
which by comparison to Newtonian hydrodynamics implies that T 0 ; = 0 is the
continuity equation. This equation represents the conservation of energy.
The ith component of the divergence is
ij
T i i0
, = T ,0 + T ,j
(v i ) (v i v j + p ij )
= + (6.8)
t xj
v i v j i
j v p
= + vi + vi + v + i
t t xj xj x
now, according to the continuity equation

(v i )
=
xi t
v i v i p
T i , = + vi vi + v j j + i
t t t x x (6.9)
Dv i p
= + i
Dt x
Dv i p
T i ; = 0 = i
Dt x
which is Eulers equation of motion. It expresses the conservation of momentum.
6.2 Einsteins curvature tensor 95

The equations T
; = 0 are general expressions for energy and momentum
conservation.

6.1.2 Perfect fluids


A perfect fluid is a fluid with no viscosity and is given by the energy-momentum
tensor
T = (c2 + p)u u + pg (6.10)
where and p are the mass density and the stress, respectively, measured in the
fluids rest frame, u are the components of the 4-velocity of the fluid.
In a comoving orthonormal basis the components of the 4-velocity are u =
(c, 0, 0, 0). Then the energy-momentum tensor is given by
2
c 0 0 0
0 p 0 0
T =
0 0 p 0
(6.11)
0 0 0 p
where p > 0 is pressure and p < 0 is tension.
There are three different types of perfect fluids that are useful.
1. dust or non-relativistic gas is given by p = 0 and the energy-momemtum
tensor T = c2 u u .
2. radiation or ultra-relativistic gas is given by a traceless energy-momemtum
tensor, i.e. T = 0. It follows that p = 31 c2 .
3. vacuum energy: If we assume that no velocity can be measured relatively
to vacuum, then all the components of the energy-momentum tensor must
be Lorentz-invariant. It follows that T g . If vacuum is defined as a
perfect fluid we get p = c2 so that T = pg = c2 g .

6.2 Einsteins curvature tensor


The field equations are assumed to have the form:

space-time curvature momentum-energy tensor


Also, it is demanded that energy and momentum conservation should follow as
a consequence of the field equation. This puts the following constraints on the
curvature tensor: It must be a symmetric, divergence free tensor of rank 2.
Bianchis 2nd identity:
R; + R; + R; = 0 (6.12)
contraction of and
R; R; + R; = 0
(6.13)
R ; R ; + R; = 0
96 Chapter 6. Einsteins Field Equations

further contraction of and gives

R; R; + R
: = 0
R; R ; + R; = 0 (6.14)
2R; = R ;

Thus, we have calculated the divergence of the Ricci tensor,

1
R; = R (6.15)
2 ;
Now we use this expression together with the fact that the metric tensor is co-
variant and divergence free to construct a new divergence free curvature tensor.

1
R; R ; = 0 (6.16)
2

Keeping in mind that (g R); = g R; we multiply (6.16) by g to get

1
g R; g R ; = 0
2
  (6.17)
1 
g R g R =0
; 2 ;

interchanging and in the first term of the last equation:


  1 
g R g R =0
; 2 ;
  (6.18)
1
R R =0
2 ;

since g R = R =R . So that R 12 R is the divergence free curvature


tensor desired.
This tensor is called the Einstein tensor and its covariant components are
denoted by E . That is

1
E = R g R (6.19)
2

NOTE THAT: E ; = 0 4 equations, giving only 6 equations from E ,


which secures a free choice of coordinate system.

6.3 Einsteins field equations


Einsteins field equations:

E = T (6.20)
6.3 Einsteins field equations 97

or
1
R g R = T (6.21)
2
Contraction gives:
1
R 4R = T , where T T
2 (6.22)
R = T

1
R = g (T ) + T , (6.23)
2
Thus the field equations may be written in the form
1
R = (T g T ) (6.24)
2
In the Newtonian limit the metric may be written
 
2
ds = 1 + 2 dt2 + (1 + hii )(dx2 + dy 2 + dz 2 )
2
(6.25)
c

where the Newtonian potential ||  c 2 . We also have T00  Tkk and T T00 .
Then the 00-component of the field equations becomes

R00 T00 (6.26)
2
Furthermore we have

R00 = R00 = Ri 0i0


= i 00,i i 0i,0
k00 1
= k
= 2 2 (6.27)
x c

Since T00 c2 eq.(6.26) can be written 2 = 21 c4 . Comparing this equation


with the Newtonian law of gravitation on local form: 2 = 4G, we see that
= 8G
c4
.
In classical vacuum we have : T = 0, which gives

E = 0 or R = 0 . (6.28)

These are the vacuum field equations. Note that R = 0 does not imply
R = 0.
98 Chapter 6. Einsteins Field Equations

Digression 6.3.1 (Lagrange (variation principle))


It was shown by Hilbert that the field equations may be deduced from a variation
principle with action Z

R gd4 x , (6.29)

where R g is the Lagrange density. One may also include a so-called cosmological
constant : Z

(R + 2) gd4 x (6.30)

The field equations with cosmological constant are


1
R g R + g = T (6.31)
2

6.4 The geodesic postulate as a consequence of the


field equations
The principle that free particles follow geodesic curves has been called the
geodesic postulate. We shall now show that the geodesic postulate follows
as a consequence of the field equations.
Consider a system of free particles in curved space-time. This system can
be regarded as a pressure-free gas. Such a gas is called dust. It is described by
an energy-momentum tensor

T = u u (6.32)

where is the rest density of the dust as measured by an observer at rest in the
dust and u are the components of the four-velocity of the dust particles.
Einsteins field equations as applied to space-time filled with dust, take the
form
1
R g R = u u (6.33)
2
Because the divergence of the left hand side is zero, the divergence of the right
hand side must be zero, too

(u u ); = 0 (6.34)

or

(u u ); = 0 (6.35)

we now regard the quantity in the parenthesis as a product of u and u . By


the rule for differentiating a product we get

(u ); u + u u; = 0 (6.36)
6.4 The geodesic postulate as a consequence of the field equations 99

Since the four-velocity of any object has a magnitude equal to the velocity of
light we have

u u = c2 (6.37)

Differentiation gives

(u u ); = 0 (6.38)

Using, again, the rule for differentiating a product, we get

u; u + u u; = 0 (6.39)

From the rule for raising an index and the freedom of changing a summation
index from to , say, we get

u ; u = u u; = g u u; = u g u; = u u; = u u; (6.40)

Thus the two terms of eq.(6.39) are equal. It follows that each of them are equal
to zero. So we have

u u; = 0 (6.41)

Multiplying eq.(6.36) by u, we get

(u ); u u + u u u; = 0 (6.42)

Using eq.(6.37) in the first term, and eq.(6.41) in the last term, which then
vanishes, we get

(u ); = 0 (6.43)

Thus the first term in eq.(6.36) vanishes and we get

u u; = 0 (6.44)

Since 6= 0 we must have

u u; = 0 (6.45)

This is just the geodesic equation. Conclusion: It follows from Einsteins


field equations that free particles move along paths corresponding to geodesic
curves of space-time.
Chapter 7

The Schwarzschild spacetime

7.1 Schwarzschilds exterior solution


This is a solution of the vacuum field equations E = 0 for a static spherically
symmetric spacetime. One can then choose the following form of the line element
(employing units so that c=1),

ds2 = e2(r) dt2 + e2(r) dr 2 + r 2 d2


(7.1)
d2 = d 2 + sin2 d2

These coordinates are chosen so that the area of a sphere with radius r is 4r 2 .
Physical distance in radial direction, corresponding to a coordinate distance

dr, is dlr = grr dr = e(r) dr.
Here follows a stepwise algorithm to determine the components of the Ein-
stein tensor by using the Cartan formalism:

1. Using orthonormal basis (ie. solving E = 0) we find

t = e(r) dt , r = e(r) dr , = rd , = r sin d (7.2)

2. Computing the connection forms by applying Cartans 1. structure equa-


tions
d = (7.3)

d t = e 0 dr dt
= e 0 e r e t
(7.4)
= e 0 t r
= tr r

tr = e 0 t + f1 r (7.5)

100
7.1 Schwarzschilds exterior solution 101

3. To determine the f-functions we apply the anti-symmetry

= (7.6)

This gives:

1
r = r = e
r
1
= = cot
r (7.7)
t r 0 t
r = + t = e
1
r = r = e
r

4. We then proceed to determine the curvature forms by applying Cartans


2nd structure equations

R = d + (7.8)

which gives:
2
Rtr = e2 (00 + 0 0 0 ) t r
1
Rt = e2 0 t
r
1
Rt = e2 0 t
r
1 2 0 r (7.9)
r
R = e
r
1
Rr = e2 0 r
r
1
R = 2 (1 e2 )

r

5. By applying the following relation

1
R = R (7.10)
2
we find the components of Riemanns curvature tensor.

6. Contraction gives the components of Riccis curvature tensor

R R (7.11)

7. A new contraction gives Riccis curvature scalar

R R (7.12)
102 Chapter 7. The Schwarzschild spacetime

8. The components of the Einstein tensor can then be found

1
E = R R , (7.13)
2
where = diag(1, 1, 1, 1). We then have:

2 2 0 1
Ett = e + (1 e2 )
r r2
2 1
Err = e2 0 (1 e2 ) (7.14)
r r2
2 0 0
E = E = e2 (00 + 0 0 0 + )
r r
We want to solve the equations E = 0. We get only 2 independent
equations, and choose to solve those:

Ett = 0 and Err = 0 (7.15)

By adding the 2 equations we get:

Ett + Err = 0

2 2 0
e ( + 0 ) = 0
r
( + )0 = 0 + = K1 (const) (7.16)
We now have:
ds2 = e2 dt2 + e2 dr 2 + r 2 d2 (7.17)
By choosing a suitable coordinate time, we can achieve

K1 = 0 =

Since we have ds2 = e2 dt2 + e2 dr 2 + r 2 d2 , this means that grr =


g1tt . We still have to solve one more equation to get the complete solution,
and choose the equation Ett = 0, which gives

2 2 0 1
e + 2 (1 e2 ) = 0
r r
This equation can be written:

1 d
[r(1 e2 )] = 0
r 2 dr (7.18)
r(1 e2 ) = K2 (const)

If we choose K2 = 0 we get = 0 giving = 0 and

ds2 = dt2 + dr 2 + r 2 d2 , (7.19)


7.1 Schwarzschilds exterior solution 103

which is the Minkowski space-time described in spherical coordinates. In


general, K2 6= 0 and 1 e2 = Kr2 Kr , giving

K
e2 = e2 = 1
r
and
K 2 dr 2
ds2 = (1 )dt + + r 2 d2 (7.20)
r 1 Kr

We can find K by going to the Newtonian limit. We calculate the gravita-


tional acceleration ( that is, the acceleration of a free particle instantanously at
rest ) in the limit of a weak field of a particle at a distance r from a spherical
mass M . Newtonian:
d2 r GM
g= 2 = 2 (7.21)
dt r
We anticipate that r >> K. Then the proper time qof a particle will be
approximately equal to the coordinate time, since d = 1 Kr dt
The acceleration of a particle in 3-space, is given by the geodesic equation:

d2 x
+ u u = 0
d 2 (7.22)
dx
u =
d
For a particle instantanously at rest in a weak field, we have d dt. Using
u = (1, 0, 0, 0), we get:
d2 r
g = 2 = rtt (7.23)
dt
This equation gives a physical interpretation of rtt as the gravitational acceler-
ation. This is a mathematical way to express the principle of equivalence: The
gravitational acceleration can be transformed to 0, since the Christoffel symbols
always can be transformed to 0 locally, in a freely falling non-rotating frame,
i.e. a local inertial frame.

1 r gt gt gtt 
rtt = g +
2 |{z} | t
{z } | t
{z } x
1
gr =0 =0
1 gtt
=
2grr r
K gtt K
gtt = (1 ) , = 2 (7.24)
r r r
K GM
g = rtt = 2 = 2
2r r
gives K = 2GM
2GM
or with c: K =
c2
104 Chapter 7. The Schwarzschild spacetime

Then we have the line element of the exterior Schwarzschild metric:


2GM dr 2
ds2 = (1 2 )c2 dt2 + 2GM
+ r 2 d2 (7.25)
c r 1 c2 r
RS 2GMc2
is the Schwarzschild radius of a mass M.
Weak field: r >> RS .
For the earth: RS 0.9cm
For the sun: RS 3km
A standard clock at rest in the Schwarzschild spacetime shows a proper time :
r
RS
d = 1 dt (7.26)
r
So the coordinate clocks showing t, are ticking with the same rate as the stan-
dard clocks far from M. Coordinate clocks are running equally fast no matter
where they are. If they hadnt, the spatial distance between simultanous events
with given spatial coordinates, would depend on the time of the measuring of
the distance. Then the metric would be time dependent. Time is not running
at the Schwarzschild radius.

Definition 7.1.1 (Physical singularity)


A physical singularity is a point where the curvature is infinitely large.

Definition 7.1.2 (Coordinate singularity)


A coordinate singularity is a point (or a surface) where at least one of the components
of the metric tensor is infinitely large, but where the curvature of spacetime is finite.

Kretschmanns curvature scalar is R R . From the Schwarzschild metric,


we get:
48G2 M 2
R R = (7.27)
r8
which diverges only at the origin. Since there is no physical singularity at
r = RS , the singularity here is just a coordinate singularity, and can be re-
moved by a transformation to a coordinate system falling inward. (Eddington -
Finkelstein coordinates, Kruskal - Szekers analytical extension of the description
of Schwarzschild spacetime to include the area inside R S ).

7.2 Radial free fall in Schwarzschild spacetime


The Lagrangian function of a particle moving radially in Schwarzschild space-
time
1 RS 2 2 1 r 2 d
L = (1 )c t + R
, (7.28)
2 r 2 (1 rS ) d
7.2 Radial free fall in Schwarzschild spacetime 105

where is the time measured on a standard clock which the particle is carrying.
The momentum conjugate pt of the cyclic coordinate t, is a constant of motion.

L RS 2
pt = = (1 )c t (7.29)
t r

4-velocity identity: u u = c2 :

RS 2 2 r 2
(1 )c t + RS
= c2 (7.30)
r 1 r

Inserting the expression for t gives:

p2t RS 2
r 2 = (1 )c (7.31)
c2 r

Boundary conditions: the particle is falling from rest at r = r 0 .


r
RS c2 RS 2
pt = (1 )q = 1 c (7.32)
r0 1 RS r0
r0
| {z }
t(r=r0 )

giving
r r
dr RS r0 r
r = = c (7.33)
d r0 r
Z r
dr RS
q = c (7.34)
r0 r r0
r

Integration with = 0 for r = 0 gives:


r r r r
r0 r0 r r r
= (arcsin 1 ) (7.35)
c RS r0 r0 r0

is the proper time that the particle spends on the part of the fall which is
from r to r=0. To the lowest order in rr0 we get:
r
2 r r
= (7.36)
3 RS c

Travelling time from r = RS to r = 0 for RS = 2km is then:

2 RS
| (RS )| = = 4 106 s (7.37)
3 c
106 Chapter 7. The Schwarzschild spacetime

7.3 Light cones in Schwarzschild spacetime


The Schwarzschild line-element (with c = 1) is
RS 2 dr 2
ds2 = (1 )dt + + r 2 d2 (7.38)
r (1 RrS )
We will look at radially moving photons (ds 2 = d2 = 0). We then get
r 1
dr RS r 2 dr r RS
q = 1 dt = 1 dt
1 rRS r r R S r 2
(7.39)
rdr
= dt
r RS
with + for outward motion and for inward motion. For inwardly moving
photons, integration yields
r
r + t + RS ln | 1| = k = constant (7.40)
RS
We now introduce a new time coordinate t 0 such that the equation of motion
for photons moving inwards takes the following form
dr
r + t0 = k = 1
dt0 (7.41)
r
t0 = t + RS ln | 1|
RS
The coordinate t0 is called an ingoing Eddington-Finkelstein coordinate. The
photons here always move with the local velocity of light, c. For photons moving
outwards we have
r
r + RS ln | 1| = t + k (7.42)
RS
Making use of t = t0 RS ln | RrS 1| we get
r
r + 2RS ln | 1| = t0 + k
RS
dr 2RS dr r + RS dr
+ =1 =1 (7.43)
dt0 r RS dt 0 r RS dt0
dr r RS
=
dt0 r + RS
Making use of ordinary Schwarzschild coordinates we would have gotten the
following coordinate velocities for inn- and outwardly moving photons:
dr RS
= (1 ) (7.44)
dt r
which shows us how light is decelerated in a gravitational field. Figure 7.1 shows
how this is viewed by a non-moving observer located far away from the mass. In
Figure 7.2 we have instead used the alternative time coordinate t 0 . The special
theory of relativity is valid locally, and all material particles thus have to remain
inside the light cone.
7.3 Light cones in Schwarzschild spacetime 107

t
Collaps of light cone
at the horizon, r = R S
?

Trajectory of transmitter

horizon
Light cone

r
RS

Figure 7.1: At a radius r = RS the light cones collapse, and nothing can any
longer escape, when we use the Schwarzschild coordinate time.

Trajectory of transmitter
horizon

r
RS

Figure 7.2: Using the ingoing Eddingto n-Finkelstein time coordinate there is no
collapse of the light cone at r = RS . Instead we get a collapse at the singularity
at r = 0. The angle between the left part of the light cone and the t 0 -axis is
always 45 degrees. We also see that once the transmitter gets inside the horizon
at r = RS , no particles can escape.
108 Chapter 7. The Schwarzschild spacetime

7.4 Analytical extension of the Schwarzschild space-


time
The Schwarzschild coordinates are comoving with a static reference frame out-
side a spherical mass distribution. If the mass has collapsed to a black hole
there exist a horizon at the Schwarzschild radius. As we have seen in section
7.3 there do not exist static observers at finite radii inside the horizon. Hence,
the Schwarzschild coordinates are well defined only outside the horizon.
Also the rr-component of the metric tensor has a coordinate singularity at
the Schwarzschild radius. The curvature of spacetime is finite here.
Kruskal and Szekeres have introduced new coordinates that are well defined
inside as well as outside the Schwarzschild radius, and with the property that
the metric tensor is non-singular for all r > 0.
In order to arrive at these coordinates we start by considering a photon
moving radially inwards. From eq. (7.40) we then have

r
t = r RS ln 1 + v (7.45)
RS
where v is a constant along the world line of the photon. We introduce a new
radial coordinate
r

r r + RS ln 1 (7.46)
RS
Then the equation of the worldline of the photon takes the form
t + r = v (7.47)
The value of the constant v does only depend upon the point of time when the
photon was emitted. We may therefore use v as a new time coordinate.
For an outgoing photon we get in the same way
t r = u (7.48)
where u is a constant of integration, which may be used as a new time coordinate
for outgoing photons. The coordinates u and v are the generalization of the light
cone coordinates of Minkowski spacetime to the Schwarzchild spacetime.
From eqs. (7.47) and (7.48) we get
1
dt = (dv + du) (7.49)
2
1
dr = (dv du) (7.50)
2
and from eq. (7.46)  
Rs
dr = 1 dr (7.51)
r
Inserting these differentials into eq. (7.38) we arrive at a new form of the
Schwarzschild line-element,
 
2 Rs
ds = 1 du dv + r 2 d2 (7.52)
r
7.4 Analytical extension of the Schwarzschild spacetime 109

The metric is still not well behaved at the horizon. Introducing the coordinates
u
U = e 2Rs (7.53)
v
V =e 2Rs (7.54)

gives
vu r Rs r
U V = e 2Rs = e Rs
= 1 e Rs (7.55)
r
and
dU dV
du dv = 4Rs2 (7.56)
UV
The line-element (7.52) then takes the form

4Rs3 Rr
ds2 = e s dU dV + r 2 d2 (7.57)
r
This is the first form of the Kruskal-Szekeres line-element. Here is no coordinate
singularity, only a physical singularity at r = 0.
We may furthermore introduce two new coordinates
 
1 r r t
T = (V + U ) = 1 e 2Rs sinh (7.58)
2 Rs 2Rs
 
1 r r t
Z = (V U ) = 1 e 2Rs cosh (7.59)
2 Rs 2Rs

Hence

V =T +Z (7.60)
U =T Z (7.61)

giving
dU dV = dT 2 dZ 2 (7.62)
Inserting this into eq. (7.57) we arrive at the second form of the Kruskal-Szekeres
line-element
4R 3 r 
ds2 = s e Rs dT 2 dZ 2 + r 2 d2 (7.63)
r
The inverse transformations of eqs. (7.58) and (7.59) is
 
r r
1 e Rs = T 2 Z 2 (7.64)
Rs
t T
tanh = (7.65)
2Rs Z

Note from eq. (7.63) that with the Kruskal-Szekeres coordinates T and Z
the equation of the radial null geodesics has the same form as in flat spacetime

Z = T + constant (7.66)
110 Chapter 7. The Schwarzschild spacetime

7.5 Embedding of the Schwarzschild metric


We will now look at a static, spherically symmetric space. A curved simultaneity
plane (dt = 0) through the equatorial plane (d = 0) has the line element

ds2 = grr dr 2 + r 2 d2 (7.67)

with a radial coordinate such that a circle with radius r has a circumference of
length 2r.
We now embed this surface in a flat 3-dimensional space with cylinder co-
ordinates (z, r, ) and line element

ds2 = dz 2 + dr 2 + r 2 d2 (7.68)

The surface described by the line element in (7.67) has the equation z = z(r).
The line element in (7.68) is therefore written as

dz 2 2
ds2 = [1 + ( ) ]dr + r 2 d2 (7.69)
dr
Demanding that (7.69) is in agreement with (7.67) we get

dz 2 dz p
grr = 1 + ( ) = grr 1 (7.70)
dr dr
Choosing the positive solution gives
p
dz = grr 1dr (7.71)

In the Schwarzschild spacetime we have


1
grr = (7.72)
1 RrS

Making use of this we find z:


Z r p
dr
z= q = 4RS (r RS ) (7.73)
RS RS
1 r

This is shown in Figure 7.3.

7.6 Deceleration of light


The speed of light in Schwarzschild coordinates is
RS
c = 1 (7.74)
r
To measure this effect one can look at how long it takes for light to get from
Mercury to the Earth. This is illustrated in Figure 7.4. The travel time from
7.6 Deceleration of light 111

Figure 7.3: Embedding of the Schwarzschild metric.

Mercury
z2 >0
r2

Sun b
x

r1

z1 <0
Earth

Figure 7.4: General relativity predicts that light traveling from Mercury to the
Earth will be delayed due to the effect of the Suns gravity field on the speed of
light. This effect has been measured.
112 Chapter 7. The Schwarzschild spacetime

z1 to z2 is
Z z2 Z z2 Z z2
dz RS RS
t =
(1 + )dz = (1 + )dz
z1 1 RrS
z1 r z1 b + z2
2
(7.75)
z2 2 + b2 + z2
= z2 + |z1 | + RS ln 2
z1 + b2 |z1 |
where RS is the Schwarzschild radius of the Sun.
The deceleration is greatest when Earth and Mercury (where the light is
reflected) are on nearly opposite sides of the Sun. The impact parameter b is
then small. A series expansion to the lowest order of b/z gives

4|z1 |z2
t = z2 + |z1 | + RS ln (7.76)
b2
The last term represents the extra traveling time due to the effect of the Suns
gravity field on the speed of light. The journey takes longer time:

RS = the Schwarzschild radius of the Sun 2km


|z1 | = the radius of Earths orbit = 15 10 10 m
z2 = the radius of Mercurys orbit = 5.8 10 10 m
b = R 8
= 7 10 m

give a delay of 1.1 104 s. In addition to this one must also, of course, take
into account among other things the effects of the curvature of spacetime near
the Sun and atmospheric effects on Earth.

7.7 Particle trajectories in Schwarzschild 3-space

1
L= g X X
2
  1 2 (7.77)
1 Rs 2 r 1 1
= 1 t + 2 Rs + r 2 2 + r 2 sin2 2
2 r 1 r 2 2

Since t is a cyclic coordinate


 
L Rs
pt = = 1 t = constant = E (7.78)
t r

where E is the particles energy as measured by an observer "far away" (r  R s ).


Also is a cyclic coordinate so that
L
p = = r 2 sin2 = constant (7.79)

where p is the particles orbital angular momentum.
7.7 Particle trajectories in Schwarzschild 3-space 113

~ 2 = g X X = 1 we transform
Making use of the 4-velocity identity U
the above to get
 
Rs r 2
1 t2 + + r 2 2 + r 2 sin2 2 = 1 (7.80)
r 1 Rrs

E p
which on substitution for t = and = r 2 sin2
becomes
1 Rrs

E2 r 2 2 2
p2
+ + r + = 1 (7.81)
1 Rrs 1 Rrs r 2 sin2

Now, refering back to the Lagrange equation


 
d L L
=0 (7.82)
d X X

we get, for

(r 2 ) = r 2 sin cos 2
p2 cos (7.83)
=
r 2 sin3

Multiplying this by r 2 we get

cos 2
(r 2 )(r 2 ) = p (7.84)
sin3

which, on integration, gives


 p 2

(r 2 )2 = k (7.85)
sin

where k is the constant of integration.


Because of the spherical geometry we are free to choose a coordinate system
such that the particle moves in the equatorial plane and along the equator at a
given time t = 0. That is = 2 and = 0 at time t = 0. This determines the
constant of integration and k = p2 such that
 
1
(r 2 )2 = p2 1 (7.86)
sin2

The RHS is negative for all 6= 2 . It follows that the particle cannot deviate
from its original (equatorial) trajectory. Also, since this particular choice of
trajectory was arbitrary we can conclude, quite generally, that any motion of
free particles in a spherically symmetric gravitational field is planar motion.
114 Chapter 7. The Schwarzschild spacetime

7.7.1 Motion in the equatorial plane

E2 r 2 p2
+ + = 1 (7.87)
1 Rrs 1 Rrs r2

that is
  !
2 2 Rs p2
r = E 1 1+ (7.88)
r r2

This corresponds to an energy equation with an effective potential V (r) given


by
  !
2 Rs p2
V (r) = 1 1+ 2
r r
r 2 + V 2 (r) = E 2
s
2 2 (7.89)
rs p Rs p
V = 1 + 2 3
r r r
2
1 Rs 1 p
u1 +
2 r 2 r2
Newtonian potential VN is defined by using the last expression in

GM p2
VN = V 1 V N = + 2 (7.90)
r 2r
The possible trajectories of particles in the Schwarzschild 3-space are shown
schematically in Figure 7.5 as functions of position and energy of the particle
in the Newtonian limit.
To take into account the relativistic effects the above picture must be mod-
ified. We introduce dimensionless variables
r p
X= and k= (7.91)
GM GM m

The potential V 2 (r) now take the form


 1/2
2 k2 2k 2
V = 1 + 2 3 (7.92)
X X X

For r equal to the Schwarzschild radius (X = 2) we have


r
k 2 2k 2
V (2) = 11+ =0 (7.93)
4 8

For k 2 < 12 particles will fall in towards r = 0.


7.7 Particle trajectories in Schwarzschild 3-space 115

Figure 7.5: Newtonian particle trajectories are functions of the position and
energy of the particle. Note the centrifugal barrier. Due to this particles
with p 6= 0 cannot arrive at r = 0.

An orbit equation is one which connects r and . So for motion in the


equatorial plane for weaks fields we have
d p d p d
= = (7.94)
dt mr 2 d mr 2 d
1
Introducing the new radial coordinate u r our equations transform to
du 1 dr 1 mr 2 dr m
= 2 = 2 = r
d r d r p dt p
(7.95)
p du
r =
m d
Substitution from above for r in the energy equation yields the orbit equation,
 2 !
du m 2 E2
+ (1 2GM u) u2 + 2 = 2 . (7.96)
d p p
Differentiating this, we find
d2 u GM m2
+ u = + 3GM u2 (7.97)
d2 p2
The last term on the RHS is a relativistic correction term.
116 Chapter 7. The Schwarzschild spacetime

1.05

20 40 60 80 100
0.95

0.9

Figure 7.6: When relativistic effects are included there is no longer a limit to
the values that r can take and collapse to a singularity is "possible". Note that
V 2 is plotted here.

7.8 Classical tests of Einsteins general theory of rel-


ativity

7.8.1 The Hafele-Keating experiment

Hafele and Keating measured the difference in time shown on moving and sta-
tionary atomic clocks. This was done by flying around the Earth in the East-
West direction comparing the time on the clock in the plane with the time on a
clock on the ground.
dxi
The proper time interval measured on a clock moving with a velocity v i = dt
in an arbitrary coordinate system with metric tensor g is given by

g 1
d = ( dx dx ) 2 , dx0 = cdt
c2
vi v2 1 (7.98)
= (g00 2gi0 2 ) 2 dt
c c
v 2 gij v i v j
7.8 Classical tests of Einsteins general theory of relativity 117

For a diagonal metric tensor (gi0 = 0) we get

v2 1
d = (g00 ) 2 dt , v 2 = gii (v i )2 (7.99)
c2

We now look at an idealized situation where a plane flies at constant altitude


and with constant speed along the equator.

RS v2 1
d = (1 2 ) 2 dt , r =R+h (7.100)
r c

RS v2
To the lowest order in r and c2
we get

RS 1 v2
d = (1 )dt (7.101)
2r 2 c2

The speed of the moving clock is

v = (R + h) + u (7.102)

where is the angular velocity of the Earth and u is the speed of the plane. A
series expansion and use of this value for v gives

GM 1 R2 2 gh 2Ru + u2 GM
= (1 + 2 )t , g= R2
Rc2 2 c2 c 2c2 R2
(7.103)

u > 0 when flying in the direction of the Earths rotation, i.e. eastwards. For a
clock that is left on the airport (stationary, h = u = 0) we get

GM 1 R 2 2
0 = (1 )t (7.104)
Rc2 2 c2

To the lowest order the relative difference in travel time is

0 gh 2Ru + u2
k= = 2 (7.105)
0 c 2c2

Measurements:
Travel time: 0 = 1.2 105 s (a little over 24h)
Traveling eastwards: ke = 1.0 1012
Traveling westwards: kw = 2.1 1012
( 0 )e = 1.2 107 s 120ns
( 0 )w = 2.5 107 s 250ns
118 Chapter 7. The Schwarzschild spacetime

7.8.2 Mercurys perihelion precession


The orbit equation for a planet orbiting a star of mass M is given by equation
(7.97),
d2 u GM m2
+ u = + ku2 (7.106)
d2 p 2
where k = 3GM . We will be slightly more general, and allow k to be a theory-
or situation dependent term. This equation has a circular solution, such that
GM m2
u0 = + ku0 2 (7.107)
p 2
With a small perturbation from the circular motion u is changed by u 1 , where
u1  u0 . To lowest order in u1 we have
d2 u1 GM m2
+ u 0 + u 1 = + ku0 2 + 2ku0 u1 (7.108)
d2 p 2
or
d2 u1 d2 u1
+ u1 = 2ku0 u1 + (1 2ku0 )u1 = 0 (7.109)
d2 d2
For ku0  1 the equilibrium orbit is stable and we get a periodic solution:
p
u1 = u0 cos[ 1 2ku0 ( 0 )] (7.110)
where  and 0 are integration constants.  is the eccentricity of the orbit. We
can choose 0 = 0 and then have
1 p
= u = u0 + u1 = u0 [1 +  cos( 1 2ku0 ) (7.111)
r

Let f 1 2ku0
1 1
= (1 +  cos f ) (7.112)
r r0
For f = 1 (k = 0, no relativistic term) this expression describes a non-precessing
elliptic orbit (a Kepler-orbit).
For f < 1 (k > 0) the ellipse is not closed. To give the same value for r as
on a given starting point, has to increase by 2 f > 2. The extra angle per
rotation is 2( f 1) = 1 .
1

1
1 = 2( 1) 2ku0 (7.113)
1 2ku0
Using general relativity we get for Mercury
GM m2
k = 3GM = 6GM u0 6GM (7.114)
p 2

GM m 2
= 6( ) per orbit. (7.115)
p

which in Mercurys case amounts to () century = 4300


7.8 Classical tests of Einsteins general theory of relativity 119

7.8.3 Deflection of light


The orbit equation for a free particle with mass m = 0 is
u 2 1
( ) + B 2 = 2 (7.116)
b
p
where b = E = impact parameter. For a photon b = 1 , p = Eb2 , where p
is the photons angular momentum.

B 2 = (1 2GM u)u2 , B 2 max = 27G2 M 2 , rmax = 3GM (7.117)

Light falling towards the body with b > B max will be deflected and pass M ,
while light with a smaller impact parameter will hit M .
The straight, dashed line shown in Figure 7.7 is given by cos = rb0 = bu0 ,
which gives the unperturbed solution for u 0 :
1
u0 = cos (7.118)
b
The photon trajectory (the curve in Figure 7.7) is a perturbation of the straight

photon

2

b
r


M, GM<<b
Sun
Figure 7.7: Light traveling close to a massive object is deflected.

(dashed) line:

u = u 0 + u1 , u1  u0 (7.119)

To the 1st order in u1 we get


u 1 u1
= sin +
b
u 1 2 u1
( )2 2 sin2 sin (7.120)
b b
1 1 2 u1
= 2 2 cos2 sin
b b b
120 Chapter 7. The Schwarzschild spacetime

Since u1 is small we get


1 2
u2 u0 2 + 2u0 u1 = 2
cos2 + cos u1
b b
2GM 1 2 GM
B 2 (1 cos )( 2 cos2 + cos u1 ) , 1 (7.121)
b b b b
1 2 2GM 3 2
2 cos cos + cos u1
b b3 b
Inserting this into the orbit equation gives
1 2 u1 1 2GM 2 1
2
sin2 sin + 2 cos2 3
cos3 + cos u1 = 2
b b b b b b
u1 GM
tan + u1 = 2 cos2
b
1 u1 cos GM cos3 GM cos
+ 2 u 1 = 2 2 = 2 ( 2 cos )
sin sin b sin b sin
u1 GM cos
d( ) = 2 ( 2 cos )d
sin b sin
(7.122)
Integration gives
u1 GM 1
= 2 ( sin ) + K
sin b sin
(7.123)
GM
u1 = 2 (1 + sin2 ) + K sin
b
where K is an integration constant. From Figure 7.7 it follows that the solution
must be symmetric about = 0. sin is antisymmetric, and we must therefore
have K = 0
GM
u1 = (1 + sin2 ) (7.124)
b2
The trajectory of the photon is
1 GM
u = u0 + u1 = [cos + (2 cos2 )] (7.125)
b b
To find out how much the light is deflected, we let r , (u 0). We will
then have 2 , and since cos 2 = 0 we can neglect the cos2 -term.
2GM
cos + 0 when r
b (7.126)
2GM
cos 0 =
b
From Figure 7.7 we get
2GM
cos( )=
2 2 b (7.127)
2GM
cos cos + sin sin = , sin
2 2 2 2 b 2 2
7.8 Classical tests of Einsteins general theory of relativity 121

4GM
|| (7.128)
b

For light traveling in a tangent line trajectory to the surface of the Sun we get

= 1.7500 (7.129)
Chapter 8

Black Holes

8.1 Surface gravity:gravitational acceleration on the


horizon of a black hole
Surface gravity is denoted by 1 and is defined by
a p
= lim a= a a (8.1)
rr+ ut

where r+ is the horizon radius, r+ = RS for the Schwarzschild spacetime, u t is


the time component of the 4-velocity.
The 4-velocity of a free particle instantanously at rest in the Schwarzschild
spacetime:
dt 1 ~et
~u = ut~et = ~et = ~et = q (8.2)
d gtt 1 RS r

The only component of the 4-acceleration different from zero, is a r . The


4-acceleration:~a = u~ ~u = u; u ~e = (u, + u )u ~e .

ar = (ur, + r u )u
= ur, u +rtt (ut )2
| {z }
=0
rtt
=
1 RrS
1 gtt RS (8.3)
rtt = = 2
2 r 2r
RS
2r 2
ar =
1 RrS
ar RS RS
ar = g rr ar = = (1 )ar = 2
grr r 2r
RS
r
The acceleration scalar: a = ar a = q 2r 2
R
(measured with standard instru-
1 rS

122
8.2 Hawking radiation:radiation from a black hole (1973) 123

ments: at the horizon, time is not running).

a RS
= 2 (8.4)
ut 2r
With c:
a c2 RS GM
t
= 2
= 2 (8.5)
u 2r r
a 1 1
= lim t = = (8.6)
rRS u 2RS 4GM
c 2
Including c the expression is = 4GM . On the horizon of a black hole with one
m
solar mass, we get = 2 10 s2 .
13

8.2 Hawking radiation:radiation from a black hole (1973)


The radiation from a black hole has a thermal spectrum. We are going to find
the temperature of a Schwarzschild black hole of mass M. The Planck spectrum
has an intensity maximum at a wavelength given by Wiens displacement law.

N ~c
= where k is the Boltzmann constant, and N=0.2014
kT
For radiation emitted from a black hole, Hawking derived the following expres-
sion for the wavelength at a maximum intensity

8N GM
= 4N RS = (8.7)
c2
Inserting from Wiens displacement law, gives:

~c3 ~c
T = = (8.8)
8GkM 2k
Inserting values for ~, c and k gives:

2 104 m
T K (8.9)
RS

For a black hole with one solar mass,we have T 107 . When the mass is
decreasing because of the radiation, the temperature is increasing.So a black
hole has a negative heat capacity. The energy loss of a black hole because of
radiation, is given by the Stefan-Boltzmann law:

dM A
= T 4 2 (8.10)
dt c
where A is the surface of the horizon.
16G2 M 2
A = 4RS2 = (8.11)
c4
124 Chapter 8. Black Holes

gives:

dM 1 ~c6 Q
= 2
dt 2
15360 G M 2 M (8.12)
M (t) = (M03 3Qt)1/3 , M0 = M (0)

A black hole with mass M0 early in the history of the universe which is about
to explode now, had to have a starting mass

M0 = (3Qt0 )1/3 1012 kg (8.13)

about the mass of a mountain. They are called mini black holes.

8.3 Rotating Black Holes: The Kerr metric


This solution was found by Roy Kerr in 1963.
A time-independent, time-orthogonal metric is known as a static metric. A
time-independent metric is known as a stationary metric. A stationary metric
allows rotation.
Consider a stationary metric which describes a axial-symmetric space

ds2 = e2 dt2 + e2 dr 2 + e2 (d dt)2 + e2 d 2 , (8.14)

where , , , and are functions of r and .


By solving the vacuum field equations for this line-element, Kerr found the
solution:
2 2 2
e2 = , e2 = , e2 = sin2 , e2 = 2 ,
2 2
2M ar
= , where 2 = r 2 + a2 cos2
2
= r 2 + a2 2M r
2 = (r 2 + a2 )2 a2 sin2
(8.15)

This is the Kerr solution expressed in Boyer-Lindquist coordinates. The function


is the angular-velocity. The Kerr-solution is the metric for space-time outside
a rotating mass-distribution. The constant a is spin per mass-unit for the mass-
distribution and M is its mass.
Line-element:
2M r 2 2 2 4M ar
ds2 = (1 )dt + dr sin2 dtd + 2 d 2
2 2
(8.16)
2M a2 r
+ (r 2 + a2 + sin2 ) sin2 d2
2

(Here M is a measure of the mass so that M = G mass, ie. G = 1)


8.3 Rotating Black Holes: The Kerr metric 125

Light emitted from the surface, r = r 0 , where g tt = 0 is infinitely redshifted


further out. Observed from the outside time stands still.
2 = 2M r0 r02 + a2 cos2 = 2M r0
p (8.17)
r0 = M M 2 a2 cos2
This is the equation for the surface which represents infinite redshift.

8.3.1 Zero-angular-momentum-observers (ZAMOs)



The Lagrange function of a free particle in the equator plane, = 2
1 1 1 1
L = (e2 e2 )t2 + e2 r 2 + e2 2 + e2 2 e2 t (8.18)
2 2 2 2
Here = 0. The momentum p of the cyclic coordinates :
L dt d
p = e2 ( t) , t = , cot = (8.19)
d d
The angular speed of the particle relative to the coordinate system:
d
= = , = t
dt t (8.20)
p = e2 t( )
p is conserved during the movement.
2M ar
= ,
(r 2
+ a 2 )2
a2 (r 2 + a2 2M r) (8.21)
0 when r
When studying the Kerr metric one finds that Kerr Minkowski for large
r. The coordinate clocks in the Kerr space-time show the same time as the
standard-clocks at rest in the asymptotic Minkowski space-time.
A ZAMO is per definition a particle or observer with p = 0. p is a constant
of motion, so the stone remains a ZAMO during the movement. A local reference
frame which coincides with the stone is a local inertial frame.
d
p = 0 = = (8.22)
dt
That is, the local inertial frame obtains an angular speed relative to the BL-
system (Boyer-Lindquist system).
Since the Kerr metric is time independent, the BL-system is stiff. The
distant observer has no motion relative to the BL-system. To this observer the
BL-system will appear stiff and non-rotating. The observer will observe that
the local inertial system of the stone obtains an angular speed a is spin
d 2M ar per mass
== 2 (8.23) unity and
dt (r + a2 )2 a2 (r 2 + a2 2M r)
M a is spin
In other words, inertial systems at finite distances from the rotating mass M
are dragged with it in the same direction. This is known as inertial dragging
or the Lense-Thirring effect (about 1920).
126 Chapter 8. Black Holes

8.3.2 Does the Kerr space have a horizon?


Definition 8.3.1 (Horizon)
a surface one can enter, but not exit.

Consider a particle in an orbit with constant r and . Its 4-velocity is:

d~x dt d~x
~u = =
d d dt (8.24)
2 12 d
= (g tt 2g t g ) (1, ) , where =
dt
To have stationary orbits the following must be true

g 2 + 2g t + g tt < 0 (8.25)

This implies that must be in the interval

min < < max , (8.26)


q g
q g tt
where min = 2 g tt , max = + 2 g since g t = g .

Outside the surface with infinite redshift g tt < 0. That is can be negative,
zero and positive. Inside the surface r = r 0 with infinite redshift g tt > 0. Here
min > 0 and static particles, = 0, cannot exist. This is due to the inertial
dragging effect. The surface r = r0 is therefore known as the static border.
The interval of , where stationary orbits are allowed, is reduced to zero
g
when min = max , that is = g tt g tt = 2 g (equation for the horizon).

For the Kerr metric we have:

g tt = 2 g e2 (8.27)

Therefore the horizon equation becomes

e2 = 0 =0 r 2 2mr + a2 = 0 (8.28)

The largest solution is r+ = M + M 2 a2 and this is the equation for a
spherical surface. The static border is r 0 = M + M 2 a2 cos .
8.3 Rotating Black Holes: The Kerr metric 127

=0
horizon
r+

r0


>0 = 2
2M

=0
 K

M
static border
ergo-sphere stationary paths

Figure 8.1: Static border and horizon of a Kerr black hole


Chapter 9

Schwarzschilds Interior Solution

9.1 Newtonian incompressible star

2 = 4G, = (r)
1 d 2 d (9.1)
(r ) = 4G
r 2 dr dr
Assuming = constant.
d
d(r 2 ) = 4Gr 2 dr
dr
r2
d
=
4
Gr 3 + K (9.2)
dr 3
= M (r) + K

Gravitational acceleration: ~g = = d
dr ~
er
M (r) K1 4 K1
g= 2
+ 2 = Gr + 2 (9.3)
r r 3 r
Finite g in r = 0 demands K1 = 0.
4 d 4
g= Gr, = Gr (9.4)
3 dr 3
Assume that the massdistribution has a radius R.
2
= Gr 2 + K2 (9.5)
3
Demands continuous potensial at r = R.
2 M (R) 4
GR2 + K2 = = GR2
3 R 3 (9.6)
K2 = 2GR2
(with zero level at infinite distance). Gives the potensial inside the mass distri-
bution:
2
= G(r 2 3R2 ) (9.7)
3

128
9.1 Newtonian incompressible star 129

The star is in hydrostatic equilibrium, that is, the pressure forces are in equi-
librium with the gravitational forces.

dm=4 r 2 dr

4 3
r
3

Figure 9.1: The shell with thickness dr, is affected by both gravitational and
pressure forces.

Consider figure 9.1. The pressure forces on the shell is 4r 2 dp. Gravitational
forces on the shell:

mass inside shell mass of the shell


G
r2
4 3 2 (9.8)
r 4r dr
=G 3
r2

Equilibrium:

4 3
4r 2 dp = G r 4dr
3
4
dp = G2 rdr
3

2G 2 2 (9.9)
p = K3 r
3
2G 2 2
P (R) = 0 gives : K3 = R
3
2G 2 2
p(r) = (R r 2 )
3

No matter how massive the star is, it is possible for the pressure forces to keep
the equilibrium with gravity. In Newtonian theory, gravitational collapse is not
a necessity.
130 Chapter 9. Schwarzschilds Interior Solution

9.2 The pressure contribution to the gravitational mass


of a static, spherical symmetric system
Gravitational acceleration :
a p
g=+ , a= a a (9.10)
ut
We have the line element:
ds2 = e2(r) dt2 + e2(r) dr 2 + r 2 d2
(9.11)
gtt = e2 , grr = e2

gives (because of the gravitational acceleration)

g = +e 0 (9.12)

From the expressions for E tt , E rr , E , E follow (see Section 7.1)

20 2
E tt E rr E E = 2e2 ( + 00 + 0 0 0 ) . (9.13)
r
We also have
0 20 2
(r 2 e 0 ) = r 2 e ( + 00 + 0 0 0 ) , (9.14)
r
which gives
Z
1
g=+ (Ett E rr E E )r 2 e+ dr . (9.15)
2r 2

By applying Einsteins field equations

E = 8GT (9.16)

we get
Z
4G
g=+ (Ttt Trr T T )r 2 e+ dr . (9.17)
r2

This is the Tolman-Whittaker expression for gravitational acceleration.


The corresponding Newtonian expression is :
Z
4G
gN = 2 r 2 dr (9.18)
r
The relativistic gravitational mass density is therefore defined as

G = T tt + T rr + T + T (9.19)

For an isotropic fluid with

T tt = , T rr = T = T = p (9.20)

9.3 The Tolman-Oppenheimer-Volkov equation 131

we get G = + 3p (with c = 1), which becomes


3p
G = + (9.21)
c2
It follows that in relativity, pressure has a gravitational effect. Greater pressure
gives increasing gravitational attraction. Strain (p < 0) decreases the gravita-
tional attraction.
In the Newtonian limit, c , pressure has no gravitational effect.

9.3 The Tolman-Oppenheimer-Volkov equation


With spherical symmetry the spacetime line-element may be written
ds2 = e2(r) dt2 + e2(r) dr 2 + r 2 d2
(9.22)
E tt = 8GT tt , T = diag(, p, p, p)
From E tt we get
1 d
[r(1 e2 )] = 8G
r 2 dr Z r (9.23)
2 2
r(1 e ) = 2G 4r dr ,
0
Rr
where m(r) = 0 4r 2 dr giving
2Gm(r) 1
e2 = 1 = (9.24)
r g rr
From E rr we have
E rr = 8GT rr
2 d 2 1 (9.25)
e 2 (1 e2 ) = 8Gp
r dr r
We get
2 d 2Gm(r) 2Gm(r)
(1 ) = 8Gp
r dr r r3
d m(r) + 4r 3 p(r)
=G (9.26)
dr r(r 2Gm(r))
The relativistic generalized equation for hydrostatic equilibrium is T r
; = 0,
giving
T r
, + T
r
+ r T = 0
1 p
T r rr
, = T ,r = p ,r =
g rr r
dp (9.27)
T r
, = e
dr
T r = r p = trt p + r p
r T = r T = rtt + r p
132 Chapter 9. Schwarzschilds Interior Solution

In orthonormal basis we have


= =
(9.28)
r = r = r = r

; = 0 now takes the form:


T r

dp
e + trt p + rtt (9.29)
dr
We have

trt = trt = rtt = rtt (9.30)

dr , giving:
and we also have rtt = e d

dp d
+ (p + ) =0 (9.31)
dr dr
Inserting Equation 9.26 into Equation 9.31 gives

dp m(r) + 4r 3 p(r)
= G( + p) (9.32)
dr r(r 2Gm(r))

This is the Tolman-Oppenheimer-Volkov (TOV) equation. The component g tt =


e2(r) may now be calculated as follows
dp
= d , = constant
+p
(9.33)
ln( + p) = K
+ p = K1 e , p = K1 e

Hence
p
e = e(R) (1 + )1 (9.34)

where R is the radius of the mass distribution.

9.4 An exact solution for incompressible stars - Schwarzschilds


interior solution
The mass inside a radius r for an incompressiable star is
4
m(r) = r 3 (9.35)
3

2Gm(r) r2
e2 = 1 1 2 (9.36)
r a
9.4 An exact solution for incompressible stars - Schwarzschilds interior solution133

where
3 r3 r2
a2 = , m(r) = , r s = 2Gm = r (9.37)
8G 2Ga2 a2
TOV equation:
4
dp r 3 + 4r 3 p(r)
= G 3 ( + p(r))
dr r(r 2G 34 r 3 )
4 + 3p(r)
= G r( + p(r))
3 1 G 38 r 2
1 + 3p(r)
= 2 r( + p(r)) (9.38)
2a 1 r22
a
Z p Z r
dp 1 r
= 2 dr
0 ( + 3p)( + p 2a R 1 r22
a
r
p+ a2 R 2
=
3p + a2 r 2
So the relativistic pressure distribution is

a 2 r 2 a2 R 2
p(r) = , r R (9.39)
3 a2 R 2 a2 r 2
also
3 a2 r
a2 = , 2 = >1a>r (9.40)
8G r rs
To satisfy the condition for hydrostatic equilibrium we must have p > 0 or
p(0) > 0 which gives

a a2 R 2
p(0) pc = >0 (9.41)
3 a2 R 2 a
in which the numerator is positive so that
p
3 a2 R 2 > a
9a2 9R2 > a2
r
8 (9.42)
R< a
9
8 8 3 1
R2 < a = =
9 9 8G 3G
Stellar mass:
4 4 1 4R
M = R3 < R =
3 3 3G 9G
(9.43)
4 1
M<
9G 3G
134 Chapter 9. Schwarzschilds Interior Solution

For a neutron star we can use 1017 g /cm3 . An upper limit on the mass is
then M < 2.5 M Substitution for p in the expression for e gives
r r
3 Rs 1 Rs
e = 1 1 3 r2 (9.44)
2 R 2 R
The line element for the interior Schwarzschild solution is
r r !2
3 R s 1 R s dr 2
ds2 = 1 1 3 r 2 dt2 + + r 2 d, r R
2 R 2 R 1 RR
s 2
3 r
(9.45)
Chapter 10

Cosmology

10.1 Comoving coordinate system


We will consider expanding homogenous and isotropic models of the universe.
We introduce an expanding frame of reference with the galactic clusters as
reference particles. Then we introduce a comoving coordinate system in this
frame of reference with spatial coordinates , , . We use time measured on
standard clocks carried by the galactic clusters as coordinate time (cosmic time).
The line element can then be written in the form:

ds2 = dt2 + a(t)2 [d2 + r()2 d2 ] (10.1)

(For standard clocks at rest in the expanding system, d = d = 0 and ds 2 =


d 2 = dt2 ). The function a(t) is called the expansion factor, and t is called
cosmic time.
The physical distance to a galaxy with coordinate distance d from an ob-
server at the origin, is:

dlx = g d = a(t)d (10.2)

Even if the galactic clusters have no coordinate velocity, they do have a radial
velocity expressed by the expansion factor.
The value determines which cluster we are observing and a(t) how it is
moving. 4-velocity of a reference particle (galactic cluster):

dx dx
u = = = (1, 0, 0, 0) (10.3)
d dt
du
This applies at an abritrary time, that is du
dt = 0. Geodesic equation: dt +

u u = 0 which is reduces to: tt = 0

0 0 0
1 z}|{ z}|{ z}|{
tt = g ( gt,t + gt,t + gtt, ) = 0 (10.4)
2
We have that gtt = 1. This shows that the reference particles are freely falling.

135
136 Chapter 10. Cosmology

10.2 Curvature isotropy - the Robertson-Walker met-


ric
Introduce orthonormal form-basis:

t = dt = a(t)d = a(t)r()d
(10.5)
= a(t)r() sin d

Using Cartans 1st equation:

d = (10.6)

to find the connection forms. Then using Cartans 2nd structure equation to
calculate the curvature forms:

R = d + (10.7)

d 0 d
Calculations give: (notation: = dt , = d )

a t
Rti = i , i = , ,
a
a2 r 00 
Rj = 2 2 j , j = , (10.8)
a ra
a2 1 r 0 2 
R = 2 + 2 2 2 2

a r a r a
The curvature of 3-space (dt = 0) can be found by putting a = 1. That is:

r 00

3 R j j
=
r
(10.9)
1 r02 
3 R = 2 2
r r
The 3-space is assumed to be isotropic and homogenous. This demands

r 00 1 r 02
= =k , (10.10)
r r2
where k represents the constant curvature of the 3-space.
p
r 00 + kr = 0 and r 0 = 1 kr 2 (10.11)

Solutions with r(0) = 0, r 0 (0) = 1 :



kr = sinh( k) (k < 0)
r = (k = 0) (10.12)

kr = sin( k) (k > 0)
10.3 Cosmic dynamics 137

The solutions can be characterized by the following 3 cases:


p
r = sinh , dr = 1 + r 2 d, (k = 1)
r = , dr = d, (k = 0) (10.13)
p
r = sin , dr = 1 r 2 d, (k = 1)

In all three cases one may write dr = 1 kr 2 d, which is just the last equation
above.
dr 2
We now set d2 = 1kr 2 into the line-element :


ds2 = dt2 + a2 (t) d2 + r 2 ()d2
 
dr 2 (10.14)
= dt2 + a2 (t) + r 2
d 2
1 kr 2

The first expression is known as the standard form of the line-element, the
second is called the Robertson-Walker line-element.
The 3-space has constant curvature. 3-space is spherical for k = 1, Euclidean
for k = 0 and hyperbolic for k = 1.
Universe models with k = 1 are known as closed and models with k = 1
are known as open. Models with k = 0 are called flat even though these
models also have curved space-time.

10.3 Cosmic dynamics


10.3.1 Hubbles law
The observer is placed in origo of the coordinate-system; 0 = 0. The proper
distance to a galaxy with radial coordinate e is D = a(t)e . The galaxy has a
radial velocity:

dD a a
v= = ae = D = HD whereH = (10.15)
dt a a

The expansion velocity v is proportional to the distance D. This is Hubbles


law.

10.3.2 Cosmological redshift of light


te : the time interval in transmitter-position at transmission-time
t0 : the time interval in receiver-position at receiving-time

Light follows curves with ds2 = 0, with d = d = 0 we have :

dt = a(t)d (10.16)
138 Chapter 10. Cosmology

t
6
t0 + t0

t0

te + te

te -
0 = 0 e

Figure 10.1: Schematic representation of cosmological redshift

Integration from transmitter-event to receiver-event :


Z t0 Z 0
dt
= d = e
te a(t) e
Z t0 +t0 Z 0
dt
= d = e ,
te +te a(t) e

which gives
Z t0 +t0 Z t0
dt dt
=0 (10.17)
te +te a te a
or
Z t0 +t0 Z te +te
dt dt
=0 (10.18)
t0 a te a

Under the integration from te to te + te the expansion factor a(t) can be


considered a constant with value a(t e ) and under the integration from t0 to
t0 + t0 with value a(t0 ), giving:

te t0
= (10.19)
a(te ) a(t0 )
10.3 Cosmic dynamics 139

t0 and te are intervals of the light at the receiving and transmitting time.
Since the wavelength of the light is = ct we have:

0 e
= (10.20)
a(t0 ) a(te )

This can be interpreted as a stretching of the electromagnetic waves due to


the expansion of space. The cosmological redshift is denoted by z and is given
by:
0 e a(t0 )
z= = 1 (10.21)
e a(te )
Using a0 a(t0 ) we can write this as:
a0
1 + z(t) = (10.22)
a

10.3.3 Cosmic fluids


The energy-momentum tensor for a perfect fluid (no viscosity and no thermal
conductivity) is
T = ( + p)u u + pg (10.23)
In an orthonormal basis

T = ( + p)u u + p (10.24)

where is the Minkowski metric. We consider 3 types of cosmic fluid:

1. dust: p = 0,
T = u u (10.25)

2. radiation: p = 13 ,

4
T =u u + p
3 (10.26)

= (4u u + )
3
The trace

T = T = (4u u + ) = 0 (10.27)
3
3. vacuum: p = ,
T = (10.28)

If vacuum can be described as a perfect fluid we have p v = v , where


is the energy density. It can be related to Einsteins cosmological constant,
= 8Gv .
One has also introduced a more general type of vacuum energy given by
the equation of state p = w , where denotes that the vacuum energy is
140 Chapter 10. Cosmology

connected to a scalar field . In a homogeneous universe the pressure and the


density are given by
1 1 2
p = 2 V (), = + V () (10.29)
2 2
where V () is the potential for the scalar field. Then we have
1 2
2 V ()
w= 1 2
(10.30)
2 + V ()

The special case = 0 gives the Lorentz invariant vacuum with w = 1. The
more general vacuum is called quintessence.

10.3.4 Isotropic and homogeneous universe models


We will discuss isotropic and homogenous universe models with perfect fluid
and a non-vanishing cosmological constant . Calculating the components of
the Einstein tensor from the line-ement (10.14) we find in an orthonormal basis

3a2 3k
Ett = + 2 (10.31)
a2 a
2a a2 k
Emm = 2 2. (10.32)
a a a
The components of the energy-momentum tensor of a perfect fluid in a comoving
orthonormal basis are
Ttt = , Tmm = p. (10.33)
Hence the tt component of Einsteins field equations is

a2 + k
3 = 8G + (10.34)
a2
mm components:
a a2 k
2 2 2 = 8Gp (10.35)
a a a
where is the energy density and p is the pressure. The equations with vanishing
cosmological constant are called the Friedmann equations. Inserting eq. (10.34)
into eq. (10.35) gives:
4G
a = a( + 3p) (10.36)
3
If we interpret as the mass density and use the speed of light c, we get
4G
a = a( + 3p/c2 ) (10.37)
3
Inserting the gravitational mass density G from eq.(9.21) this equation takes
the form
4G
a = aG (10.38)
3
10.3 Cosmic dynamics 141

Inserting p = wc2 into (9.21) gives

G = (1 + 3w) (10.39)

which is negative for w < 1/3, i.e. for 2 < V (). Special cases:

dust: w = 0, G =

radiation: w = 13 , G = 2

Lorentz-invariant vacuum: w = 1, G = 2

In a universe dominated by a Lorentz-invariant vacuum the acceleration of the


cosmic expansion is
8G
av = av > 0, (10.40)
3
i.e. accelerated expansion. This means that vacuum acts upon itself with repul-
sive gravitation.
The field equations can be combined into
 2
2 a 8G k
H = m + 2 (10.41)
a 3 3 a

where m is the density of matter, = 8G where is the vacuum energy


with constant density. = m + is the total mass density. Then we may
write
8G k
H2 = 2 (10.42)
3 a
The critical density cr is the density in a universe with euclidean spacelike
geometry, k = 0, which gives
3H 2
cr = (10.43)
8G
We introduce the relative densities
m
m = , = (10.44)
cr cr
Furthermore we introduce a dimensionless parameter that describes the curva-
ture of 3-space
k
k = 2 2 (10.45)
a H
Eq. (10.42) can now be written

m + + k = 1 (10.46)

From the Bianchi identity and Einsteins field equations follow that the energy-
momentum density tensor is covariant divergence free. The time-component
expresses the equation of continiuty and takes the form

[( + p)ut u ]; + (p t ); = 0 (10.47)
142 Chapter 10. Cosmology

Since ut = 1, um = 0 and tt = 1, tm = 0, we get

( + p). + ( + p)u; p = 0 (10.48)

or
+ ( + p)(u, + t ) = 0 (10.49)

Here u, = 0 and t tt = 0. Calculating mtm for d = we get


a
mtm = rtr + t + = 3 (10.50)
t a
Hence
a
+ 3( + p) =0 (10.51)
a
which may be written
(a3 ). + p(a3 ). = 0 (10.52)

Let V = a3 be a comoving volume in the universe and U = V be the energy


in the comoving volume. Then we may write

dU + pdV = 0 (10.53)

This is the first law of thermodynamics for an adiabatic expansion. It follows


that the universe expands adiabatically. The adiabatic equation can be written

a
= 3 (10.54)
+p a

Assuming p = w we get

d da
= 3(1 + w)
a
 3(1+w)
a
ln = ln
0 a0

It follows that
 3(1+w)
a
= 0 (10.55)
a0
This equation tells how the density of different types of matter depends on the
expansion factor
a3(1+w) = constant (10.56)

Special cases:

dust: w = 0 gives d a3 = constant


Thus, the mass in a comoving volume is constant.
10.4 Some cosmological models 143

radiation: w = 31 gives r a4 = constant


Thus, the radiation energy density decreases faster than the
case with dust when the universe is expanding. The energy
in a comoving volume is decreasing because of the thermo-
dynamic work on the surface. In a remote past, the density
of radiation must have exceeded the density of dust:

d0 a30 =d a3
r0 a40 =r a4
r a4 r0 a40
=
d a3 d0 a30

The expansion factor when r = d :


r0
a(t1 ) = a0
d0
Lorentz-invariant vacuum: w = 1 gives = constant.
The vacuum energy in a comoving volume is increasing a 3 .

10.4 Some cosmological models


10.4.1 Radiation dominated model
The energy-momentum tensor for radiation is trace free. According to the Ein-
stein field equations the Einstein tensor must then be trace free:

aa + a2 + k = 0
(10.57)
(aa + kt) = 0

Integration gives
aa + kt = B (10.58)
Another integration gives

1 2 1 2
a + kt = Bt + C (10.59)
2 2
The initial condition a(0) = 0 gives C = 0. Hence
p
a= 2Bt kt2 (10.60)

For k = 0 we have
r
B
a = 2Bt , a = (10.61)
2t

The expansion velocity reaches infinity at t = 0, (lim t0 a = )


144 Chapter 10. Cosmology

Figure 10.2: In a radiation dominated universe the expansion velocity reaches


infinity at t = 0.


R a4 = K , a= 2Bt
(10.62)
4R B 2 t2 = K
According to the Stefan-Boltzmann law we then have
R = T 4 4B 2 T 4 t2 = K
r
K1 K1 (10.63)
t= 2 T =
T t
where T is the temperature of the background radiation.

10.4.2 Dust dominated model


From the first of the Friedmann equations we have
8G 2
a2 + k =a (10.64)
3
We now introduce a time parameter given by
dt d 1 d
= a() =
d dt a d
(10.65)
da 1 da
So: a = =
dt a d
8G
We also introduce A 3 0 a0 .
3 The first Friedmann equation then gives
8G 3 8G
aa2 + ka = a = 0 a0 3 = A (10.66)
3 3
10.4 Some cosmological models 145

Using we get

1 da 2
( ) = A ka
a d
1 da 2 A
( ) = k (10.67)
a2 d a
r r r
1 da A A a
= k = 1 k
a d a a A

where
p a we chose the positive root. We now introduce u, given by a = Au , u =
2

A . We then get

da du
= 2Au (10.68)
d d

which together with the equation above give

1 du 1p
2Au = 1 ku2
Au2 d u
(10.69)
du 1
= d
1 ku2 2

This equation will first be integrated for k < 0. Then k = |k|, so that
Z
du
p = +K (10.70)
1 + |k|u 2 2

or arcsinh( ku) = 2 + K. The condition u(0) = 0 gives K = 0. Hence

k 1
a = sinh2 = (cosh 1) (10.71)
A 2 2
or
A
a= (cosh 1) (10.72)
2k
From eqs. (10.43), (10.44) and (10.66) we have

8G m0
A= m0 = H02 = H02 m0 (10.73)
3 cr0

From egs. (10.45) and (10.46) we get

k = H02 (m0 1) (10.74)

Hence, the scale factor of the negatively curved, dust dominated universe model
is
1 m0
a() = (cosh 1) (10.75)
2 1 m0
146 Chapter 10. Cosmology

Inserting this into eq. (10.65) and integrating with t(0) = (0) leads to

m0
t() = (sinh ) (10.76)
2H0 (1 m0 )3/2

Integrating eg. (10.69) for k = 0 leads to an Einstein-deSitter universe

t 2
a(t) = ( )3 (10.77)
t0

Finally integrating eg. (10.69) for k > 0 gives, in a similar way as for k < 0

1 m0
a() = (1 cos ) (10.78)
2 1 m0
m0
t() = ( sin ) (10.79)
2H0 (m0 1)3/2

We see that this is a parametric representation of a cycloid.

In the Einstein-deSitter model the Hubble factor is

a 21 2 1 2
H= = , t= = tH (10.80)
a 3t 3H 3

The critical density in the Einstein-deSitter model is given by the first Fried-
mann equation:

8G
H2 = cr , k=0
3
(10.81)
3H 2
cr = , =
8G cr

Example 10.4.1 (Age-redshift relation for dust dominated universe with k = 0)

a0 a0
1+z = a=
a 1+z
a0 a (10.82)
da = dz = dz
(1 + z)2 1+z

Eq. (10.34) gives

a 2 8G 8G 0 a30
= =
a 3 3 a3 (10.83)
8G
= 0 (1 + z)3
3
10.4 Some cosmological models 147

Figure 10.3: For k = 1 the density is larger than the critical density, and the
universe is closed. For k = 0 we have = cr and the expansion velocity of
the universe will approach zero as t . For k = 1 we have < cr . The
universe is then open, and will continue expanding forever.

3
8G a da
Using H02 = 3 0 gives a = H0 (1 + z) 2 . From a = dt we get:

da da dz
dt = = a = 5 (10.84)
a aa H0 (1 + z) 2

Integration gives the age of the universe:

Z 0
1 dz 2 1  1 0
t0 = 5 = 3 (10.85)
H0 (1 + z) 2 3 H0 (1 + z) 2

t0 = 32 tH where the Hubble-time tH H10 is the age of the universe, if the expansion
rate had been constant. Look-back-time to a source with redshift z is:

Z z
dz 2  1 
t = tH 5 = tH 1 3 (10.86)
0 (1 + z) 2 3 (1 + z) 2
148 Chapter 10. Cosmology

exp.factor a

tangent

today

t
0 time t
t
H
Figure 10.4: tH is the age of the universe if the expansion had been constant,
BUT:The exp.rate was faster closer to the Big Bang, so the age is lower.

1
t = t0 [1 ] (10.87)
(1 + z)3/2

1
z= t 2/3
1 (10.88)
(1 t0 )

t
= 0, 99 z = 20, 5 (10.89)
t0

10.4.3 Friedmann-Lematre model


The dynamics of galaxies and clusters of galaxies has made it clear that far
stronger gravitational fields are needed to explain the observed motions than
those produced by visible matter (McGaugh 2001). At the same time it has
become clear that the density of this dark matter is only about 30% of the critical
10.4 Some cosmological models 149

density, although it is a prediction by the usual versions of the inflationary


universe models that the density ought to be equal to the critical density (Linde
2001). Also the recent observations of the temperature fluctuations of the cosmic
microwave radiation have shown that space is either flat or very close to flat
(Bernadis et.al 2001, Stompor et al. 2001, Pryke et al. 2001). The energy that
fills up to the critical density must be evenly distributed in order not to affect
the dynamics of the galaxies and the clusters.
Furthermore, about two years ago observations of supernovae of type Ia with
high cosmic red shifts indicated that the expansion of the universe is accelerating
(Riess et al. 1998, Perlmutter et al. 1999). This was explained as a result of
repulsive gravitation due to some sort of vacuum energy. Thereby the missing
energy needed to make space flat, was identified as vacuum energy. Hence, it
seems that we live in a flat universe with vacuum energy having a density around
70% of the critical density and with matter having a density around 30% of the
critical density.
Until the discovery of the accelerated expansion of the universe the standard
model of the universe was assumed to be the Einstein-DeSitter model, which is
a flat universe model dominated by cold matter. This universe model is thor-
oughly presented in nearly every text book on general relativity and cosmology.
Now it seems that we must replace this model with a new "standard model"
containing both dark matter and vacuum energy.
Recently several types of vacuum energy or so called quintessence energy
have been discussed (Zlatev, Wang and Steinhardt 1999, Carroll 1998). How-
ever, the most simple type of vacuum energy is the Lorentz invariant vacuum
energy (LIVE), which has constant energy density during the expansion of the
universe (Zeldovich 1968, Grn 1986). This type of energy can be mathemati-
cally represented by including a cosmological constant in Einsteins gravitational
field equations. The flat universe model with cold dark matter and this type of
vacuum energy is the Friedmann-Lematre model.
The field equations for the flat Friedmann-Lematre is found by putting
k = p = 0 in equation (10.35). This gives

a a2
2 + 2 = (10.90)
a a
Integration leads to
3
aa2 = a +K (10.91)
3
where K is a constant of integration. Since the amount of matter in a volume
comoving with the cosmic expansion is constant, M a3 = M 0 a30 , where the
index 0 refers to measured values at the present time. Normalizing the expansion
factor so that a0 = 1 and comparing eqs.(10.42) and(10.91) then gives K =
(8G/3)M 0 . Introducing a new variable x by a3 = x2 and integrating once
more with the initial condition a(0) = 0 we obtain
 
3 3K 2 t 2
a = sinh , t = (10.92)
t 3
150 Chapter 10. Cosmology

The vacuum energy has a constant density given by

= 8G (10.93)

The critical density, which is the density making the 3-space of the universe flat,
is
3H 2
cr = (10.94)
8G
The relative density, i.e. the density measured in units of the critical density, of
the matter and the vacuum energy, are respectively
8GM
M = = (10.95)
cr 3H 2

= = (10.96)
cr 3H 2
Since the present universe model has flat space, the total density is equal to
the critical density, i.e. M + = 1. Eq. (10.91) with the normalization
a(t0 ) = 1, where t0 is the present age of the universe, gives 3H 02 = 3K + .
Eq. (10.34) with k = 0 gives 8G0 = 3H02 . Hence K = 8G0 /3 and
8G0
3K
= = 0 = M
0
0
. In terms of the values of the relative densities at the
present time the expression for the expansion factor then takes the form
 
1/3 2/3 t M 0 1 0
a = A sinh , A= = (10.97)
t 0 0
p
Using the identity sinh(x/2) = (cosh x 1)/2 this expression may be written
   
3 A 2t
a = cosh 1 (10.98)
2 t
The age t0 of the universe
is found from a(t0 ) = 1, which by use of the formula
arc tanh x = arc sinh(x/ 1 x2 ), leads to the expression
p
t0 = t arc tanh 0 (10.99)

Inserting typical values t0 = 15 109 years, 0 = 0.7 we get A = 0.43, t =


12109 years. With these values the expansion factor is a = 0.75 sinh 2/3 (1.2t/t0 ).
This function is plotted in fig. 10.5. The Hubble parameter as a function of
time is
H = (2/3t ) coth(t/t ) (10.100)
Inserting t0 = 1.2t we get Ht0 = 0.8 coth(1.2t/t0 ), which is plotted in fig. 10.6
The Hubble parameter decreases all the time and approaches a constant value
H = 2/3t in the infinite future. The present value of the Hubble parameter
is
2
H0 = (10.101)
3t 0

The corresponding Hubble age is tH0 = (3/2)t 0 . Inserting our numerical
values gives H0 = 64km/secMpc1 and tH0 = 15.7 109 years. In this universe
10.4 Some cosmological models 151

a
1.5

1.25

0.75

0.5

0.25

0.2 0.4 0.6 0.8 1 1.2 1.4 t/t0

Figure 10.5: The expansion factor as function of cosmic time in units of the age
of the universe.

Figure 10.6: The Hubble parameter as function of cosmic time.


152 Chapter 10. Cosmology

model the age of the universe is nearly as large as the Hubble age, while in
the Einstein-DeSitter model the corresponding age is t 0ED = (2/3)tH0 = 10.5
109 years. The reason for this difference is that in the Einstein-DeSitter model
the expansion is decelerated all the time, while in the Friedmann-Lematre model
the repulsive gravitation due to the vacuum energy have made the expansion
accelerate lately (see below). Hence, for a given value of the Hubble parameter
the previous velocity was larger in the Einstein-DeSitter model than in the
Friedmann-Lematre model.
The ratio of the age of the universe and its Hubble age depends upon the
present relative density of the vacuum energy as follows,

t0 2 arc tanh 0
= H 0 t0 = (10.102)
tH0 3 0
This function is depicted graphically in fig. 10.7 The age of the universe increases

Figure 10.7: The ratio of the age of the universe and the Hubble age as function
of the present relative density of the vacuum energy.

with increasing density of vacuum energy. In the limit that the density of the
vacuum approaches the critical density, thereis no dark matter, and the universe
model approaches the DeSitter model with exponential expansion and no Big
Bang. This model behaves in the same way as the Steady State cosmological
model and is infinitely old.
A dimensioness quantity representing the rate of change of the cosmic ex-
pansion velocity is the deceleration parameter, which is defined as q = a/aH 2 .
For the present universe model the deceleration parameter as a function of time
is
1
q = [1 3 tanh2 (t/t )] (10.103)
2
10.4 Some cosmological models 153

which is shown graphically in fig. 10.8 The inflection point of time t 1 when

Figure 10.8: The deceleration parameter as function of cosmic time.

deceleration turned into acceleration is given by q = 0. This leads to



t1 = t arc tanh(1/ 3) (10.104)

or expressed in terms of the age of the universe



arc tanh(1/ 3)
t1 = t0 (10.105)
arc tanh 0

The corresponding cosmic red shift is


 1/3
a0 20
z(t1 ) = 1 = 1 (10.106)
a(t1 ) 1 0

Inserting 0 = 0.7 gives t1 = 0.54t0 and z(t1 = 0.67.


The results of analysing the observations of supernova SN 1997 at z = 1.7,
corresponding to an emission time t e = 0.30t0 = 4.5 109 years, have provided
evidence that the universe was decelerated at that time (Riess n.d.). M.Turner
and A.G.Riess (Turner and Riess 2001) have recently argued that the other
supernova data favour a transition from deceleration to acceleration for a red
shift around z = 0.5.
Note that the expansion velocity given by Hubbles law, v = Hd, always
decreases as seen from fig. 10.6. This is the velocity away from the Earth of the
cosmic fluid at a fixed physical distance d from the Earth. The quantity a on the
other hand, is the velocity of a fixed fluid particle comoving with the expansion
154 Chapter 10. Cosmology

of the universe. If such a particle accelerates, the expansion of the universe is


said to accelerate. While H tells how fast the expansion velocity changes at a
fixed distance from the Earth, the quantity a represents the acceleration of a
free particle comoving with the expanding universe. The connection between
these two quantities are a = a(H + H 2 ).
The ratio of the inflection point of time and the age of the universe, as given
in eq.(10.105), is depicted graphically as function of the present relative density
of vacuum energy in fig. 10.9 The turnover point of time happens earlier the

Figure 10.9: The ratio of the point of time when cosmic decelerations turn over
to acceleration to the age of the universe.

greater the vacuum density is. The change from deceleration to acceleration
would happen at the present time if 0 = 1/3.
The red shift of the inflection point given in eq.(10.106) as a function of
vacuum energy density, is plotted in fig. 10.10 Note that the red shift of future
points of time is negative, since then a > a 0 . If 0 < 1/3 the transition to
acceleration will happen in the future.
The critical density is

cr = tanh2 (t/t ) (10.107)

This is plotted in fig. 10.11. The critical density decreases with time.
Eq. (10.106) shows that the relative density of the vacuum energy is

= tanh2 (t/t ) (10.108)

which is plotted in fig. 10.12. The density of the vacuum energy approaches
the critical density. Since the density of the vacuum energy is constant, this is
10.4 Some cosmological models 155

Figure 10.10: The cosmic red shift of light emitted at the turnover time from
deceleration to acceleration as function of the present relative density of vacuum
energy.

Figure 10.11: The critical density in units of the constant density of the vacuum
energy as function of time.
156 Chapter 10. Cosmology

Figure 10.12: The relative density of the vacuum energy density as function of
time.

better expressed by saying that the critical density approaches the density of
the vacuum energy. Furthermore, since the total energy density is equal to the
critical density all the time, this also means that the density of matter decreases
faster than the critical density. The density of matter as function of time is

M = sinh2 (t/t ) (10.109)

which is shown graphically in fig. 10.13 The relative density of matter as func-
tion of time is
M = cosh2 (t/t ) (10.110)
which is depicted in fig. 10.14 Adding the relative densities of fig. 10.13 and
fig. 10.14 or the expressions (10.107) and (10.109) we get the total relative
density T OT = M + = 1.
The universe became vacuum dominated at a point of time t 2 when (t2 ) =
M (t2 ). From eq.(10.109) follows that this point of time is given by sinh(t 2 /t ) =
1. According to eq.(10.99) we get
arc sinh(1)
t2 = t0 (10.111)
arc tanh( 0 )
From eq.(10.97) follows that the corresponding red shift is

z(t2 ) = A1/3 1 (10.112)

Inserting 0 = 0.7 gives t2 = 0.73t0 and z(t2 ) = 0.32. The transition to


accelerated expansion happens before the universe becomes vacuum dominated.
10.4 Some cosmological models 157

Figure 10.13: The density of matter in units of the density of vacuum energy as
function of time.

Figure 10.14: The relative density of matter as function of time.


158 Chapter 10. Cosmology

Note from eqs.(10.103) and (10.108) that in the case of the flat Friedmann-
Lematre universe model, the deceleration parameter may be expressed in terms
of the relative density of vacuum only, q = (1/2)(1 3 ). The supernova Ia
observations have shown that the expansion is now accelerating. Hence if the
universe is flat, this alone means that 0 > 1/3.
As mentioned above, many different observations indicate that we live in a
universe with critical density, where cold matter contributes with about 30%
of the density and vacuum energy with about 70%. Such a universe is well
described by the Friedmann-Lematre universe model that have been presented
above.
However, this model is not quite without problems in explaining the observed
properties of the universe. In particular there is now much research directed at
solving the so called coincidence problem. As we have seen, the density of the
vacuum energy is constant during the expansion, while the density of the mat-
ter decreases inversely proportional to a volume comoving with the expanding
matter. Yet, one observes that the density of matter and the density of the
vacuum energy are of the same order of magnitude at the present time. This
seems to be a strange and unexplained coincidence in the model. Also just at
the present time the critical density is approaching the density of the vacuum
energy. At earlier times the relative density was close to zero, and now it changes
approaching the constant value 1 in the future. S. M. Carroll (Carroll 2001) has
illustrated this aspect of the coincidence problem by plotting as a function
of ln(t/t0 ). Differentiating the expression (10.108) we get
t d sinh(t/t )
= (10.113)
2 dt cosh3 (t/t )
which is plotted in fig. 10.15
Putting = 0 we find that the rate of change of was maximal at
the point of time t1 when the deceleration of the cosmic expansion turned into
acceleration. There is now a great activity in order to try to explain these coinci-
dences by introducing more general forms of vacuum energy called quintessence,
and with a density determined dynamically bythe evolution of a scalar field
(Turner 2001).
However, the simplest type of vacuum energy is the LIVE. One may hope
that a future theory of quantum gravity may settle the matter and let us un-
derstand the vacuum energy. In the meantime we can learn much about the
dynamics of a vacuum dominated universe by studying simple and beautiful
universe models such as the Friedmann-Lematre model.

10.5 Inflationary Cosmology


10.5.1 Problems with the Big Bang Models
The Horizon Problem
The Cosmic Microwave Background (CMB) radiation from two points A and B
in opposite directions has the same temperature. This means that it has been
10.5 Inflationary Cosmology 159

Figure 10.15: Rate of change of as function of ln( tt0 ). The value ln( tt0 ) = 40
corresponds to the cosmic point of time t 0 1s.

radiated by sources of the same temperature in these points. Thus, the universe
must have been in thermic equilibrium at the decoupling time, t d = 3 105 years.
This implies that points A and B, at opposite sides of the universe, had been
in causal contact already at that time. I.e., a light signal must have had time to
move from A to B during the time from t = 0 to t = 3 10 5 years. The points
A and B must have been within each others horizons at the decoupling.
Consider a photon moving radially in space descibed by the Robertson-
Walker metric (10.14) with k = 0. Light follows a null geodesic curve, i.e. the
curve is defined by ds2 = 0. We get
dt
dr = . (10.114)
a(t)
The coordinate distance the photon has moved during the time t is
Z t
dt
r = . (10.115)
0 a(t)
The physical distance the light has moved at the time t is called the horizon
distance, and is Z t
dt
lh = a(t)t = a(t) . (10.116)
0 a(t)
To find a quantitative expression for the horizon problem, we may consider
a model with critical mass density (Euclidian spacelike geometry.) Using p = w
and = 1, integration of equation (10.36) gives
2
a t 3+3w . (10.117)
160 Chapter 10. Cosmology

Inserting this into the expression for l h and integrating gives


3w + 3
lh = t. (10.118)
3w + 1
Let us call the volume inside the horizon the horizon volume and denote it by
VH . From equation (10.118) follows that V H t3 . At the decoupling time, the
horizon volume may therefore be written
 3
td
(VH )d = V0 , (10.119)
t0
where V0 is the size of the present horizon volume. Events within this volume
are causally connected, and a volume of this size may be in thermal equilibrium
at the decoupling time.
Let (V0 )d be the size, at the decoupling, of the part of the universe that
corresponds to the present horizon volume, i.e. the observable universe. For our
Euclidean universe, the equation (10.117) holds, giving
  2
a3 (td ) td w+1
(V0 )d = 3 V0 = V0 . (10.120)
a (t0 ) t0
From equations (10.119) and (10.120), we get
  3w+1
(V0 )d td w+1
= . (10.121)
(VH )d t0

Using that td = 104 t0 and inserting w = 0 for dust, we find (VV0H))dd = 104 .
Thus, there was room for 104 causally connected areas at the decoupling time
within what presently represents our observable universe. Points at opposite
sides of our observable universe were therefore not causally connected at the
decoupling, according to the Friedmann models of the universe. These models
can therefore not explain that the temperature of the radiation from such points
is the same.

The Flatness Problem



According to eq. (10.42), the total mass parameter = cr is given by

k
1= . (10.122)
a2
By using the expansion factor (10.117) for a universe near critical mass
density, we get
 2( 3w+1 )
1 t 3w+3
= . (10.123)
0 1 t0
For a radiation dominated universe, we get
1 t
= . (10.124)
0 1 t0
10.5 Inflationary Cosmology 161

Measurements indicate that 0 1 is of order of magnitude 1. The age of


the universe is about t0 = 1017 s. When we stipulate initial conditions for the
universe, it is natural to consider the Planck time, t P = 1043 s, since this is
the limit to the validity of general relativity. At earlier time, quantum effects
will be important, and one can not give a reliable description without using
quantum gravitation. The stipulated initial condition for the mass parameter
then becomes that 1 is of order 1060 at the Planck time. Such an extreme
fine tuning of the initial value of the universes mass density can not be explained
within standard Big Bang cosmology.

Other Problems
The Friedman models can not explain questions about why the universe is nearly
homogeneous and has an isotropic expansion, nor say anything about why the
universe is expanding.

10.5.2 Cosmic Inflation


Spontaneous Symmetry Breaking and the Higgs Mechanism
The particles responsible for the electroweak force, W and Z 0 are massive
(causing the weak force to only have short distance effects). This was originally
a problem for the quantum field theory describing this force, since it made it
difficult to create a renormalisable theory 1 . This was solved by Higgs and Kibble
in 1964 by introducing the so-called Higgs mechanism.
The main idea is that the massive bosons W and Z 0 are given a mass by
interacting with a Higgs field . The effect causes the mass of the particles to be
proportional to the value of the Higgs field in vacuum. It is therefore necessary
for the mechanism that the Higgs field has a value different from zero in the
vacuum (the vacuum expectation value must be non-zero).
Let us see how the Higgs field can get a non-zero vacuum expectation value.
The important thing for our purpose is that the potential for the Higgs field
may be temperature dependent. Let us assume that the potential for the Higgs
field is described by the function

1 1
V () = 2 2 + 4 , (10.125)
2 4
where the sign of 2 depends on whether the temperature is above or below a
critical temperature Tc . This sign has an important consequence for the shape
of the potential V . The potential is shown in figure 10.16 for two different
temperatures. For T > Tc , 2 > 0, and the shape is like in fig. 10.16(a), and
there is a stable minimum for = 0. However, for T < T c , 2 < 0, and the
shape is like in fig. 10.16(b). In this case the potential has stable minima for
||
= 0 =
and an unstable maximum at = 0. For both cases, the
1
The problem is that the Lagrangian for the gauge bosons can not include terms like m2 W2 ,
which are not gauge invariant
162 Chapter 10. Cosmology

potential V () is invariant under the symmetry transformation 7 (i.e.


V () = V ()).

The real vacuum state of the system is at a stable minimum of the poten-
tial. For T > Tc , the minimum is in the symmetric state = 0. On the other
hand, for T < Tc this state is unstable. It is therefore called a false vacuum.
The system will move into one of the stable minimas at = 0 . When the
system is in one of these states, it is no longer symmetric under the change of
sign of . Such a symmetry, which is not reflected in the vacuum state, is called
spontaneously broken. Note that from figure 10.16(b) we see that the energy of
the false vacuum is larger than for the real vacuum.

Figure 10.16: The shape of the potential depends on the sign of 2 .


(a): Higher temperature than the critical, with 2 > 0.
(b): Lower temperature than the critical, with 2 < 0.

The central idea, which originated the inflationary cosmology, was to take
into consideration the consequences of the unified quantum field theories, the
gauge theories, at the construction of relativistic models for the early universe.
According to the Friedmann models, the temperature was extremely high in the
early history of the universe. If one considers Higgs fields associated with GUT
models (grand unified theories), one finds a critical temperature T c correspond-
ing to the energy kTc = 1014 GeV , where k is Boltzmanns constant. Before
the universe was about t1 = 1035 s old, the temperature was larger than this.
Thus, the Higgs field was in the symmetric ground state. According to most of
the inflation models, the universe was dominated by radiation at this time.

When the temperature decreases, the Higgs potential changes. This could
happen as shown in figure 10.17. Here, there is a potential barrier at the critcal
temperature, which means that there can not be a classical phase transition.
The transition to the stable minimum must happen by quantum tunneling.
This is called a first order phase transition.
10.5 Inflationary Cosmology 163

Figure 10.17: The temperature dependence of a Higgs potential with a first


order phase transition.

Guths Inflation Model


Alan Guths original inflation model (Guth 1981) was based on a first order
phase transition.
According to most of the inflationary models, the universe was dominated
by radiation during the time before 10 35 s. The universe was then expanding so
fast that there was no causal contact between the different parts of the universe
that became our observable universe. Probably, the universe was rather homo-
geneous, with considerable spacelike variations in temperature. There was also
areas of false vacuum, with energy densities characteristic of the GUT energy
scale, which also controls its critical temperature. While the energy density
of the radiation decreased quickly, as a 4 , the energy density of vacuum was
constant. At the time t = 1035 s, the energy density of the radiation became
less than that of the vacuum.
At the same time, the potential started to change, such that the vacuum went
from being stable to being an unstable false vacuum. Thus, there was a first
order phase transition to the real vacuum. Because of the inhomogeneouty of
the universes initial condition, this happened with different speed at differing
places. The potential barrier slowed down the process, which happened by
tunneling, and the universe was at several places considerably undercooled, and
there appeared bubbles dominated by the energy of the false vacuum. These
areas acted on themselves with repulsive gravity.
By integrating the equation of motion for the expansion factor in such a
vacuum dominated bubble, one gets
r
Ht 8Gc
a=e , H= . (10.126)
3
164 Chapter 10. Cosmology

By inserting the GUT value above, we get H = 6.6 10 34 s1 , i.e. H 1 =


1.5 1035 s. With reference to field theoretical works by Sindney Coleman and
others, Guth argumented that a realistic duration of the nucleation process hap-
pening during the phase transition is 10 33 s. During this time, the expansion
factor increases by a factor of 1028 . This vacuum dominated epoch is called the
inflation era.
Let us look closer at what happens with the energy of the universe in the
course of its development, according to the inflationary models. To understand
this we first have to consider what happens at the end of the inflationary era.
When the Higgs field reaches the minimum corresponding to the real vacuum,
it starts to oscillate. According to the quantum description of the oscillating
field, the energy of the false vacuum is converted into radiation and particles. In
this way the equation of state for the energy dominating the development of the
expansion factor changes from p = , characteristic for vacuum, to p = 13 ,
characteristic of radiation.
The energy density and the temperature of the radiation is then increased
enourmously. Before and after this short period around the time t = 10 33 s the
radiation energy increases adiabatically, such that a 4 = constant. According to
Stefan-Boltzmanns law of radiadion, T 4 . Therefore, aT = constant during
adiabatic expansion. This means that during the inflationary era, while the
expansion factor increases exponentially, the energy density and temperature
of radiation decreases exponentially. At the end of the inflationary era, the
radiation is reheated so that it returns to the energy it had when the inflationary
era started.
It may be interesting to note that the Newtonian theory of gravitation does
not allow an inflationary era, since stress has no gravitational effect according
to it.

The Inflation Models Answers to the Problems of the Friedmann


Models

The horizon problem will here be investigated in the light of this model. The
problem was that there was room for about 10000 causally connected areas
inside the area spanned by our presently observable universe at the time. Let us
calculate the horizon radius lh and the radius a of the region presently within
the horizon, lh = 15 109 ly = 1.5 1026 cm, at the time t1 = 1035 s when the
inflation started. From equation (10.118) for the radiation dominated period
before the inflatinary era, one gets

lh = 2t1 = 6 1025 cm. (10.127)

The radius, at time t1 , of the region corresponding to our observable universe,


is found by using that a eHt during the inflation era from t1 = 1035 s to
1
t2 = 1033 s, a t 2 in the radiation dominated period from t 2 to t3 = 1011 s,
2
and a t 3 in the matter dominated period from t 3 until now, t0 = 1017 s. This
10.5 Inflationary Cosmology 165

gives
 1  2
eHt1 t2 2 t3 3
a1 = Ht2 lh (t0 ) = 1.5 1028 cm. (10.128)
e t3 t0
We see that at the beginning of the inflationary era the horizon radius,
lh , was larger than the radius a of the region corresponding to our observable
universe. The whole of this region was then causally connected, and thermic
equilibrium was established. This equilibrium has been kept since then, and
explains the observed isotropy of the cosmic background radiation.
We will now consider the flatness problem. This problem was the necessity,
in the Friedmann models, of fine tuning the initial density in order to obtain
the closeness of the observed mass density to the critical density. Again, the
inflationary models give another result. Inserting the expansion factor (10.126)
into equation (10.122), we get

k 2Ht
1= e , (10.129)
H2
where H is constant and given in eq. (10.126). The ratio between 1 at the
end of and the beginning of the inflationary era becomes
2 1
= e2H(t2 t1 ) = 1056 . (10.130)
1 1
Contrary to in the Friedmann models, where the mass density moves away
from the critical density as time is increasing, the density approaches the critical
density exponentially during the inflationary era. Within a large range of initial
conditions, this means that according to the inflation models the universe should
still have almost critical mass density.
Bibliography

Bernadis et.al, P. d.: 2001, Multiple peaks in the angular power spectrum of
the cosmic microwave background: Significance and consequences for cos-
mology, astro-ph/0105296.
Brill, D. R. and Cohen, J. M.: 1966, Phys. Rev. 143, 1011.
Carroll, S. M.: 1998, Quintessence and the rest of the world., Phys. Rev. Lett.
81, 30673070.
Carroll, S. M.: 2001, Dark energy and the preposterous universe., astro-
ph/0107571.
Grn, .: 1986, Repulsive gravitation and inflationary universe models., 54, 46
52.
Grn, . and Fl, H. K.: 1984, Differensialgeometri, Departmental. Course
compendium.
Guth, A. H.: 1981, The inflationary universe: a possible solution to the horizon
and flatness problems, Phys. Rev. D23, 347356.
Linde, A.: 2001, Inflation and string cosmology, astro-ph/0107176.
McGaugh, S.: 2001, Constraints on the radial mass distribution of dark matter
halos from rotation curves, astro.ph/0107490.
Misner, C. W., Thorne, K. S. and Wheeler, J. A.: 1973, Gravitation, first edn,
W. H. Freeman and company, San Francisco.
Perlmutter et al., S.: 1999, Measurements of omega ans lambda from 42 high-
redshift supernovae., Astrophys. J. 517, 565586.
Pound, R. V. and Rebka, G. A., J.: 1960, Phys. Review Letters 4, 337.
Pryke et al., C.: 2001, Cosmological parameter extraction from the first season
of observations with dasi., astro-ph/0104490.
Ravndal, F.: 1978, Infring i generell relativitetsteori, Departmental. Course
compendium.
Riess, A. G.: n.d., The farthest known supernova: Support for an accelerating
universe and a glimpse of the epoch of deceleration., astro-ph/0104455.

166
BIBLIOGRAPHY 167

Riess et al., A. G.: 1998, Observational evidence from supernovae for an accel-
erating universe and a cosmological constant., Astron. J. 116, 10091038.

Stompor et al., R.: 2001, Cosmological implications of the maxima-i high


resolution cosmic microwave background anisotropy measurement., astro-
ph/0105062.

Turner, M. S.: 2001, Dark energy and the new cosmology., astro-ph/0108103.

Turner, M. S. and Riess, A. G.: 2001, Do Sn Ia provide direct evidence for past
deceleration of the universe?, astro-ph/0106051.

Zeldovich, Y.: 1968, The cosmological constant and the theory of elementary
particles, Sov. Phys. Usp. 11, 381393.

Zlatev, I., Wang, L. and Steinhardt, P. J.: 1999, Quintessence, cosmic coinci-
dence, and the cosmological constant, Phys. Rev. Lett. 82, 896899.
Index

Symbols mini . . . . . . . . . . . . . . . . . . . . . . . 122


, 1 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . .50 radiation. . . . . . . . . . . . . . . . . . .121
S = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . 59 temperature . . . . . . . . . . . . . . . 121
(~r ) (Dirac delta function) . . . . . . . 4 Born-stiff motion . . . . . . . . . . . . . . . . 44
ij . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
(tensor product) . . . . . . . . . . . . . . 26 C
d2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 canonical momentum . . . . . . . . . . . . 59
d (solid angle) . . . . . . . . . . . . . . . . . . 5 Cartan
dt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 formalism . . . . . . . . . . . . . . . . . . 100
dl2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 structure equations
[ ] (antisymmetric combination) . 32 1st . . . . . . . . . . . . . . . . . . . 79, 100
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 2nd . . . . . . . . . . . . . . . . . . 85, 101
3-space. . . . . . . . . . . . . . . . . . . . . . . . . .34 surface curvature. . . . . . . . . . . .89
4-acceleration . . . . . . . . . . . . . . 16, 120 centrifugal acceleration . . . . . . . . . . 57
4-momentum . . . . . . . . . . . . . . . . . . . . 15 centrifulgal barrier . . . . . . . . . . . . . 112
4-velocity . . . . . . . . . . . . . . . . . . . . . . . 15 Christoffel symbols . . . . . . . . . . . . . . 61
identity . . . . . . . . . . . . . . . . . 16, 59 in coordinate basis . . . . . . . . . . 53
physical interpretation . . . . . 103
A
rotating refernece frame . . . . . 74
accelerated frame
comoving orthonormal basis . . . . . 31
hyperbolically . . . . . . . . . . . . . . . 44
connection coefficients
acceleration of particle
in a Riemannian space . . . . . . 74
in non-rotating frame . . . . . . . 56
in coordinate basis . . . . . . . . . . 53
in rotating frame . . . . . . . . . . . . 57
rotating reference frame . . . . . 73
acceleration scalar . . . . . . . . . . . . . 120
connection forms . . . . . . . . . . . . . . . 100
action . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
conservation
Archimedean spiral. . . . . . . . . . . . . .41
energy-momentum . . . . . . . . . . 93
B constant of motion . . . . . . . . . . . . . . 59
basis continuity equation. . . . . . . . . . . . . .93
1-form . . . . . . . . . . . . . . . . . . . . . . 25 contravariant tensor . . . . . . . . . . . . . 26
coordinate basis vector . . 18, 21 convective derivative . . . . . . . . . . . . 93
orthonormal . . . . . . . . . . . . . 18, 22 coordinate
Bianchi cyclic . . . . . . . . . . . . . . . . . . . 59, 60
1st identity . . . . . . . . . . . . . . 90, 92 Gaussian. . . . . . . . . . . . . . . . . . . . 54
component form . . . . . . . . . . 91 transformation . . . . . . . . . . 20, 27
2nd identity . . . . . . . . . . . . . 91, 95 coordinate basis . . . . . . . . . . . . . 18, 21
black hole connection coefficients. . . . . . .53
horizon . . . . . . . . . . . . . . . . . . . . 120 coordinate system . . . . . . . . . . . . . . . 18

168
INDEX 169

coriolis acceleration . . . . . . . . . . . . . 57 Einstein


cosmological redshift . . . . . . . . . . . 135 field equations . . . . . . . . . . . . . . 96
covariance principle. . . . . . . . . .11, 51 principle of equivalence. . . . . . . 9
covariant equations. . . . . . . . . . . . . .28 synchronization . . . . . . . . . . . . . 35
covariant tensor . . . . . . . . . . . . . . . . . 26 tensor. . . . . . . . . . . . . . . . . . . . . . .96
curvature . . . . . . . . . . . . . . . . . . . . . . . 81 Einstein deSitter universe . . . . . . 144
forms . . . . . . . . . . . . . . . . . . 85, 101 embedding
of surface . . . . . . . . . . . . . . . . . 89 of the Schwarzschild metric 108
Gaussian . . . . . . . . . . . . . . . . 88, 89 energy-momentum tensor. . . . . . . .93
geodesic . . . . . . . . . . . . . . . . . . . . 87 equivalence, principle of . . . . . . . . . . 9
normal . . . . . . . . . . . . . . . . . . . . . . 87 Euler equation of motion . . . . . . . . 93
principal . . . . . . . . . . . . . . . . . . . . 88 Euler-Lagrange equations . . . . . . . 59
Ricci . . . . . . . . . . . . . . . . . . . . . . . . 91 exterior differentiation. . . . . . . . . . .49
Riemann tensor . . . . . . . . . . . . . 83
surface (Cartan formalism) . . 89 F
field equations
D Einstein . . . . . . . . . . . . . . . . . . . . 96
deflection of light . . . . . . . . . . . . . . 117 vacuum . . . . . . . . . . . . . . . . . . . . . 97
differential forms first order phase transition . . . . . 160
1-form . . . . . . . . . . . . . . . . . . . . . . 25 flatness problem . . . . . . . . . . . 158, 163
2-form . . . . . . . . . . . . . . . . . . . . . . 32 forms . . . . . . . . . see differential forms
components . . . . . . . . . . . . . . . . . 33 free fall . . . . . . . . . . . . . . . . . . . . . . . . 104
connection . . . . . . . . . . . . . . . . . 100
covariant differentiation of . . 76 G
curvature . . . . . . . . . . . . . . 85, 101 Gauss
of surface . . . . . . . . . . . . . . . . . 89 coordinates . . . . . . . . . . . . . . . . . 54
differentiation of . . . . . . . . . . . . 49 curvature. . . . . . . . . . . . . . . .88, 89
directional derivatives of . . . . 76 integral theorem . . . . . . . . . . . . . 4
null form. . . . . . . . . . . . . . . . . . . .33 theorema egregium . . . . . . . . . . 89
p-form . . . . . . . . . . . . . . . . . . . . . . 32 general principle of relativity . . . . 10
transformation . . . . . . . . . . . . . . 27 geodesic
differentiation curvature . . . . . . . . . . . . . . . . . . . 87
convective . . . . . . . . . . . . . . . . . . 93 curve . . . . . . . . . . . . . . . . . . . . 57, 61
covariant . . . . . . . . . . . . . . . . 53, 55 spatial . . . . . . . . . . . . . . . . . . . . 65
exterior . . . . . . . . . . . . . . . . . . . . . 49 equation . . . . . . . . . . . . . . . . . . . . 56
of form . . . . . . . . . . . . . . . . . . . . . 49 postulate . . . . . . . . . . . . . . . . . . . 98
Dirac delta function . . . . . . . . . . . . . . 4 gravitational acceleration. . . . . . .103
distance Tolman-Whittaker expression128
coordinate . . . . . . . . . . . . . . . . . 100 gravitational collapse. . . . . . . . . . .112
physical. . . . . . . . . . . . . . . . . . . .100 gravitational time dilation . . . . . . . 40
Doppler effect . . . . . . . . . . . . . . . . . . . 23 Grn, yvind.. . . . . . . . . . . . . . . . . . . .3
Doppler effect, gravitational . . . . . 71 Guth, Alan . . . . . . . . . . . . . . . . . . . . 161
dust dominated universe . . . . . . . 142
H
E Hafele-Keating experiment . . . . . 114
Eddington - Finkelstein coordinates Hamilton-principle . . . . . . . . . . . . . . 58
104 Hawking radiation . . . . . . . . . . . . . 121
170 INDEX

heat capacity of black hole . . . . . 121 from Schwarzschild solution 103


Higgs mechanism . . . . . . . . . . . . . . 159 mixed tensor . . . . . . . . . . . . . . . . . . . . 26
horizon problem . . . . . . . . . . . 156, 162 momentum-energy tensor. . . . . . . .93
Hubbles law. . . . . . . . . . . . . . . . . . . .135
hydrostatic equilibrium . . . . . . . . 127 N
hyperbolically accelerated frame . 44 Newton
fluid . . . . . . . . . . . . . . . . . . . . . . . . 94
I hydrodynamics. . . . . . . . . . . . . . 93
incompressible stars . . . . . . . . . . . . 130 law of gravitation . . . . . . . . . . . . 1
inertial reference frame . . . . . . . . . . 18 continuous distribution . . . . . 2
inertial dragging . . . . . . . . . . . . . . . 123 local form . . . . . . . . . . . . . . . . . . 2
inertial reference frame . . . . . . . . . . 54 Newtonian incompressible star . 126
inflation . . . . . . . . . . . . . . . . . . . . . . . 161
O
K orbit equation. . . . . . . . . . . . . . . . . .113
Kerr metric . . . . . . . . . . . . . . . . . . . . 122 orthonormal basis . . . . . . . . . . . . . . . 18
Kretschmanns curvature scalar 104
Kruskal - Szekers. . . . . . . . . . . . . . .104 P
parallel transport . . . . . . . . . . . . 55, 81
L particle trajectory
Lagrange Schwarzschild 3-space . . . . . . 110
covariant dynamics. . . . . . . . . .57 Poincars lemma . . . . . . . . . . . . . . . 51
equations . . . . . . . . . . . . . . . . . . . 59 precession
free particle . . . . . . . . . . . . . . . 61 of Mercurys perihelion . . . . 115
function . . . . . . . . . . . . . . . . . . . . 58 pressure forces . . . . . . . . . . . . . . . . . 127
free particles . . . . . . . . . . . . . . 61 principle of equivalence . . . . . . 9, 103
Laplace equation . . . . . . . . . . . . . . . . . 3 projectile motion . . . . . . . . . . . . . . . . 68
Levi-Civita connection . . . . . . . . . . 53 proper distance . . . . . . . . . . . . . . . . 135
light
deceleration of . . . . . . . . . . . . . 108 R
deflection of . . . . . . . . . . . . . . . 117 radiation dominated universe. . .141
light cones . . . . . . . . . . . . . . . . . . . . . 106 rank . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
line element . . . . . . . . . . . . . . . . . . . . . 31 rapidity . . . . . . . . . . . . . . . . . . . . . . . . . 45
free particle . . . . . . . . . . . . . . . . . 60 redshift
in space . . . . . . . . . . . . . . . . . . . . . 34 cosmological . . . . . . . . . . . . . . . 135
spacetime . . . . . . . . . . . . . . . . . . . 35 Kerr spacetime . . . . . . . . . . . . 123
Lorentz transformation . . . . . . . . . . 23 reference frame . . . . . . . . . . . . . . . . . . 17
inertial . . . . . . . . . . . . . . . . . . 18, 54
M relativity, general principle of. . . .10
Machs principle . . . . . . . . . . . . . . . . 12 Ricci
metric curvature . . . . . . . . . . . . . . . . . . . 91
covariant differentiation . . . . . 77 tensor, divergence . . . . . . . . . 96
Kerr (rotating) . . . . . . . . . . . . 122 identity . . . . . . . . . . . . . . . . . . . . . 90
static,stationary . . . . . . . . . . . . 60 component form . . . . . . . . . . 90
tensor. . . . . . . . . . . . . . . . . . . . . . .28 Riemann curvature tensor . . 83, 101
mini black holes. . . . . . . . . . . . . . . . 122 components . . . . . . . . . . . . . . . . . 84
Minkowski space-time rotating black hole . . . . . . . . . . . . . 122
INDEX 171

rotating reference frame . . . . . . . . . 65 product . . . . . . . . . . . . . . . . . . . . . 26


connection coefficients. . . . . . .73 rank . . . . . . . . . . . . . . . . . . . . . . . . 26
Ricci curvature. . . . . . . . . . . . . .91
S Riemann curvature . . . . . 83, 101
Sagnac effect . . . . . . . . . . . . . . . . . . . . 42 transformation properties . . . 27
Schwarzschild theorema egregium . . . . . . . . . . . . . . 89
3-space . . . . . . . . . . . . . . . . . . . . 110 tidal forces . . . . . . . . . . . . . . . . . . . . . . . 5
energy . . . . . . . . . . . . . . . . . . . 110 time dilation, gravitational . . . . . . 40
momentum . . . . . . . . . . . . . . 110 Tolman-Oppenheimer-Volkov eq.130
light cones . . . . . . . . . . . . . . . . . 106 Tolman-Whittaker expression . . 128
metric . . . . . . . . . . . . . . . . . . . . . 104 travelling time . . . . . . . . . . . . . . . . . 105
embedding . . . . . . . . . . . . . . . 108 Twin paradox . . . . . . . . . . . . . . . . . 69
radius . . . . . . . . . . . . . . 2, 104, 120
solution U
interior . . . . . . . . . . . . . . . . . . 130 universe model
outer . . . . . . . . . . . . . . . . . . . . 100 dust dominated . . . . . . . . . . . . 142
spacetime . . . . . . . . . . . . . 104, 120 inflationary . . . . . . . . . . . . . . . . 161
simultaneity radiation dominated . . . . . . . 141
on Einstein synchronized clocks V
35 vacuum field equations . . . . . . . . . . 97
space, geometry of . . . . . . . . . . 35 vector
singularity basis . . . . . . . . . . . . . . . . . . . . . . . . 14
coordinate . . . . . . . . . . . . . . . . . 104 covariant differentiation of . . 75
physical. . . . . . . . . . . . . . . . . . . .104 dimension . . . . . . . . . . . . . . . . . . . 13
solid angle . . . . . . . . . . . . . . . . . . . . . . . 4 linearly independent . . . . . . . . 13
spherical geometry . . . . . . . . . . . . . . 38
spontaneous symmetry breaking160 Z
structure coefficients . . . . . . . . . . . . 23 ZAMO . . . . . . . . . . . . . . . . . . . . . . . . . 123
in a Riemannian space . . . . . . 74
in coordinate basis . . . . . . . . . . 25
surface curvature . . . . . . . . . . . . . . . . 89
surface gravity . . . . . . . . . . . . . . . . . 120

T
tangent plane . . . . . . . . . . . . . . . . . . . 17
tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
antisymmetric . . . . . . . . . . . . . . 31
construction example . . . . . . . 27
contraction of . . . . . . . . . . . . . . . 91
contravariant . . . . . . . . . . . . . . . 26
covariant. . . . . . . . . . . . . . . . . . . .26
covariant differentiation of . . 77
Einstein . . . . . . . . . . . . . . . . . . . . 96
metric . . . . . . . . . . . . . . . . . . . . . . 28
mixed . . . . . . . . . . . . . . . . . . . 26, 28
momentum energy . . . . . . . . . . 93

Você também pode gostar