186
Lecture Notes On General Relativity Øyvind Grøn 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

LectureNotes On GeneralRelativity

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: LectureNotes On GeneralRelativity

Lecture NotesOn

General Relativity

Øyvind Grøn

Oslo College, Department of engineering, Cort Adelers gt. 30, N-0254 Oslo,Norwayand

Department of Physics, University of Oslo, Box 1048 Blindern, N-0316, Norway

Page 2: LectureNotes On GeneralRelativity

May 8, 2006

Page 3: LectureNotes On GeneralRelativity

Preface

These notes are a transcript of lectures delivered by Øyvind Grøn during thespring of 1997 at the University of Oslo. Two compendia, (Grøn and Flø 1984)and (Ravndal 1978) were provided by Grøn as additional reference materialduring the lectures.

The present version of this document is an extended and corrected version ofa set of Lecture Notes which were typesetted by S. Bard, Andreas O. Jaunsen,Frode Hansen and Ragnvald J. Irgens using LATEX2ε. Svend E. Hjelmeland hasmade 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 allinterested readers and accept all feedback with thanks.

All comment may be sent to the author either by e-mail or snail mail.Øyvind GrønFysisk InstituttUniversitetet i OsloP.O.Boks 1048, Blindern0315 OSLOE-mail: [email protected]

Page 4: LectureNotes On GeneralRelativity
Page 5: LectureNotes On GeneralRelativity

Contents

List of Figures v

List of Definitions ix

List of Examples xi

1 Newton’s law of universal gravitation 11.1 The force law of gravitation . . . . . . . . . . . . . . . . . . . . . 11.2 Newton’s law of gravitation in its local form . . . . . . . . . . . . 21.3 Tidal Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.4 The Principle of Equivalence . . . . . . . . . . . . . . . . . . . . 91.5 The general principle of relativity . . . . . . . . . . . . . . . . . . 101.6 The covariance principle . . . . . . . . . . . . . . . . . . . . . . . 111.7 Mach’s principle . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Vectors, Tensors and Forms 132.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.1.1 4-vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142.1.2 Tangent vector fields and coordinate vectors . . . . . . . . 172.1.3 Coordinate transformations . . . . . . . . . . . . . . . . . 202.1.4 Structure coefficients . . . . . . . . . . . . . . . . . . . . . 23

2.2 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.2.1 Transformation of tensor components . . . . . . . . . . . . 272.2.2 Transformation of basis 1-forms . . . . . . . . . . . . . . . 272.2.3 The metric tensor . . . . . . . . . . . . . . . . . . . . . . 28

2.3 Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3 Accelerated Reference Frames 343.1 Rotating reference frames . . . . . . . . . . . . . . . . . . . . . . 34

3.1.1 Space geometry . . . . . . . . . . . . . . . . . . . . . . . . 343.1.2 Angular acceleration in the rotating frame . . . . . . . . . 383.1.3 Gravitational time dilation . . . . . . . . . . . . . . . . . 403.1.4 Path of photons emitted from axes in the rotating refer-

ence frame (RF) . . . . . . . . . . . . . . . . . . . . . . . 413.1.5 The Sagnac effect . . . . . . . . . . . . . . . . . . . . . . . 42

3.2 Hyperbolically accelerated reference frames . . . . . . . . . . . . 43

i

Page 6: LectureNotes On GeneralRelativity

4 Covariant Differentiation 494.1 Differentiation of forms . . . . . . . . . . . . . . . . . . . . . . . . 49

4.1.1 Exterior differentiation . . . . . . . . . . . . . . . . . . . . 494.1.2 Covariant derivative . . . . . . . . . . . . . . . . . . . . . 51

4.2 The Christoffel Symbols . . . . . . . . . . . . . . . . . . . . . . . 534.3 Geodesic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564.4 The covariant Euler-Lagrange equations . . . . . . . . . . . . . . 574.5 Application of the Lagrangian formalism to free particles . . . . . 59

4.5.1 Equation of motion from Lagrange’s equation . . . . . . . 604.5.2 Geodesic world lines in spacetime . . . . . . . . . . . . . . 614.5.3 Gravitational Doppler effect . . . . . . . . . . . . . . . . . 70

4.6 The Koszul connection . . . . . . . . . . . . . . . . . . . . . . . . 714.7 Connection coefficients Γαµν and structure coefficients cαµν in ... . 744.8 Covariant differentiation of vectors, forms and tensors . . . . . . 75

4.8.1 Covariant differentiation of a vector in an arbitrary basis . 754.8.2 Covariant differentiation of forms . . . . . . . . . . . . . . 754.8.3 Generalization for tensors of higher rank . . . . . . . . . . 77

4.9 The Cartan connection . . . . . . . . . . . . . . . . . . . . . . . . 77

5 Curvature 815.1 The Riemann curvature tensor . . . . . . . . . . . . . . . . . . . 815.2 Differential geometry of surfaces . . . . . . . . . . . . . . . . . . . 86

5.2.1 Surface curvature, using the Cartan formalism . . . . . . . 895.3 The Ricci identity . . . . . . . . . . . . . . . . . . . . . . . . . . 905.4 Bianchi’s 1st identity . . . . . . . . . . . . . . . . . . . . . . . . . 905.5 Bianchi’s 2nd identity . . . . . . . . . . . . . . . . . . . . . . . . 91

6 Einstein’s Field Equations 936.1 Energy-momentum conservation . . . . . . . . . . . . . . . . . . . 93

6.1.1 Newtonian fluid . . . . . . . . . . . . . . . . . . . . . . . . 936.1.2 Perfect fluids . . . . . . . . . . . . . . . . . . . . . . . . . 95

6.2 Einstein’s curvature tensor . . . . . . . . . . . . . . . . . . . . . . 956.3 Einstein’s field equations . . . . . . . . . . . . . . . . . . . . . . . 966.4 The “geodesic postulate” as a consequence of the field equations . 98

7 The Schwarzschild spacetime 1007.1 Schwarzschild’s exterior solution . . . . . . . . . . . . . . . . . . . 1007.2 Radial free fall in Schwarzschild spacetime . . . . . . . . . . . . . 1047.3 Light cones in Schwarzschild spacetime . . . . . . . . . . . . . . . 1067.4 Analytical extension of the Schwarzschild spacetime . . . . . . . . 1087.5 Embedding of the Schwarzschild metric . . . . . . . . . . . . . . . 1107.6 Deceleration of light . . . . . . . . . . . . . . . . . . . . . . . . . 1107.7 Particle trajectories in Schwarzschild 3-space . . . . . . . . . . . 112

7.7.1 Motion in the equatorial plane . . . . . . . . . . . . . . . 1147.8 Classical tests of Einstein’s general theory of relativity . . . . . . 116

7.8.1 The Hafele-Keating experiment . . . . . . . . . . . . . . . 116

ii

Page 7: LectureNotes On GeneralRelativity

7.8.2 Mercury’s perihelion precession . . . . . . . . . . . . . . . 1187.8.3 Deflection of light . . . . . . . . . . . . . . . . . . . . . . . 119

8 Black Holes 1218.1 ’Surface gravity’:gravitational acceleration on the horizon of a

black hole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1218.2 Hawking radiation:radiation from a black hole (1973) . . . . . . . 1228.3 Rotating Black Holes: The Kerr metric . . . . . . . . . . . . . . . 123

8.3.1 Zero-angular-momentum-observers (ZAMO’s) . . . . . . . 1248.3.2 Does the Kerr space have a horizon? . . . . . . . . . . . . 125

9 Schwarzschild’s Interior Solution 1279.1 Newtonian incompressible star . . . . . . . . . . . . . . . . . . . 1279.2 The pressure contribution to the gravitational mass of a static,

spherical symmetric system . . . . . . . . . . . . . . . . . . . . . 1299.3 The Tolman-Oppenheimer-Volkov equation . . . . . . . . . . . . 1309.4 An exact solution for incompressible stars - Schwarzschild’s inte-

rior solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

10 Cosmology 13410.1 Comoving coordinate system . . . . . . . . . . . . . . . . . . . . 13410.2 Curvature isotropy - the Robertson-Walker metric . . . . . . . . 13510.3 Cosmic dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

10.3.1 Hubbles law . . . . . . . . . . . . . . . . . . . . . . . . . . 13610.3.2 Cosmological redshift of light . . . . . . . . . . . . . . . . 13610.3.3 Cosmic fluids . . . . . . . . . . . . . . . . . . . . . . . . . 13810.3.4 Isotropic and homogeneous universe models . . . . . . . . 139

10.4 Some cosmological models . . . . . . . . . . . . . . . . . . . . . . 14210.4.1 Radiation dominated model . . . . . . . . . . . . . . . . . 14210.4.2 Dust dominated model . . . . . . . . . . . . . . . . . . . . 14310.4.3 Friedmann-Lemaître model . . . . . . . . . . . . . . . . . 147

10.5 Inflationary Cosmology . . . . . . . . . . . . . . . . . . . . . . . . 15710.5.1 Problems with the Big Bang Models . . . . . . . . . . . . 15710.5.2 Cosmic Inflation . . . . . . . . . . . . . . . . . . . . . . . 160

Bibliography 165

iii

Page 8: LectureNotes On GeneralRelativity
Page 9: LectureNotes On GeneralRelativity

List of Figures

1.1 Newton’s law of universal gravitation . . . . . . . . . . . . . . . . 11.2 Newton’s law of gravitation in its local form . . . . . . . . . . . . 21.3 The definition of solid angle dΩ . . . . . . . . . . . . . . . . . . . 51.4 Tidal Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61.5 A small Cartesian coordinate system at a distance R from a mass

M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.6 An elastic, circular ring falling freely in the Earth’s gravitational

field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 Closed polygon (linearly dependent) . . . . . . . . . . . . . . . . 132.2 Carriage at rest (top) and with velocity ~v (bottom) . . . . . . . . 142.3 World-lines in a Minkowski diagram . . . . . . . . . . . . . . . . 162.4 No position vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 172.5 Tangentplane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172.6 Proper time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202.7 Coordinate transformation, flat space. . . . . . . . . . . . . . . . 222.8 Basis-vectors ~e1 and ~e2 . . . . . . . . . . . . . . . . . . . . . . . . 292.9 The covariant- and contravariant components of a vector . . . . . 30

3.1 Simultaneity in rotating frames . . . . . . . . . . . . . . . . . . . 363.2 Rotating system: Distance between points on the circumference . 363.3 Rotating system: Discontinuity in simultaneity . . . . . . . . . . 373.4 Rotating system: Angular acceleration . . . . . . . . . . . . . . . 393.5 Rotating system: Distance increase . . . . . . . . . . . . . . . . . 393.6 Rotating system: Lorentz contraction . . . . . . . . . . . . . . . . 403.7 The Sagnac effect . . . . . . . . . . . . . . . . . . . . . . . . . . . 423.8 Hyperbolic acceleration . . . . . . . . . . . . . . . . . . . . . . . 443.9 Simultaneity and hyperbolic acceleration . . . . . . . . . . . . . . 463.10 The hyperbolically accelerated reference system . . . . . . . . . . 48

4.1 Parallel transport . . . . . . . . . . . . . . . . . . . . . . . . . . . 554.2 Different world-lines connecting P1 and P2 in a Minkowski diagram 574.3 Geodesic on a flat surface . . . . . . . . . . . . . . . . . . . . . . 594.4 Geodesic on a sphere . . . . . . . . . . . . . . . . . . . . . . . . . 594.5 Timelike geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . 614.6 Projectiles in 3-space . . . . . . . . . . . . . . . . . . . . . . . . . 634.7 Geodesics in rotating reference frames . . . . . . . . . . . . . . . 64

v

Page 10: LectureNotes On GeneralRelativity

4.8 Coordinates on a rotating disc . . . . . . . . . . . . . . . . . . . . 654.9 Projectiles in accelerated frames . . . . . . . . . . . . . . . . . . . 664.10 The twin “paradox” . . . . . . . . . . . . . . . . . . . . . . . . . . 684.11 Rotating coordinate system . . . . . . . . . . . . . . . . . . . . . 72

5.1 Parallel transport of ~A . . . . . . . . . . . . . . . . . . . . . . . . 815.2 Parallel transport of a vector along a triangle of angles 90 is

rotated 90 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825.3 Geometry of parallel transport . . . . . . . . . . . . . . . . . . . 835.4 Surface geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

7.1 Light cones in Schwarzschild spacetime . . . . . . . . . . . . . . . 1077.2 Light cones in Schwarzschild spacetime . . . . . . . . . . . . . . . 1077.3 Embedding of the Schwarzschild metric . . . . . . . . . . . . . . . 1117.4 Deceleration of light . . . . . . . . . . . . . . . . . . . . . . . . . 1117.5 Newtonian centrifugal barrier . . . . . . . . . . . . . . . . . . . . 1157.6 Gravitational collapse . . . . . . . . . . . . . . . . . . . . . . . . 1167.7 Deflection of light . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

8.1 Static border and horizon of a Kerr black hole . . . . . . . . . . . 126

9.1 Hydrostatic equilibrium . . . . . . . . . . . . . . . . . . . . . . . 128

10.1 Schematic representation of cosmological redshift . . . . . . . . . 13710.2 Expansion of a radiation dominated universe . . . . . . . . . . . 14310.3 The size of the universe . . . . . . . . . . . . . . . . . . . . . . . 14610.4 Expansion factor . . . . . . . . . . . . . . . . . . . . . . . . . . . 14710.5 The expansion factor as function of cosmic time in units of the

age of the universe. . . . . . . . . . . . . . . . . . . . . . . . . . . 15010.6 The Hubble parameter as function of cosmic time. . . . . . . . . 15010.7 ....... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15110.8 The deceleration parameter as function of cosmic time. . . . . . . 15210.9 The ratio of the point of time when cosmic decelerations turn

over to acceleration to the age of the universe. . . . . . . . . . . . 15310.10The cosmic red shift of light emitted at the turnover time from

deceleration to acceleration as function of the present relativedensity of vacuum energy. . . . . . . . . . . . . . . . . . . . . . . 154

10.11The critical density in units of the constant density of the vacuumenergy as function of time. . . . . . . . . . . . . . . . . . . . . . . 154

10.12The relative density of the vacuum energy density as function oftime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

10.13The density of matter in units of the density of vacuum energyas function of time. . . . . . . . . . . . . . . . . . . . . . . . . . . 156

10.14The relative density of matter as function of time. . . . . . . . . 15610.15Rate of change of ΩΛ as function of ln( t

t0). The value ln( t

t0) =

−40 corresponds to the cosmic point of time t0 ∼ 1s. . . . . . . . 15810.16The shape of the potential depends on the sign of µ2. . . . . . . . 161

vi

Page 11: LectureNotes On GeneralRelativity

10.17The temperature dependence of a Higgs potential with a firstorder phase transition. . . . . . . . . . . . . . . . . . . . . . . . . 162

vii

Page 12: LectureNotes On GeneralRelativity
Page 13: LectureNotes On GeneralRelativity

List of Definitions

1.2.1 Solid angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42.1.1 4-velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.1.2 4-momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.1.3 4-acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162.1.4 Reference frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172.1.5 Coordinate system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.1.6 Comoving coordinate system . . . . . . . . . . . . . . . . . . . . . . 182.1.7 Orthonormal basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.1.8 Coordinate basis vectors. . . . . . . . . . . . . . . . . . . . . . . . . 182.1.9 Coordinate basis vectors. . . . . . . . . . . . . . . . . . . . . . . . . 212.1.10Orthonormal basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.1.11Commutators between vectors . . . . . . . . . . . . . . . . . . . . . . 232.1.12Structure coefficients cρµν . . . . . . . . . . . . . . . . . . . . . . . . 242.2.1 Multilinear function, tensors . . . . . . . . . . . . . . . . . . . . . . . 262.2.2 Tensor product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262.2.3 The metric tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282.2.4 Contravariant components . . . . . . . . . . . . . . . . . . . . . . . . 292.3.1 p-form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323.2.1 Born-stiff motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444.3.1 Geodesic curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564.6.1 Koszul’s connection coeffecients in an arbitrary basis . . . . . . . . . 724.8.1 Covariant derivative of a vector . . . . . . . . . . . . . . . . . . . . . 754.8.2 Covariant directional derivative of a one-form field . . . . . . . . . . 754.8.3 Covariant derivative of a one-form . . . . . . . . . . . . . . . . . . . 764.8.4 Covariant derivative of a tensor . . . . . . . . . . . . . . . . . . . . . 774.9.1 Exterior derivative of a basis vector . . . . . . . . . . . . . . . . . . . 774.9.2 Connection forms Ων

µ . . . . . . . . . . . . . . . . . . . . . . . . . . 784.9.3 Scalar product between vector and 1-form . . . . . . . . . . . . . . . 785.5.1 Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917.1.1 Physical singularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1047.1.2 Coordinate singularity . . . . . . . . . . . . . . . . . . . . . . . . . . 1048.3.1 Horizon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

ix

Page 14: LectureNotes On GeneralRelativity
Page 15: LectureNotes On GeneralRelativity

List of Examples

2.1.1 Photon clock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142.1.2 Coordinate transformation . . . . . . . . . . . . . . . . . . . . . . . . 212.1.3 Relativistic Doppler Effect . . . . . . . . . . . . . . . . . . . . . . . . 232.1.4 Structure coefficients in planar polar coordinates . . . . . . . . . . . 252.2.1 Example of a tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 272.2.2 A mixed tensor of rank 3 . . . . . . . . . . . . . . . . . . . . . . . . 282.2.3 Cartesian coordinates in a plane . . . . . . . . . . . . . . . . . . . . 292.2.4 Basis-vectors in plane polar-coordinates . . . . . . . . . . . . . . . . 292.2.5 Non-diagonal basis-vectors . . . . . . . . . . . . . . . . . . . . . . . . 292.2.6 Cartesian coordinates in a plane . . . . . . . . . . . . . . . . . . . . 312.2.7 Plane polar coordinates . . . . . . . . . . . . . . . . . . . . . . . . . 312.3.1 antisymmetric combinations . . . . . . . . . . . . . . . . . . . . . . . 322.3.2 antisymmetric combinations . . . . . . . . . . . . . . . . . . . . . . . 322.3.3 A 2-form in 3-space . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324.1.1 Outer product of 1-forms in 3-space . . . . . . . . . . . . . . . . . . 504.1.2 The derivative of a vector field with rotation . . . . . . . . . . . . . . 524.2.1 The Christoffel symbols in plane polar coordinates . . . . . . . . . . 544.3.1 vertical motion of free particle in hyperb. acc. ref. frame . . . . . . . 564.5.1 How geodesics in spacetime can give parabolas in space . . . . . . . 614.5.2 Spatial geodesics described in the reference frame of a rotating disc. 624.5.3 Christoffel symbols in a hyperbolically accelerated reference frame . 654.5.4 Vertical projectile motion in a hyperbolically accelerated reference

frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664.5.5 The twin “paradox” . . . . . . . . . . . . . . . . . . . . . . . . . . . . 684.5.6 Measurements of gravitational Doppler effects (Pound and Rebka 1960) 714.6.1 The connection coefficients in a rotating reference frame. . . . . . . . 724.6.2 Acceleration in a non-rotating reference frame (Newton) . . . . . . . 734.6.3 The acceleration of a particle, relative to the rotating reference frame 734.9.1 Cartan-connection in an orthonormal basis field in plane polar coord. 796.1.1 Energy momentum tensor for a Newtonian fluid . . . . . . . . . . . . 9410.4.1Age-redshift relation for dust dominated universe with k = 0 . . . . 145

xi

Page 16: LectureNotes On GeneralRelativity
Page 17: LectureNotes On GeneralRelativity

Chapter 1

Newton’s law of universalgravitation

1.1 The force law of gravitation

M

m

Fr

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

~F = −mGMr3~r = −mGM

r2~er (1.1)

Let V be the potential energy of m (see figure 1.1). Then

~F = −∇V (~r), Fi = −∂V∂xi

(1.2)

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

infinitely far from the center of M . Newton’s 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 energyof a test particle is very small compared to its rest mass energy. (Note thathere one is interested only in the absolute values of the above quantities andnot their sign).

mGM

r mc2 ⇒ r GM

c2. (1.3)

1

Page 18: LectureNotes On GeneralRelativity

2 Chapter 1. Newton’s law of universal gravitation

The Schwarzschild radius for an object of mass M is Rs = 2GMc2

. Faroutside the Schwarzschild radius we have a weak field. To get a feeling formagnitudes consider that Rs u 1 cm for the Earth which is to be comparedwith RE u 6400 km. That is, the gravitational field at the Earth’s surface canbe said to be weak! This explains, in part, the success of the Newtonian theory.

1.2 Newton’s law of gravitation in its local form

Let P be a point in the field (see figure 1.2) with position vector ~r = xi~ei andlet the gravitating point source be at ~r′ = xi

′~ei′ . Newton’s law of gravitation

for a continuous distribution of mass is

~F = −mG∫ ∞

rρ(~r′)

~r − ~r′|~r − ~r′|3

d3r′

= −∇V (~r)

(1.4)

See figure (1.2) for symbol definitions.

~r′

~r

~r − ~r′P

Figure 1.2: Newton’s law of gravitation in its local form.

Page 19: LectureNotes On GeneralRelativity

1.2 Newton’s law of gravitation in its local form 3

Let’s consider equation (1.4) term by term.

∇ 1

|~r − ~r′|= ~ei

∂xi

1[(xj − xj′)(xj − xj′)

]1/2

= ~ei∂

∂xi

[(xj − xj′)(xj − xj′)

]−1/2

= ~ei−1

22(xj − xj′)

∂xj

∂xi

[(xk − xk′)(xk − xk′)

]−3/2

= −~ei(xj − xj′)δij

[(xk − xk′)(xk − xk′)]3/2

= −~ei(xi − xi′)

[(xj − xj′)(xj − xj′)

]3/2

= − ~r − ~r′|~r − ~r′|3

(1.5)

Now equations (1.4) and (1.5) together ⇒

V (~r) = −mG∫

ρ(~r′)

|~r − ~r′|d3r′ (1.6)

Gravitational potential at point P :

φ(~r) ≡ V (~r)

m= −G

∫ρ(~r′)

|~r − ~r′|d3r′

⇒ ∇φ(~r) = G

∫ρ(~r′)

~r − ~r′|~r − ~r′|3

d3r′

⇒ ∇2φ(~r) = G

∫ρ(~r′)∇· ~r −

~r′

|~r − ~r′|3d3r′

(1.7)

The above equation simplifies considerably if we calculate the divergence in theintegrand. Note that “∇”

operates on ~ronly!∇· ~r −

~r′

|~r − ~r′|3=

∇·~r|~r − ~r′|3

+ (~r − ~r′) · ∇ 1

|~r − ~r′|3

=3

|~r − ~r′|3− (~r − ~r′) · 3(~r − ~r′)

|~r − ~r′|5

=3

|~r − ~r′|3− 3

|~r − ~r′|3= 0 ∀ ~r 6= ~r′

(1.8)

We conclude that the Newtonian gravitational potential at a point in a gravi-tational field outside a mass distribution satisfies Laplace’s equation

∇2φ = 0 (1.9)

Page 20: LectureNotes On GeneralRelativity

4 Chapter 1. Newton’s law of universal gravitation

Digression 1.2.1 (Dirac’s delta function)The Dirac delta function has the following properties:

1. δ(~r − ~r′) = 0 ∀ ~r 6= ~r′

2.∫δ(~r − ~r′)d3r′ = 1 when ~r = ~r′ is contained in the integration domain. Theintegral is identically zero otherwise.

3.∫f(~r′)δ(~r − ~r′)d3r′ = f(~r)

A calculation of the integral∫∇· ~r−~r′|~r−~r′|3d

3r′ which is valid also in the case wherethe field point is inside the mass distribution is obtained through the use ofGauss’ integral theorem:

v

∇· ~Ad3r′ =∮

s

~A · d~s, (1.10)

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

Definition 1.2.1 (Solid angle)

dΩ ≡ ds′⊥|~r − ~r′|2

(1.11)

where ds′⊥ is the projection of the area ds′ normal to the line of sight. ~ds′⊥ is thecomponent vector of ~ds′ along the line of sight which is equal to the normal vectorof ds′⊥ (see figure (1.3)).

Now, let’s apply Gauss’ integral theorem.∫

v

∇· ~r −~r′

|~r − ~r′|3d3r′ =

s

~r − ~r′|~r − ~r′|3

· d~s′ =∮

s

ds′⊥|~r − ~r′|2

=

s

dΩ (1.12)

So that,∫

v

∇· ~r −~r′

|~r − ~r′|3d3r′ =

4π if P is inside the mass distribution,0 if P is outside the mass distribution.

(1.13)

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

∇· ~r −~r′

|~r − ~r′|3= 4πδ(~r − ~r′) (1.14)

Page 21: LectureNotes On GeneralRelativity

1.3 Tidal Forces 5

~r′

~r~r − ~r′

P

d~s′⊥

d~s′ normal to bounding surface

d~s′⊥ = ~r−~r′|~r−~r′| · d

~s′

Figure 1.3: The solid angle dΩ is defined such that the surface of a spheresubtends 4π at the center

We now have

∇2φ(~r) = G

∫ρ(~r′)∇· ~r −

~r′

|~r − ~r′|3d3r′

= G

∫ρ(~r′)4πδ(~r − ~r′)d3r′

= 4πGρ(~r)

(1.15)

Newton’s theory of gravitation can now be expressed very succinctly indeed!

1. Mass generates gravitational potential according to

∇2φ = 4πGρ (1.16)

2. Gravitational potential generates motion according to

~g = −∇φ (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 agravitational field. The tidal force is due to the inhomogeneity of a gravitationalfield.

In figure 1.4 two points have a separation vector ~ζ. The position vectors of 1and 2 are ~r and ~r+ ~ζ, respectively, where |~ζ| |~r|. The gravitational forces on

Page 22: LectureNotes On GeneralRelativity

6 Chapter 1. Newton’s law of universal gravitation

F

F

2

1

2

Figure 1.4: Tidal Forces

a mass m at 1 and at 2 are ~F (~r) and ~F (~r+ ~ζ). By means of a Taylor expansionto lowest order in |~ζ| we get for the i-component of the tidal force

fi = Fi(~r + ~ζ)− Fi(~r) = ζj

(∂Fi∂xj

)

~r

. (1.18)

The corresponding vector equation is

~f = (~ζ · ∇)~r ~F . (1.19)

Using that~F = −m∇φ, (1.20)

the tidal force may be expressed in terms of the gravitational potential accordingto

~f = −m(~ζ · ∇)∇φ. (1.21)

It follows that in a local Cartesian coordinate system, the i-coordinate of therelative acceleration of the particles is

d2ζidt2

= −(

∂2φ

∂xi∂xj

)

~r

ζj. (1.22)

Let us look at a few simple examples. In the first one ~ζ has the same directionas ~g. Consider a small Cartesian coordinate system at a distance R from a massM (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

Fz(+z) = −m GM

(R+ z)2(1.23)

while an identical particle at the origin will be acted upon by the force

Fz(0) = −mGM

R2. (1.24)

Page 23: LectureNotes On GeneralRelativity

1.3 Tidal Forces 7

R

F

F

z

z

(0)

( )+z

m

M

y

z

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

If this little coordinate system is falling freely towards M , an observer atthe origin will say that the particle at (0, 0,+z) is acted upon by a force

fz = Fz(z) − Fz(0) ≈ 2mzGM

R3(1.25)

directed away from the origin, along the positive z-axis. We have assumedz R. This is the tidal force.

In the same way particles at the points (+x, 0, 0) and (0,+y, 0) are attractedtowards the origin by tidal forces

fx = −mxGMR3

, (1.26)

fy = −myGMR3

. (1.27)

Eqs. (1.25)–(1.27) have among others the following consequence: If an elastic,circular ring is falling freely in the Earth’s gravitational field, as shown in figure1.6, it will be stretched in the vertical direction and compressed in the horizontaldirection.

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 massof the Moon, and choose a coordinate system with origin at the Earth’s center.The tidal force per unit mass at a point is the negative gradient of the tidalpotential

φ(~r) = −GMR3

(z2 − 1

2x2 − 1

2y2

)= −GM

2R3r2(3 cos2 θ − 1), (1.28)

Page 24: LectureNotes On GeneralRelativity

8 Chapter 1. Newton’s law of universal gravitation

Figure 1.6: An elastic, circular ring falling freely in the Earth’s gravitationalfield

where we have introduced spherical coordinates, z = r cos θ, x2 + y2 = r2 sin2 θ,R is the distance between the Earth and the Moon, and the radius r of thespherical 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 theattraction of the Moon. Thus,

Θ(r) = gh − GM

2R3r2(3 cos2 θ − 1). (1.29)

At equilibrium, the surface of the Earth will be an equipotential surface,given by Θ = constant. The height of the water at flood, θ = 0 or θ = π, istherefore

hflood = h0 +GM

gR

( rR

)2, (1.30)

where h0 is an unknown constant. The height of the water at ebb (θ = π2 or

θ = 3π2 ) is

hebb = h0 −1

2

GM

gR

( rR

)2. (1.31)

The height difference between flood and ebb is therefore

∆h =3

2

GM

gR

( rR

)2. (1.32)

For a numerical result we need the following values:

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

R = 3.85 · 105km, rEarth = 6378km. (1.34)

With these values we find ∆h = 53cm, which is typical of tidal height differences.

Page 25: LectureNotes On GeneralRelativity

1.4 The Principle of Equivalence 9

1.4 The Principle of Equivalence

Galilei investigated experimentally the motion of freely falling bodies. He foundthat they moved in the same way, regardless what sort of material they consistedof and what mass they had.

In Newton’s theory of gravitation mass appears in two different ways; asgravitational mass, mG, in the law of gravitation, analogously to charge inCoulomb’s law, and as inertial mass, mI in Newton’s 2nd law.

The equation of motion of a freely falling particle in the field of gravity froma spherical body with mass M then takes the form

d2~r

dt2= −GmG

mI

M

r3~r. (1.35)

The results of Galilei’s measurements imply that the quotient between gravita-tional and inertial mass must be the same for all bodies. With a suitable choiceof units, we then obtain

mG = mI . (1.36)

Measurements performed by the Hungarian baron Eötvös around the turnof the century indicated that this equality holds with an accuracy better than10−8. More recent experiments have given the result | mImG

− 1| < 9 · 10−13.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 ofa gravitational force by being in free fall. In order to clarify this, Einsteinconsidered a homogeneous gravitational field in which the acceleration of gravity,g, is independent of the position. In a freely falling, non-rotating reference framein this field, all free particles move according to

mId2~r

dt2= (mG −mI)~g = 0, (1.37)

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 withconstant velocities along straight paths. In other words, such a reference frameis inertial.

Einstein’s heuristic reasoning suggests equivalence between inertial frames inregions far from mass distributions, where there are no gravitational fields, andinertial frames falling freely in a gravitational field. This equivalence between alltypes of inertial frames is so intimately connected with the equivalence betweengravitational and inertial mass, that the term “principle of equivalence” is usedwhether one talks about masses or inertial frames. The equivalence of differenttypes of inertial frames encompasses all types of physical phenomena, not onlyparticles 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

Page 26: LectureNotes On GeneralRelativity

10 Chapter 1. Newton’s law of universal gravitation

an accelerated reference frame in a region far from any mass distributions, willobtain identical results when they perform similar experiments. An inertialfield caused by the acceleration of the reference frame, is equivalent to a field ofgravity 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 generalprinciple 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 allinertial frames.

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

These two formulations may be interpreted as different formulations of asingle principle. But the generalization of S1 and S2 to the general case, whichencompasses accelerated motion and non-inertial frames, leads to two differentprinciples 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 ofrelativity. But G2 is a stronger principle (i.e. stronger restriction on naturalphenomena) than G1. Generally the course of events of a physical processin a certain reference frame, depends upon the laws of physics, the boundaryconditions, the motion of the reference frame and the geometry of space-time.The two latter properties are described by means of a metrical tensor. Byformulating the physical laws in a metric independent way, one obtains that G1is valid for all types of physical phenomena.

Even if the laws of Nature are the same in all reference frames, the course ofevents of a physical process will, as mentioned above, depend upon the motionof the reference frame. As to the spreading of light, for example, the law is thatlight follows null-geodesic curves (see ch. 4). This law implies that the path ofa light particle is curved in non-inertial reference frames and straight in inertialframes.

The question whether G2 is true in the general theory of relativity has beenthoroughly discussed recently, and the answer is not clear yet.

Page 27: LectureNotes On GeneralRelativity

1.6 The covariance principle 11

1.6 The covariance principle

The principle of relativity is a physical principle. It is concerned with physicalphenomena. This principle motivates the introduction of a formal principle,called the covariance principle: The equations of a physical theory shall havethe same form in every coordinate system.

This principle is not concerned directly with physical phenomena. Theprinciple 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 obtaina description of what happens in the presence of gravitation. We then startwith the physical laws as formulated in the special theory of relativity. Thenthe 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 adescription 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 examplethe principle of relativity, is concerned with observable relationships. Therefore,when one is going to deduce the observable consequences of an equation, onehas to establish relations between the tensor-components of the equation andobservable physical quantities. Such relations have to be defined; they are notdetermined by the covariance principle.

From the tensor equations, that are covariant, and the defined relationsbetween the tensor components and the observable physical quantities, one candeduce equations between physical quantities. The special principle of relativity,for example, demands that the laws which these equations express must be thesame 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 avector within Newtonian kinematics. However, in the relativistic kinematics offour-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 fourcomponents. Equations between physical quantities are not covariant in general.

For example, Maxwell’s 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 willchange the form of the equations.

If all equations of a theory are tensor equations, the theory is said to be givena manifestly covariant form. A theory that is written in a manifestly covariantform, will automatically fulfill the covariance principle, but it need not fulfillthe principle of relativity.

Page 28: LectureNotes On GeneralRelativity

12 Chapter 1. Newton’s law of universal gravitation

1.7 Mach’s principle

Einstein gave up Newton’s idea of an absolute space. According to Einstein allmotion is relative. This may sound simple, but it leads to some highly non-trivialand fundamental questions.

Imagine that there are only two particles connected by a spring, in theuniverse. What will happen if the two particles rotate about each other? Willthe spring be stretched due to centrifugal forces? Newton would have confirmedthat this is indeed what will happen. However, when there is no longer anyabsolute space that the particles can rotate relatively to, the answer is not soobvious. 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 nowkinematically 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 motionof a particle in an empty universe is not defined. All motion is motion relativelyto something else, i.e. relatively to other masses. According to Mach this impliesthat inertial forces must be due to a particle’s acceleration relatively to the greatmasses of the universe. If there were no such cosmic masses, there would notexist inertial forces, like the centrifugal force. In our example with two particlesconnected by a string, there would not be any stretching of the spring, if therewere no cosmic masses that the particles could rotate relatively to.

Another example may be illustrated by means of a turnabout. If we stayon 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 toMach identical centrifugal forces should appear if the turnabout is static andthe heavenly bodies rotate.

Einstein was strongly influenced by Mach’s arguments, which probably hadsome influence, at least with regards to motivation, on Einstein’s constructionof his general theory of relativity. Yet, it is clear that general relativity does notfulfill all requirements set by Mach’s principle. For example there exist generalrelativistic, rotating cosmological models, where free particles will tend to rotaterelative to the cosmic masses of the model.

However, some Machian effects have been shown to follow from the equationsof the general theory of relativity. For example, inside a rotating, massiveshell the inertial frames, i.e. the free particles, are dragged on and tend torotate in the same direction as the shell. This was discovered by Lense andThirring in 1918 and is therefore called the Lense-Thirring effect. More recentinvestigations of this effect have, among others, lead to the following result (Brilland Cohen 1966): “A massive shell with radius equal to its Schwarzschild radiushas often been used as an idealized model of our universe. Our result showsthat in such models local inertial frames near the center cannot rotate relativelyto the mass of the universe. In this way our result gives an explanation inaccordance with Mach’s principle, of the fact that the “fixed stars” is at rest onheaven as observed from an inertial reference frame.”

Page 29: LectureNotes On GeneralRelativity

Chapter 2

Vectors, Tensors and Forms

2.1 Vectors

An expression on the form aµ~eµ, where aµ, µ = 1, 2, ..., n are real numbers, isknown as a linear combination of the vectors ~eµ.

The vectors ~e1, ..., ~en are said to be linearly independent if there does notexist real numbers aµ 6= 0 such that aµ~eµ = 0.

Figure 2.1: Closed polygon (linearly dependent)

Geometrical interpretation: A set of vectors are linearly independent if itis not possible to construct a closed polygon of the vectors (even by adjustingtheir lengths).

A set of vectors ~e1, . . . , ~en are said to bemaximally linearly independentif ~e1, . . . , ~en, ~v are linearly dependent for all vectors ~v 6= ~eµ. We define thedimension of a vector-space as the number of vectors in a maximally linearlyindependent set of vectors of the space. The vectors ~eµ in such a set are known

13

Page 30: LectureNotes On GeneralRelativity

14 Chapter 2. Vectors, Tensors and Forms

as the basis-vectors of the space.

~v + aµ~eµ = 0

⇓~v = −aµ~eµ (2.1)

The components of ~v are the numbers vµ defined by vµ = −aµ ⇒ ~v = vµ~eµ.

2.1.1 4-vectors

4-vectors are vectors which exist in (4-dimensional) space-time. A 4-vectorequation represents 4 independent component equations.

L c

cL

v ∆ t

2

v

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

Example 2.1.1 (Photon clock)Carriage at rest:

∆t0 =2L

c

Page 31: LectureNotes On GeneralRelativity

2.1 Vectors 15

Carriage with velocity ~v:

∆t =2√

(v∆t2 )2 + L2

c⇓

c2∆t2 = v2∆t2 + 4L2

∆t =2L√c2 − v2

=2L/c√

1− v2/c2=

∆t0√1− v2/c2

(2.2)

The proper time-interval is denoted by dτ (above it was denoted ∆t0). Theproper time-interval for a particle is measured with a standard clock whichfollows the particle.

Definition 2.1.1 (4-velocity)

~U = cdt

dτ~et +

dx

dτ~ex +

dy

dτ~ey +

dz

dτ~ez, (2.3)

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

~U = Uµ~eµ =dxµ

dτ~eµ, xµ = (ct, x, y, z), x0 ≡ ct

dt

dτ=

1√1− v2

c2

≡ γ (2.4)

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

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 = γm0~v = m~v and E is the relativistic energy.

The 4-force or Minkowski-force ~F ≡ d~Pdτ and the ’common force’ ~f = d~p

dt .Then

~F = γ(1

c~f · ~v, ~f) (2.6)

Page 32: LectureNotes On GeneralRelativity

16 Chapter 2. Vectors, Tensors and Forms

lightcone

world line of a material particle

ct

y

x

should have v > ctachyons, if they exist,

Figure 2.3: World-lines in a Minkowski diagram

Definition 2.1.3 (4-acceleration)

~A =d~U

dτ(2.7)

The 4-velocity has the scalar value c so that

~U · ~U = −c2 (2.8)

The 4-velocity identity eq. 2.8 gives ~U · ~A = 0, in other words ~A ⊥ ~U and ~A isspace-like.

The line element for Minkowski space-time (flat space-time) with Cartesiancoordinates is

ds2 = −c2dt2 + dx2 + dy2 + dz2 (2.9)

In general relativity theory, gravitation is not considered a force. Gravitationis instead described as motion in a curved space-time.

A particle in free fall, is in Newtonian gravitational theory said to be onlyinfluenced by the gravitational force. According to general relativity theory theparticle is not influenced by any force.

Such a particle has no 4-acceleration. ~A 6= 0 implies that the particle is notin 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.

Page 33: LectureNotes On GeneralRelativity

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 figure2.4).

P

N(North pole)

Figure 2.4: In curved space,vectors can only exist in tangent planes.The vectorsin the tangent plane of N,do not contain the vector

−−→NP (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, infinitesimalposition vectors d~r do exist.

P

tangent plane of point P:

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

Page 34: LectureNotes On GeneralRelativity

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 worldlines in spacetime.

We can view a reference frame as a set of reference particles with a specifiedmotion. 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 coordinatesfor events in spacetime.

Definition 2.1.6 (Comoving coordinate system)A comoving coordinate system in a frame is a coordinate system where theparticles in the reference frame have constant spatial coordinates.

Definition 2.1.7 (Orthonormal basis)An orthonormal basis ~eµ in spacetime is defined by

~et · ~et = −1(c = 1)

~ei · ~ej = δij(2.10)

where i and j are space indices.

Definition 2.1.8 (Coordinate basis vectors.)Temporary definition of coordinate basis vector:

Assume any coordinate system xµ.

~eµ ≡∂~r

∂xµ(2.11)

A vector field is a continuum of vectors in a space, where the components arecontinuous and differentiable functions of the coordinates. Let ~v be a tangentvector to the curve ~r(λ):

~v =d~r

dλwhere ~r = ~r[xµ(λ)] (2.12)

Page 35: LectureNotes On GeneralRelativity

2.1 Vectors 19

The chain rule for differentiation yields:

~v =d~r

dλ=

∂~r

∂xµdxµ

dλ=dxµ

dλ~eµ = vµ~eµ (2.13)

Thus, the components of the tangent vector field along a curve, parameterisedby λ, is given by:

vµ =dxµ

dλ(2.14)

In the theory of relativity, the invariant parameter is often chosen to be theproper time. Tangent vector to the world line of a material particle:

uµ =dxµ

dτ(2.15)

These are the components of the 4-velocity of the particle!

Digression 2.1.1 (Proper time of the photon.)Minkowski-space:

ds2 = −c2dt2 + dx2

= −c2dt2(1− 1

c2(dxdt

)2)

= −(1− v2

c2)c2dt2

(2.16)

For a photon,v = c so:

limv→c

ds2 = 0 (2.17)

Thus, the spacetime interval between two points on the world line of a photon, iszero! This also means that the proper time for the photon is zero!! (See example2.1.2).

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

ds2 = −c2dτ2 (2.18)

where dτ is the proper time interval between two events, measured on a clockmoving in a way, such that it is present on both events (figure 2.6).

−c2dτ2 = −c2(1− v2

c2)dt2

⇒ dτ =

√1− v2

c2dt (2.19)

Page 36: LectureNotes On GeneralRelativity

20 Chapter 2. Vectors, Tensors and Forms

x

ct

d

P 1

P 2τ

Figure 2.6: P1 and P2 are two events in spacetime, separated by a proper timeinterval dτ .

The time interval between to events in the laboratory, is smaller measured on amoving clock than measured on a stationary one, because the moving clock isticking slower!

2.1.3 Coordinate transformations

Given two coordinate systems xµ and xµ′.

~eµ′ =∂~r

∂xµ′(2.20)

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

~eµ′ =∂~r

∂xµ′ =

∂~r

∂xµ∂xµ

∂xµ′ = ~eµ

∂xµ

∂xµ′ (2.21)

This is the transformation equation for the basis vectors. ∂xµ

∂xµ′are elements

of the transformation matrix. Indices that are not sum-indices are called ’freeindices’.

Rule: In all terms on each side in an equation, the free indices shouldbehave identically (high or low), and there should be exactly the sameindices in all terms!

Page 37: LectureNotes On GeneralRelativity

2.1 Vectors 21

Applying this rule, we can now find the inverse transformation

~eµ = ~eµ′∂xµ

∂xµ

~v = vµ′~eµ′ = vµ~eµ = vµ

′~eµ∂xµ

∂xµ′

So, the transformation rules for the components of a vector becomes

vµ = vµ′ ∂xµ

∂xµ′; vµ

′= vµ

∂xµ′

∂xµ(2.22)

The directional derivative along a curve, parametrised by λ:

d

dλ=

∂xµdxµ

dλ= vµ

∂xµ(2.23)

where vµ = dxµ

dλ are the components of the tangent vector of the curve. Direc-tional derivative along a coordinate curve:

λ = xν∂

∂xµ∂xµ

∂xν= δµν

∂xµ=

∂xν(2.24)

In the primed system:

∂xµ′=∂xµ

∂xµ′∂

∂xµ(2.25)

Definition 2.1.9 (Coordinate basis vectors.)We define the coordinate basis vectors as:

~eµ =∂

∂xµ(2.26)

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

Example 2.1.2 (Coordinate transformation)From Figure 2.7 we see that

x = r cos θ, y = r sin θ (2.27)

Coordinate basis vectors were defined by

~eµ ≡∂

∂xµ(2.28)

Page 38: LectureNotes On GeneralRelativity

22 Chapter 2. Vectors, Tensors and Forms

y

x

y

e

er

θ

e

e

y

x

r

θ

x

Figure 2.7: Coordinate transformation, flat space.

This means that we have

~ex =∂

∂x, ~ey =

∂y, ~er =

∂r, ~eθ =

∂θ

~er =∂

∂r=∂x

∂r

∂x+∂y

∂r

∂y

(2.29)

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

~er = cos θ ~ex + sin θ ~ey

~eθ =∂

∂θ=∂x

∂θ

∂x+∂y

∂θ

∂y

= −r sin θ ~ex + r cos θ ~ey

(2.30)

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

~er · ~er = cos2θ + sin2θ = 1 (2.31)

So ~er is a unit vector, |~er| = 1.

~eθ · ~eθ = r2(cos2θ + sin2θ) = r2 (2.32)

and we see that ~eθ is not a unit vector, |~eθ| = r. But we have that ~er · ~eθ = 0⇒~er⊥~eθ. Coordinate basis vectors are not generally unit vectors.

Page 39: LectureNotes On GeneralRelativity

2.1 Vectors 23

Definition 2.1.10 (Orthonormal basis)An orthonormal basis is a vector basis consisting of unit vectors that are normal toeach other. To show that we are using an orthonormal basis we will use ’hats’ overthe indices, ~eµ.

Orthonormal basis associated with planar polar coordinates:

~er = ~er , ~eθ =1

r~eθ (2.33)

Example 2.1.3 (Relativistic Doppler Effect)The Lorentz transformation is known from special relativity and relates the referenceframes of two systems where one is moving with a constant velocity v with regardto the other,

x′ = γ(x− vt)t′ = γ(t− vx

c2)

According to the vector component transformation (2.22), the 4-momentum for aparticle moving in the x-direction, P µ = (Ec , p, 0, 0) transforms as

P µ′

=∂xµ

∂xµP µ,

E′ = γ(E − vp).

Using the fact that a photon has energy E = hν and momemtum p = hνc , where

h is Planck’s constant and ν is the photon’s frequency, we get the equation for thefrequency shift known as the relativistic Doppler effect,

ν ′ = γ(ν − v

cν) =

(1− v

c

)ν√(

1− vc

) (1 + v

c

)

ν ′

ν=

√c− vc+ v

(2.34)

2.1.4 Structure coefficientsDefinition 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)

Page 40: LectureNotes On GeneralRelativity

24 Chapter 2. Vectors, Tensors and Forms

where ~u~v is defined as

~u~v ≡ uµ ~eµ(vν ~eν) = uµ∂

∂xµ(vν

∂xν) (2.36)

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

~u~v = uµ∂vν

∂xµ∂

∂xν+ uµvν

∂2

∂xµ∂xν

= uµ∂vν

∂xµ~eν + uµvν

∂2

∂xµ∂xν

(2.37)

Due to the last term, ~u~v is not a vector.

~v~u = vν∂

∂xν(uµ

∂xµ)

= vν∂uµ

∂xν~eµ + vνuµ

∂2

∂xν∂xµ

~u~v − ~v~u = uµ∂vν

∂xµ~eν − vν

∂uµ

∂xν~eµ

︸ ︷︷ ︸vµ ∂u

ν

∂xµ~eν

= (uµ∂vν

∂xµ− vµ ∂u

ν

∂xµ)~eν

(2.38)

Here we have used that∂2

∂xµ∂xν=

∂2

∂xν∂xµ(2.39)

The Einstein comma notation ⇒~u~v − ~v~u = (uµvν,µ − vµuν,µ)~eν (2.40)

As we can see, the commutator between two vectors is itself a vector.

Definition 2.1.12 (Structure coefficients cρµν)The structure coefficients cρµν in an arbitrary basis ~eµ are defined by:

[ ~eµ , ~eν ] ≡ cρµν ~eρ (2.41)

Structure coefficients in a coordinate basis:

[ ~eµ , ~eν ] = [∂

∂xµ,

∂xν]

=∂

∂xµ(∂

∂xν)− ∂

∂xν(∂

∂xµ)

=∂2

∂xµ∂xν− ∂2

∂xν∂xµ= 0

(2.42)

Page 41: LectureNotes On GeneralRelativity

2.2 Tensors 25

The commutator between two coordinate basis vectors is zero, so the structurecoefficients 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

~er = ~er , ~eθ =1

r~eθ (2.43)

We will now use this to find the structure coefficients.

[~er , ~eθ] = [∂

∂r,

1

r

∂θ]

=∂

∂r(1

r

∂θ)− 1

r

∂θ(∂

∂r)

= − 1

r2

∂θ+

1

r

∂2

∂r∂θ− 1

r

∂2

∂θ∂r

= − 1

r2~eθ = −1

r~eθ

(2.44)

To find the structure coefficients in coordinate basis we must use [~e r , ~eθ] = −1r~eθ.

[~eµ , ~eν ] = cρµν~eρ (2.45)

Using (2.44) and (2.45) we get

cθrθ

= −1

r(2.46)

From the definition of cρµν ([~u , ~v] = −[~v , ~u]) we see that the structure coefficientsare anti symmetric in their lower indices:

cρµν = −cρνµ (2.47)

cθθr

=1

r= −cθ

rθ(2.48)

2.2 Tensors

A 1-form-basis ω1, . . ., ωn is defined by:

ωµ(~eν) = δµν (2.49)

An arbitrary 1-form can be expressed, in terms of its components, as a linearcombination of the basis forms:

α = αµωµ (2.50)

Page 42: LectureNotes On GeneralRelativity

26 Chapter 2. Vectors, Tensors and Forms

where αµ are the components of α in the given basis.Using eqs.(2.49) and (2.50), we find:

α(~eν) = αµωµ(~eν) = αµδ

µν = αν

α(~v) = α(vµ~eµ) = vµα(~eµ) = vµαµ = v1α1 + v2α2 + . . .(2.51)

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 mapsone-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(NN ′)maps N one-forms and N ′ vectors into R. It is usual to

say that a tensor is of rank (N +N ′). A one-form, for example, is a covarianttensor of rank 1:

α(~v) = vµαµ (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 aretwo tensors of rank m and n, the tensor product is defined by:

T ⊗ S( ~u1,..., ~um, ~v1,..., ~vn) ≡ T ( ~u1,..., ~um)S(~v1,..., ~vn) (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 = Rµ1,...,µqωµ1 ⊗ ωµ2 ⊗ · · · ⊗ ωµq (2.54)

Notice that S ⊗ T 6= T ⊗ S. We get the components of a tensor (R) by usingthe tensor on the basis vectors:

Rµ1,...,µq = R( ~eµ1 ,..., ~eµq ) (2.55)

The indices of the components of a contravariant tensor are written as upperindices, and the indices of a covariant tensor as lower indices.

Page 43: LectureNotes On GeneralRelativity

2.2 Tensors 27

Example 2.2.1 (Example of a tensor)Let ~u and ~v be two vectors and α and β two 1-forms.

~u = uµ~eµ; ~v = vµ~eµ; α = αµωµ; β = βµω

µ (2.56)

From these we can construct tensors of rank 2 through the relation R = ~u ⊗ ~v asfollows: The components of R are

Rµ1µ2 = R(ωµ1 , ωµ2)

= ~u⊗ ~v(ωµ1 , ωµ2)

= ~u(ωµ1)~v(ωµ2)

= uµ~eµ(ωµ1)vν~eν(ωµ2)

= uµδµ1µ v

νδµ2ν

= uµ1vµ2

(2.57)

2.2.1 Transformation of tensor components

We shall not limit our discussion to coordinate transformations. Instead, wewill consider arbitrary transformations between bases, ~eµ −→

~eµ′. The

elements of transformation matrices are denoted by M µµ′ such that

~eµ′ = ~eµMµµ′ and ~eµ = ~eµ′M

µ′µ (2.58)

where Mµ′µ are elements of the inverse transformation matrix. Thus, it follows

that

Mµµ′M

µ′ν = δµν (2.59)

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

Mµ′µ =

∂xµ′

∂xµ(2.60)

2.2.2 Transformation of basis 1-forms

ωµ′

= Mµ′µω

µ

ωµ = Mµµ′ω

µ′(2.61)

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 aremultiplied with one another.

Page 44: LectureNotes On GeneralRelativity

28 Chapter 2. Vectors, Tensors and Forms

Example 2.2.2 (A mixed tensor of rank 3)

Tα′µ′ν′ = Mα′

αMµµ′M

νν′T

αµν (2.62)

The components in the primed basis are linear combinations of the componentsin 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 agiven basis then it has non-vanishing components in all bases. This meansthat tensor equations have a basis independent form. Tensor equations areinvariant. A basis transformation might result in the vanishing of one or moretensor components. Equations in component form may differ from one basis toanother. But an equation expressed in tensor components can be transformedfrom one basis to another using the tensor component transformation rules. Anequation that is expressed only in terms of tensor components is said to becovariant.

2.2.3 The metric tensorDefinition 2.2.3 (The metric tensor)The scalar product of two vectors ~u and ~v is denoted by g(~u,~v) and is defined asa 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 scalarproducts of each pair of basis-vectors in a basis.

g is a symmetric covariant tensor of rank 2. This tensor is known as themetric tensor. The components of this tensor are

g(~eµ, ~eν) = g µν (2.63)

~u · ~v = g(~u,~v) = g(uµ~eµ, vν~eν) = uµvνg(~eµ, ~eν) = uµuνg µν (2.64)

Usual notation:

~u · ~v = g µνuµvν (2.65)

The absolute value of a vector:

|~v| =√g(~v, ~v) =

√|g µνvµvν | (2.66)

Page 45: LectureNotes On GeneralRelativity

2.2 Tensors 29

e 1

2e

Θ

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)

g µν =

(1 00 1

)

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

~er · ~er = 1, ~eθ · ~eθ = r2, ~er · ~eθ = 0, (2.68)

The metric tensor in plane polar-coordinates:

g µν =

(1 00 r2

)(2.69)

Example 2.2.5 (Non-diagonal basis-vectors)

~e1 · ~e1 = 1, ~e2 · ~e2 = 1, ~e1 · ~e2 = cos θ = ~e2 · ~e1

g µν =

(1 cos θ

cos θ 1

)=

(g 11 g 12

g 21 g 22

) (2.70)

Definition 2.2.4 (Contravariant components)The contravariant components gµα of the metric tensor are defined as:

gµαg αν ≡ δµν gµν = ~wµ · ~wν , (2.71)

where ~wµ is defined by

~wµ · ~wν ≡ δµν . (2.72)

gµν is the inverse matrix of g µν .

Page 46: LectureNotes On GeneralRelativity

30 Chapter 2. Vectors, Tensors and Forms

x = constant

x = constant

1

2

A

A ω22

A e

e

22

2

ω 2

A2

1A e1 1e

1A

A ω11 ω 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:

vµ = g(~v, ~eµ) = g(vα~eα, ~eµ) = vαg(~eα, ~eµ) = vαg αµ (2.73)

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

vµ = gµαvα (2.74)

The mixed components of the metric tensor becomes:

gµν = gµαg αν = δµν (2.75)

We now define distance along a curve. Let the curve be parameterized by λ(proper-time τ for time-like curves). Let ~v be the tangent vector-field of thecurve.

The squared distance ds2 between the points along the curve is defined as:

ds2 ≡ g(~v, ~v)dλ2 (2.76)

gives

ds2 = g µνvµvνdλ2. (2.77)

Page 47: LectureNotes On GeneralRelativity

2.3 Forms 31

The tangent vector has components vµ = dxµ

dλ , which gives:

ds2 = g µνdxµdxν (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, gxy = gyx = 0

ds2 = dx2 + dy2(2.79)

Example 2.2.7 (Plane polar coordinates)

g rr = 1, g θθ = r2

ds2 = dr2 + r2dθ2(2.80)

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

ds2 = −c2dt2 + dx2 + dy2 + dz2 (2.81)

In an arbitrary curved space, an orthonormal basis can be adopted in anypoint. If ~et is tangent vector to the world line of an observer, then ~e t = ~uwhere ~u is the 4-velocity of the observer. In this case, we are using what we callthe comoving orthonormal basis of the observer. In a such basis, we have theMinkowski-metric:

ds2 = ηµνdxµdxν (2.82)

2.3 Forms

An antisymmetric tensor is a tensor whose sign changes under an arbitraryexchange of two arguments.

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

The components of an antisymmetric tensor change sign under exchange oftwo indices.

A···µ···ν··· = −A···ν···µ··· (2.84)

Page 48: LectureNotes On GeneralRelativity

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:

ω[µ1 ⊗ · · · ⊗ ωµp] ∧ ω[ν1 ⊗ · · · ⊗ ωνq ] ≡ (p+ q)!

p!q!ω[µ1 ⊗ · · · ⊗ ωνq ] (2.85)

where [ ] denotes antisymmetric combinations defined by:

ω[µ1 ⊗ · · · ⊗ ωµp] ≡ 1

p!· (the sum of terms with

all possible permutationsof indices with, “+” for evenand “-” for odd permutations)

(2.86)

Example 2.3.1 (antisymmetric combinations)

ω[µ1 ⊗ ωµ2] =1

2(ωµ1 ⊗ ωµ2 − ωµ2 ⊗ ωµ1) (2.87)

Example 2.3.2 (antisymmetric combinations)

ω[µ1 ⊗ ωµ2 ⊗ ωµ3] =

1

3!(ωµ1 ⊗ ωµ2 ⊗ ωµ3 + ωµ3 ⊗ ωµ1 ⊗ ωµ2 + ωµ2 ⊗ ωµ3 ⊗ ωµ1

− ωµ2 ⊗ ωµ1 ⊗ ωµ3 − ωµ3 ⊗ ωµ2 ⊗ ωµ1 − ωµ1 ⊗ ωµ3 ⊗ ωµ2)

=1

3!εijk(ω

µi ⊗ ωµj ⊗ ωµk) (2.88)

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

α = α12ω1⊗ω2 +α21ω

2⊗ω1 +α13ω1⊗ω3 +α31ω

3⊗ω1 +α23ω2⊗ω3 +α32ω

3⊗ω2

(2.89)

Now the antisymmetry of α means that

+α21 = −α12; +α31 = −α13; +α32 = −α23 (2.90)

Page 49: LectureNotes On GeneralRelativity

2.3 Forms 33

α =α12(ω1 ⊗ ω2 − ω2 ⊗ ω1)

+ α13(ω1 ⊗ ω3 − ω3 ⊗ ω1)

+ α23(ω2 ⊗ ω3 − ω3 ⊗ ω2)

= α|µν|2ω[µ ⊗ ων]

(2.91)

where |µν| means summation only for µ < ν (see (Misner, Thorne and Wheeler1973)). We now use the definition of ∧ with p = q = 1. This gives

α = α|µν|ωµ ∧ ων

ωµ ∧ ων is theform basis.We can also write

α =1

2αµνω

µ ∧ ων

A tensor of rank 2 can always be split up into a symmetric and an anti-symmetric part. (Note that tensors of higher rank can not be split up in thisway.)

Tµν =1

2(Tµν − Tνµ) +

1

2(Tµν + Tνµ)

= Aµν + Sµν

(2.92)

We thus have:

SµνAµν =

1

4(Tµν + Tνµ)(T µν − T νµ)

=1

4(TµνT

µν − TµνT νµ + TνµTµν − TνµT νµ)

= 0

(2.93)

In general, summation over indices of a symmetric and an antisymmetric quan-tity vanishes. In a summation TµνAµν where Aµν is antisymmetric and Tµν hasno symmetry, only the antisymmetric part of Tµν contributes. So that, in

α =1

2αµνω

µ ∧ ων (2.94)

only the antisymmetric elements ανµ = −αµν , contribute to the summation.These antisymmetric elements are the form components

Forms are antisymmetric covariant tensors. Because of this antisymmetrya form with two identical components must be a null form (= zero). e.g.α131 = −α131 ⇒ α131 = 0

In an n-dimensional space all p-forms with p > n are null forms.

Page 50: LectureNotes On GeneralRelativity

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 ina reference frame R. A set of simultanous events in R, defines a 3 dimensionalspace called ’3-space’ in R. This space is orthogonal to ~e0. We are going tofind the metric tensor γij in this space, expressed by the metric tensor gµν ofspacetime.

In an arbitrary coordinate basis ~eµ, ~ei is not necessarily orthogonal to~e0. We choose ~e0‖~e0. Let ~e⊥i be the component of ~ei orthogonal to ~e0, thatis:~e⊥i · ~e0 = 0. The metric tensor of space is defined by:

γij = ~e⊥i · ~e⊥j , γi0 = 0, γ00 = 0

~e⊥i = ~ei − ~e‖i~e‖i =

~ei · ~e0

~e0 · ~eo~e0 =

gi0g00

~e0

γij = (~ei − ~e‖i) · (~ej − ~e‖j)= (~ei −

gi0g00

~e0) · (~ej −gj0g00

~e0)

= ~ei · ~ej −gj0g00

~e0 · ~ei −gi0g00

~e0 · ~ej +gi0gj0g2

00

~e0 · ~e0

= gij −gi0gj0g00

− gi0gj0g00

+gi0gj0g00

⇒ γij = gij −gi0gj0g00

(3.1)

(Note:gij = gji ⇒ γij = γji)The line element in space:

dl2 = γijdxidxj =

(gij −

gi0gj0g00

)dxidxj (3.2)

34

Page 51: LectureNotes On GeneralRelativity

3.1 Rotating reference frames 35

gives the geometry of a “simultaneity space” in a reference frame where themetric tensor of spacetime in a comoving coordinate system is gµν .

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 withmetric γij .

dt2 = dl2 − ds2 = (γµν − gµν)dxµdxν

= (γij − gij)dxidxj + 2(γi0 − gi0)dxidx0 + (γ00 − g00)dx0dx0

= (gij −gi0gj0g00

− gij)dxidxj − 2gi0dxidx0 − g00(dx0)2

= −g00

[(dx0)2 + 2

gi0g00

dx0dxi +gi0gj0g2

00

dxidxj]

=

[(−g00)1/2(dx0 +

gi0g00

dxi)

]2

So finally we get

dt = (−g00)1/2(dx0 +gi0g00

dxi) (3.4)

The 3-space orthogonal to the world lines of the reference particles in R, dt = 0,corresponds to a coordinate time interval dt = − gi0

g00dxi. This is not an exact

differential , that is , the line integral of dt around a closed curve is in generalnot equal to 0. Hence you can not in general define simultaneity (given bydt = 0) around closed curves. It can, however, be done if the spacetime metricis diagonal, gi0 = 0. The condition dt = 0 means simultaneity on Einsteinsynchronized clocks . Conclusion:It is in general (gi0 6= 0) not possible toEinstein synchronize clocks around closed curves.

In particular, it is not possible to Einstein-synchronize clocks around a closedcurve 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 ofsimultaneity. (See figure 3.1)

The distance in the laboratory frame between two points is:

L0 =2πr

n(3.5)

Lorentz transformation from the instantaneous rest frame (x′, t′) to the labo-ratory system (x, t):

∆t = γ(∆t′ +v

c2∆x′), γ =

1√1− r2ω2

c2

∆x = γ(∆x′ + v∆t′)

(3.6)

Page 52: LectureNotes On GeneralRelativity

36 Chapter 3. Accelerated Reference Frames

r

t & t+n ∆ t

t+2∆ t

t+∆ t

t+(n-1) ∆ t

S

S

S

S3

n-1

n

1

2

2

n

ω

(discontinuity)

n-1

Figure 3.1: Events simultanous in the rotating reference frame. 1 comes before2, before 3, etc. . . Note the discontinuity at t.

0L v = rω

Figure 3.2: The distance between two points on the circumference is L0.

Page 53: LectureNotes On GeneralRelativity

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,and proper distance ∆x′ = γL0, we get in the laboratory frame

∆t = γ2 rω

c2L0 = γ2 rω

c22πr

n(3.7)

The fact that ∆t′ = 0 and ∆t 6= 0 is an expression of the relativity of simul-taneity. Around the circumference this is accumulated to

n∆t = γ2 2πr2ω

c2(3.8)

and we get a discontinuity in simultaneity, as shown in figure 3.3. Let IF be aninertial frame with cylinder coordinates (T, R, Θ, Z). The line element is thengiven by

ds2 = −dT 2 + dR2 +R2dΘ2 + dZ2 (c = 1) (3.9)

In a rotating reference frame, RF, we have cylinder coordinates (t, r, θ, z). Wethen have the following coordinate transformation :

t = T, r = R, θ = Θ− ωT, z = Z (3.10)

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

ds2 = −dt2 + dr2 + r2(dθ + ωdt)2 + dz2

= −(1− r2ω2)dt2 + dr2 + r2dθ2 + dz2 + 2r2ωdθdt (c = 1)(3.11)

Page 54: LectureNotes On GeneralRelativity

38 Chapter 3. Accelerated Reference Frames

The metric tensor have the following components:

gtt = −(1− r2ω2), grr = 1, gθθ = r2, gzz = 1

gθt = gtθ = r2ω(3.12)

dt = 0 gives

ds2 = dr2 + r2dθ2 + dz2 (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 lineelement:

dl2 = (gij −gi0gj0g00

)dxidxj

γrr = grr = 1, γzz = gzz = 1,

γθθ = gθθ −g2θ0

g00

= r2 − (r2ω)2

−(1− r2ω2)=

r2

1− r2ω2

⇒ dl2 = dr2 +r2dθ2

1− r2ω2+ dz2 (3.14)

So we have a non Euclidean spatial geometry in RF. The circumference of acircle with radius r is

lθ =2πr√

1− r2ω2> 2πr (3.15)

We see that the quotient between circumference and radius > 2π which meansthat the spatial geometry is hyperbolic. (For spherical geometry we have lθ <2πr.)

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 inFigure 3.4. All points on a circle are accelerated simultaneously in IF (thelaboratory system). We let the angular velocity increase from ω to ω + dω,measured in IF. Lorentz transformation to an instantaneous rest frame for apoint on the circumference then gives an increase in velocity in this system:

rdω′ =rdω

1− r2ω2, (3.16)

where we have used that the initial velocity in this frame is zero.

Page 55: LectureNotes On GeneralRelativity

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 therods (the front end is accelerated first) in the instantaneous rest frame is:

∆t′ =rωL0√

1− r2ω2(3.17)

where L0 is the distance between points on the circumference when at rest (= thelength of the rods when at rest), L0 = 2πr

n . In IF all points on the circumferenceare accelerated simultaneously. In RF, however, this is not the case. Here thedistance between points on the circumference will increase, see Figure 3.5. Therest distance increases by

dL′ = rdω′∆t′ =r2ωL0dω

(1− r2ω2)3/2. (3.18)

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

∆ t’ t’dv’

t’+

Figure 3.5: In RF two points on the circumference are accelerated at differenttimes. Thus the distance between them is increased.

Page 56: LectureNotes On GeneralRelativity

40 Chapter 3. Accelerated Reference Frames

rest frame) is

L′ = r2L0

∫ ω

0

ωdω

(1− r2ω2)3/2= (

1√1− r2ω2

− 1)L0. (3.19)

Hence, after the acceleration there is a proper distance L′ between the rods. Inthe laboratory system (IF) the distance between the rods is

L =√

1− r2ω2L′ =√

1− r2ω2(1√

1− r2ω2− 1)L0 = L0 − L0

√1− r2ω2,

(3.20)

where L0 is the rest length of the rods and L0

√1− r2ω2 is their Lorentz con-

tracted length. We now have the situation shown in Figure 3.6.

Lorenz contractedStandard measuring rod,

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

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

3.1.3 Gravitational time dilation

ds2 = −(1− r2ω2

c2)c2dt2 + dr2 + r2dθ2 + dz2 + 2r2ωdθdt (3.21)

We now look at standard clocks with constant r and z.

ds2 = c2dt2[−(1− r2ω2

c2) +

r2

c2(dθ

dt)2 + 2

r2ω

c2dθ

dt] (3.22)

Page 57: LectureNotes On GeneralRelativity

3.1 Rotating reference frames 41

Let dθdt ≡ θ be the angular velocity of the clock in RF. The proper time interval

measured by the clock is then

ds2 = −c2dτ2 (3.23)

From this we see that

dτ = dt

1− r2ω2

c2− r2θ2

c2− 2

r2ωθ

c2(3.24)

A non-moving standard clock in RF: θ = 0⇒

dτ = dt

√1− r2ω2

c2(3.25)

Seen from IF, the non-rotating laboratory system, (3.25) represents the velocitydependent time dilation from the special theory of relativity.

But how is (3.25) interpreted in RF? The clock does not move relative toan observer in this system, hence what happens can not bee interpreted as avelocity dependent phenomenon. According to Einstein, the fact that standardclocks slow down the farther away from the axis of rotation they are, is due toa gravitational effect.

We will now find the gravitational potential at a distance r from the axis.The sentripetal acceleration is v2/r, v = rω so:

Φ = −∫ r

0g(r)dr = −

∫ r

0rω2dr = −1

2r2ω2

We then get:

dτ = dt

√1− r2ω2

c2= dt

√1 +

c2(3.26)

In RF the position dependent time dilation is interpreted as a gravitationaltime 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 areradial. Consider a photon path with Θ = 0, R = T with light source at R = 0.Transforming to RF:

t = T, r = R, θ = Θ− ωT⇒ r = t, θ = −ωt (3.27)

The orbit equation is thus θ = −ωr which is the equation for an Archimedeanspiral. The time used by a photon out to distance r from axis is t = r

c .

Page 58: LectureNotes On GeneralRelativity

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 thedisc’s rotation as shown in Figure 3.7. Photons are emitted/received in/fromopposite directions. Let t1 be the travel time of photons which move with therotation.

+r

ω

X

Emitter/Receiver

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

Then⇒ 2πr + rωt1 = ct1

⇒ t1 =2πr

c− rω(3.28)

Let t2 be the travel time for photons moving against the rotation of the disc.The difference in travel time isA is the area

enclosed by thephoton path or

orbit.∆t = t1 − t2 = 2πr

(1

c− rω −1

c+ rω

)

=2πr2rω

c2 − r2ω2

= γ2 4Aω

c2

(3.29)

RF description:ds2 = 0 along the

world line of aphoton ds2 = −

(1− r2ω2

c2

)c2dt2 + r2dθ2 + 2r2ωdθdt

let θ =dθ

dt

r2θ2 + 2r2ωθ − (c2 − r2ω2) = 0

θ =−r2ω ±

√(r4ω2 + r2c2 − r4ω2)

r2

Page 59: LectureNotes On GeneralRelativity

3.2 Hyperbolically accelerated reference frames 43

θ = −ω ± rc

r2

= −ω ± c

r

(3.30)

The speed of light: v± = rθ = −rω ± c. We see that in the rotating frame RF,the measured (coordinate) velocity of light is NOT isotropic. The difference inthe travel time of the two beams is

∆t =2πr

c− rω −2πr

c+ rω

= γ2 4Aω

c2

(3.31)

(See Phil. Mag. series 6, vol. 8 (1904) for Michelson’s article)

3.2 Hyperbolically accelerated reference frames

Consider a particle moving along a straight line with velocity u and accelerationa = du

dT . Rest acceleration is a.

⇒ a =(1− u2/c2

)3/2a. (3.32)

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

du

dT=(1− u2/c2

)3/2g. (3.33)

Which on integration with u(0) = 0 gives

u =gT

(1 + g2T 2

c2

)1/2=dX

dT

⇒ X =c2

g

(1 +

g2T 2

c2

)1/2

+ k

⇒ c4

g2= (X − k)2 − c2T 2 (3.34)

In its final form the above equation describes a hyperbola in the Minkowskidiagram as shown in figure(3.8).

Page 60: LectureNotes On GeneralRelativity

44 Chapter 3. Accelerated Reference Frames

cT

X

Figure 3.8: Hyperbolically accelerated reference frames are so called becausethe loci of particle trajectories in space-time are hyperbolae.

The proper time interval as measured by a clock which follows the particle:

dτ =

(1− u2

c2

)1/2

dT (3.35)

Substitution for u(T ) and integration with τ(0) = 0 gives

τ =c

garcsinh

(gT

c

)

or T =c

gsinh

(gτc

)

and X =c2

gcosh

(gτc

)+ k

(3.36)

We now use this particle as the origin of space in an hyperbolically acceleratedreference frame.

Definition 3.2.1 (Born-stiff motion)Born-stiff motion of a system is motion such that every element of the system hasconstant rest length. We demand that our accelerated reference frame is Born-stiff.

Let the inertial frame have coordinates (T,X, Y, Z) and the acceleratedframe have coordinates (t, x, y, z). We now denote the X-coordinate of the“origin particle” by X0.

1 +gX0

c2= cosh

gτ0

c(3.37)

Page 61: LectureNotes On GeneralRelativity

3.2 Hyperbolically accelerated reference frames 45

where τ0 is the proper time for this particle and k is set to −c2

g . (These areMøller coordinates. Setting k = 0 gives Rindler coordinates).

Let us denote the accelerated frame by Σ. The coordinate time at an ar-bitrary point in Σ is defined by t = τ0. That is coordinate clocks in Σ runidentically with the standard clock at the “origin particle”. Let ~X0 be the posi-tion 4-vector of the “origin particle”. Decomposed in the laboratory frame, thisbecomes

~X0 =

c2

gsinh

gt

c,c2

g

(cosh

gt

c− 1

), 0, 0

(3.38)

P is chosen such that P and P0 are simultaneous in the accelerated frame Σ. Thedistance (see figure(3.9)) vector from P0 to P , decomposed into an orthonormalcomoving basis of the “origin particle” is X = (0, x, y, z) where x, y and z arephysical distances measured simultaneously in Σ. The space coordinates in Σare defined by

x ≡ x, y ≡ y, z ≡ z. (3.39)

The position vector of P is ~X = ~X0 + ~X. The relationship between basis vectorsin IF and the comoving orthonormal basis is given by a Lorentz transformationin the x-direction.

~eµ = ~eµ∂xµ

∂xµ

= (~eT , ~eX , ~eY , ~eZ , )

cosh θ sinh θ 0 0sinh θ cosh θ 0 0

0 0 1 00 0 0 1

(3.40)

where θ is the rapidity defined by

tanh θ ≡ U0

c(3.41)

U0 being the velocity of the “origin particle”.

U0 =dX0

dT0= c tanh

gt

c

∴ θ =gt

c

(3.42)

Page 62: LectureNotes On GeneralRelativity

46 Chapter 3. Accelerated Reference Frames

cT

X

~X0

P0

~et

P

~eX

~X

~X

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

So the basis vectors can be written as follows

~et = ~eT coshgt

c+ ~eX sinh

gt

c

~ex = ~eT sinhgt

c+ ~eX cosh

gt

c~ey = ~eY

~ez = ~eZ

(3.43)

The equation ~X = ~X0 + ~X can now be decomposed in IF:

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

c

gsinh

gt

c~eT +

c2

g

(cosh

gt

c− 1

)~eX +

x

csinh

gt

c~eT + x cosh

gt

c~eX + y~eY + z~eZ

(3.44)

Page 63: LectureNotes On GeneralRelativity

3.2 Hyperbolically accelerated reference frames 47

This then, gives the coordinate transformations

T =c

gsinh

gt

c+x

csinh

gt

c

X =c2

g

(cosh

gt

c− 1

)+ x cosh

gt

c

Y = y

Z = z

⇒ gT

c=(

1 +gx

c2

)sinh

gt

c

1 +gX

c2=(

1 +gx

c2

)cosh

gt

c

Now dividing the last two of the above equations we get

gT

c=

(1 +

gX

c2

)tanh

gt

c(3.45)

showing that the coordinate curves t = constant are straight lines in the T,X-frame passing through the point T = 0, X = − c2

g . Using the identity cosh2 θ−sinh2 θ = 1 we get

(1 +

gX

c2

)2

−(gT

c

)2

=(

1 +gx

c2

)2(3.46)

showing that the coordinate curves x = constant are hypebolae in the T,X-diagram.

The line element (the metric) gives : ds2 is aninvariantquantityds2 = −c2dT 2 + dX2 + dY 2 + dZ2

= −(1 +gx

c2)2c2dt2 + dx2 + dy2 + dz2 (3.47)

Note: When the metric is diagonal the unit vectors are orthogonal.Clocks as rest in the accelerated system:

dx = dy = dz = 0, ds2 = −c2dτ2

−c2dτ2 = −(1 +gx

c2)2c2dt2

dτ = (1 +gx

c2)dt (3.48)

Here dτ is the proper time and dt the coordinate time.An observer in the accelerated system Σ experiences a gravitational field in

the negative x-direction. When x < 0 then dτ < dt. The coordinate clocks

Page 64: LectureNotes On GeneralRelativity

48 Chapter 3. Accelerated Reference Frames

X

cT

light t=constant

x=constant

g-c 2

horizon

Figure 3.10: The hyperbolically accelerated reference system

tick equally fast independently of their position. This implies that time passesslower further down in a gravitational field.

Consider a standard clock moving in the x-direction with velocity v = dx/dt.Then

−c2dτ2 = −(

1 +gx

c2

)2c2dt2 + dx2

= −[(

1 +gx

c2

)2− v2

c2

]c2dt2 (3.49)

Hence

dτ =

√(1 +

gx

c2

)2− v2

c2dt (3.50)

This expresses the combined effect of the gravitational- and the kinematic timedilation.

Page 65: LectureNotes On GeneralRelativity

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:

df =∂f

∂xµωµ = f,µω

µ (4.1)

where ωµ are coordinate basis forms:

ωµ(∂

∂xν) = δµν (4.2)

We then (in general) get:

ωµ = δµνων =

∂xµ

∂xνων = dxµ (4.3)

In coordinate basis we can always write the basis forms as exterior derivativesof the coordinates. The differential dxµ is given by

dxµ(d~r) = dxµ (4.4)

where d~r is an infinitesimal position vector. dxµ are not infinitesimal quantities.In coordinate basis the exterior derivative of a p-form

α =1

p!αµ1···µpdx

µ1 ∧ · · · ∧ dxµp (4.5)

will have the following component form:

d α =1

p!αµ1···µp,µ0dx

µ0 ∧ dxµ1 ∧ · · · ∧ dxµp (4.6)

49

Page 66: LectureNotes On GeneralRelativity

50 Chapter 4. Covariant Differentiation

where , µ0 ≡ ∂∂xµ0 . The exterior derivative of a p-form is a (p+ 1)-form.

Consider the exterior derivative of a p-form α.

dα =1

p!αµ1 ···µp,µ0dx

µ0 ∧ · · · ∧ dxµp . (4.7)

Let (dα)µ0 ···µp be the form components of dα. They must, by definition, beantisymmetric under an arbitrary interchange of indices.

dα =1

(p+ 1)!(dα)µ0 ···µpdx

µ0 ∧ · · · ∧ dxµp

which, by (4.7)⇒ =1

p!α[αµ1···µp,µ0 ]dx

µ0 ∧ · · · ∧ dxµp

∴ (dα)µ0···µp = (p+ 1)α[µ1 ···µp,µ0] (4.8)

The form equation dα = 0 in component form is

α[µ1···µp,µ0] = 0 (4.9)

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

α = αidxi xi = (x, y, z)

dα = αi,jdxj ∧ dxi

(4.10)

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

α[i,j] = 0 ⇒ αi,j − αj,i = 0

⇒ ∂αx∂y− ∂αy

∂x= 0,

∂αx∂z− ∂αz

∂x= 0,

∂αy∂z− ∂αz

∂y= 0

(4.11)

which corresponds to

∇×~α = 0 (4.12)

The outer product of an outer product!

d2α ≡ d(dα)

d2α =1

p!αµ1···µp,ν1ν2dx

ν2 ∧ dxν1 ∧ · · · ∧ dxµp(4.13)

,ν1ν2 ≡∂2

∂xν1∂xν2(4.14)

Page 67: LectureNotes On GeneralRelativity

4.1 Differentiation of forms 51

Since

,ν1ν2 ≡∂2

∂xν1∂xν2=,ν2ν1≡

∂2

∂xν2∂xν1(4.15)

summation over ν1 and ν2 which are symmetric in αµ1···µp,ν1ν2 and antisymmetricin the basis we get Poincaré’s lemma (valid only for scalar fields)

d2α = 0 (4.16)

This corresponds to the vector equation

∇ · (∇× ~A) = 0 (4.17)

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

d(α ∧ β) = dα ∧ β + (−1)pα ∧ dβ (4.18)

4.1.2 Covariant derivative

The general theory of relativity contains a covariance principle which statesthat all equations expressing laws of nature must have the same form irrespectiveof the coordinate system in which they are derived. This is achieved by writingall equations in terms of tensors. Let us see if the partial derivative of vectorcomponents transform as tensor components. Given a vector ~A = Aµ~eµ =Aµ′~eµ′ with the transformation of basis given by

∂xν′=∂xν

∂xν′∂

∂xν(4.19)

So that

Aµ′

,ν′ ≡∂

∂xν′

(Aµ′)

=∂xν

∂xν′∂

∂xν

(Aµ′)

=∂xν

∂xν′∂

∂xν

(∂xµ′

∂xµAµ)

=∂xν

∂xν′∂xµ′

∂xµAµ,ν +

∂xν

∂xν′Aµ

∂2xµ′

∂xν∂xµ(4.20)

The first term corresponds to a tensorial transformation. The existence of thelast term shows that Aµ,ν does not, in general, transform as the components ofa tensor. Note that Aµ,ν will transform as a tensor under linear transformationssuch as the Lorentz transformations.

The partial derivative must be generalized such as to ensure that when it isapplied to tensor components it produces tensor components.

Page 68: LectureNotes On GeneralRelativity

52 Chapter 4. Covariant Differentiation

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

~A = kr~eθ

The chain rule for derivation gives:

d

dτ=

∂xν· dx

ν

dτ= uν

∂xν

d ~A

dτ= uν (Aµ~eµ),ν

= uν(Aµ,ν~eµ +Aµ~eµ,ν

)

The change of the vector field with a displacement along a coordinate-curve isexpressed by:

∂ ~A

∂xν= ~A,ν = Aµ,ν~eµ +Aµ~eµ,ν

The change in ~A with the displacement in the θ-direction is:

∂ ~A

∂θ= Aµ,θ~eµ +Aµ~eµ,θ

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

∂ ~A

∂θ= Aθ,θ︸︷︷︸

=0

~eθ +Aθ~eθ,θ

and since Aθ,θ = 0 because Aθ = kr we end up with

∂ ~A

∂θ= Aθ~eθ,θ = kr~eθ,θ

Page 69: LectureNotes On GeneralRelativity

4.2 The Christoffel Symbols 53

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

x = r cos θ y = r sin θ

Using ~eµ = ∂∂xµ we can write:

~eθ =∂

∂θ=∂x

∂θ

∂x+∂y

∂θ

∂y

= −r sin θ~ex + r cos θ~ey

~er =∂

∂r= cos θ~ex + sin θ~ey

Gives:

~eθ,θ = −r cos θ~ex − r sin θ~ey

= −r(cos θ~ex + sin θ~ey) = −r~er

This gives us finally:

∂ ~A

∂θ= −kr2~er

Thus ∂ ~A∂θ 6= 0 even if ~A = Aθ~eθ and Aθ,θ = 0.

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 partiallyderivative Aµ,ν to a derivative written as Aµ;ν and which transforms tensorially,

Aµ′

;ν′ ≡∂xµ

∂xµ· ∂x

ν

∂xν′Aµ;ν . (4.21)

The covariant derivative of the contravariant vector components are written as:

Aµ;ν ≡ Aµ,ν +AαΓµαν (4.22)

This equation defines the Christoffel symbols Γµαν , which are also called the“connection coefficients in coordinate basis”. From the transformation formulaefor the two first terms follows that the Christoffel symbols transform as:

Γα′µ′ν′ =

∂xν

∂xν′∂xµ

∂xµ′∂xα

∂xαΓαµν +

∂xα′

∂xα∂2xα

∂xµ′∂xν′(4.23)

The Christoffel symbols do not transform as tensor components. It is possible tocancel all Christoffel symbols by transforming into a locally Cartesian coordinate

Page 70: LectureNotes On GeneralRelativity

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-rotatingframe in free fall. The Christoffel symbols are 0 (zero) in a locally Cartesiancoordinate system which is co-moving in a local inertial frame. Local Gaussiancoordinates are indicated with a bar over the indices, giving

Γαµν = 0 (4.24)

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

Γα′µ′ν′ =

∂xα′

∂xα∂2xα

∂xµ′∂xν′(4.25)

This equation shows that the Christoffel symbols are symmetric in the two lowerindices, ie.

Γα′µ′ν′ = Γα

′ν′µ′ (4.26)

Example 4.2.1 (The Christoffel symbols in plane polar coordinates)

x = r cos θ, y = r sin θ

r =√x2 + y2, θ = arctan

y

x

∂x

∂r= cos θ,

∂x

∂θ= −r sin θ

∂r

∂x=x

r= cos θ,

∂r

∂y= sin θ

∂y

∂r= sin θ,

∂y

∂θ= r cos θ

∂θ

∂x= −sin θ

r,

∂θ

∂y=

cos θ

r

Γrθθ =∂r

∂x

∂2x

∂θ2+∂r

∂y

∂2y

∂θ2

= cos θ(−r cos θ) + sin θ(−r sin θ)

= −r(cos θ2 + sin θ2) = −r

Γθrθ = Γθθr =∂θ

∂x

∂2x

∂θ∂r+∂θ

∂y

∂2y

∂θ∂r

= −sin θ

r(− sin θ) +

cos θ

r(cos θ)

=1

r

Page 71: LectureNotes On GeneralRelativity

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 λ, ie.xµ = xµ(λ). λ is invariant and chosen to be the curve length.

The tangent vector field of the curve is ~u = (dxµ/dλ)~eµ. The curve passesthrough a vector field ~A. The covariant directional derivative of the vector fieldalong the curve is defined as:

∇~u ~A =d ~A

dλ≡ Aµ;ν

dxν

dλ~eµ = Aµ;νu

ν~eµ (4.27)

The vectors in the vector field are said to beconnected by parallel transport along the curveif

Aµ;νuν = 0

λ)A(

B

)∆λA +λ(

u

λ+∆λ

Q

P

A( λ+∆λ)

λ

Figure 4.1: Parallel transport from P to Q. The vector ~B = Aµ;νuν∆λ~eµ

~u =dxµ

dλ~eµ (4.28)

According to the geometrical interpretation of Levi-Civita, the covariant direc-tional derivative is:

∇~u ~A = Aµ;νuν~eµ = lim

∆λ→0

~A‖(λ+ ∆λ)− ~A(λ)

∆λ(4.29)

where ~A‖(λ+ ∆λ) means the vector ~A parallel transported from Q to P .

Page 72: LectureNotes On GeneralRelativity

56 Chapter 4. Covariant Differentiation

4.3 Geodesic curvesDefinition 4.3.1 (Geodesic curves)A geodesic curve is defined in such a way that,the vectors of the tangent vector fieldof 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(λ) are connected by parallell transport by a dis-placement along a vector ~u , we have Aµ;νuν = 0. For geodesic curves, we thenhave:

uµ;νuν = 0 (4.30)

which is the geodesic equation.

(uµ,ν + Γµανuα)uν = 0 (4.31)

Then we are using that ddλ ≡ dxν

dλ∂∂xν = uν ∂

∂xν :

duµ

dλ= uν

∂uµ

∂xν= uνuµ,ν (4.32)

The geodesic equation can also be written as:

duµ

dλ+ Γµανu

αuν = 0 (4.33)

Usual notation: ˙ = ddλ

uµ =dxµ

dλ= xµ (4.34)

xµ + Γµαν xαxν = 0 (4.35)

By comparing eq.4.35 with the equation of motion(4.53) for a free particle (whichwe 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 Γxtt = (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:

x+ (1 +gx

c2)gt2 = 0

Page 73: LectureNotes On GeneralRelativity

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 betweentwo 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 λ(proper time τ is used for particles with a rest mass).

The Lagrange-function is a function of coordinates and their derivatives,

L = L(xµ, xµ), xµ ≡ dxµ

dλ. (4.36)

(Note: if λ = τ then xµ are the 4-velocity components)The action-integral is S =

∫L(xµ, xµ)dλ. 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 keepP1 and P2 rigid, ie.

δ

∫ λ2

λ1

L(xµ, xµ)dλ = 0, (4.37)

where λ1 and λ2 are the parameter-values at P1 and P2. For all the variations

P

P

1

2

ct

x

Figure 4.2: Different world-lines connecting P1 and P2 in a Minkowski diagram

the following condition applies:

δxµ(λ1) = δxµ(λ2) = 0 (4.38)

We write Eq. (4.37) as

δ

∫ λ2

λ1

Ldλ =

∫ λ2

λ1

[∂L

∂xµδxµ +

∂L

∂xµδxµ]dλ (4.39)

Page 74: LectureNotes On GeneralRelativity

58 Chapter 4. Covariant Differentiation

Partial integration of the last term∫ λ2

λ1

∂L

∂xµδxµdλ =

[∂L

∂xµδxµ]λ2

λ1

−∫ λ2

λ1

d

(∂L

∂xµ

)δxµdλ (4.40)

Due to the conditions δxµ(λ1) = δxµ(λ2) = 0 the first term becomes zero. Thenwe have :

δS =

∫ λ2

λ1

[∂L

∂xµ− d

(∂L

∂xµ

)]δxµdλ (4.41)

The world-line the particle follows is determined by the condition δS = 0 forany variation δxµ. Hence, the world-line of the particle must be given by

∂L

∂xµ− d

(∂L

∂xµ

)= 0 (4.42)

These are the covariant Euler-Lagrange equations.The canonical momentum pµ conjugated to a coordinate xµ is defined as

pµ ≡∂L

∂xµ(4.43)

The Lagrange-equations can now be written as

dpµdλ

=∂L

∂xµor pµ =

∂L

∂xµ. (4.44)

A coordinate which the Lagrange-function does not depend on is known as acyclic coordinate. Hence, ∂L

∂xµ = 0 for a cyclic coordinate. From this follows:

The canonical momentum conjugated to a cycliccoordinate is a constant of motion

ie. pµ = C (constant) if xµ is cyclic.A free particle in space-time (curved space-time includes gravitation) has

the Lagrange function

L =1

2~u · ~u =

1

2xµx

µ =1

2gµν x

µxν (4.45)

An integral of the Lagrange-equations is obtained readily from the 4-velocityidentity:

xµxµ = −c2 for a particle with rest-mass

xµxµ = 0 for light

(4.46)

The line-element is:

ds2 = gµνdxµdxν = gµν x

µxνdλ2 = 2Ldλ2 . (4.47)

Thus the Lagrange function of a free particle is obtained from the line-element.

Page 75: LectureNotes On GeneralRelativity

4.5 Application of the Lagrangian formalism to free particles 59

flat surface:

P

Q

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

sphere:

P

Q

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

4.5 Application of the Lagrangian formalism to freeparticles

To describe the motion of a free particle, we start by setting up the line elementof the space-time in the chosen coordinate system. There are coordinates onwhich the metric does not depend. For example, given axial symmetry we maychoose the angle θ which is a cyclic coordinate here and the conjugate (covariant)impulse Pθ is a constant of the motion (the orbital spin of the particle). If, inaddition, the metric is time independent (stationary metric) then t is alsocyclic 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. Examplesof static metrics are Minkowski and hyperbolically accelerated frames. Therotating cylindrical coordinate system is stationary.

Page 76: LectureNotes On GeneralRelativity

60 Chapter 4. Covariant Differentiation

4.5.1 Equation of motion from Lagrange’s equation

The Lagrange function for a free particle is:

L =1

2g µν x

µxν(4.48)

where g µν = g µν(xλ). And the Lagrange equations are

∂L

∂xβ− d

(∂L

∂xβ

)= 0,

∂L

∂xβ= g βν x

ν ,

∂L

∂xβ=

1

2g µν,βx

µxν .

(4.49)

d

(∂L

∂xβ

)≡(∂L

∂xβ

)•= g βν x

ν + g βν xν

= g βν,µxµxν + g βν x

ν .

(4.50)

Now, (4.50) and (4.49) together give:

1

2g µν,β x

µxν − g βν,µxµxν − g βν xν = 0. (4.51)

The second term on the left hand side of (4.51) may be rewritten making use ofthe fact that xµxν is symmetric in µν, as as follows

g βν,µxµxν =

1

2(g βµ,ν + g βν,µ)xµxν

⇒ g βν xν +

1

2(g βµ,ν + g βν,µ − g µν,β)xµxν = 0.

(4.52)

Finally, since we are free to multiply (4.52) through by gαβ , we can isolate xα

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

xα + Γαµν xµxν = 0 (4.53)

where the Christoffel symbols Γαµν in (4.53) are defined by

Γαµν ≡1

2gαβ(g βµ,ν + g βν,µ − g µν,β). (4.54)

Equation(4.53) describes a geodesic curve .

Page 77: LectureNotes On GeneralRelativity

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 theyare drawn in a Minkowski diagram which refers to an inertial reference frame.

P

O

non-geodetic curve between O and P

X

geodetic curve

t

t

0

1

v(t)

cT

Figure 4.5: Timelike curves in spacetime.

The general interpretation of the line-element for a time-like interval is; Thespacetime distance between O and P (See figure 4.5) equals the proper timeinterval 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 = −c2dτ2 (4.55)

which gives

τ0−1 =

∫ T1

T0

√1− v2(T )

c2dT (4.56)

We can see that τ0−1 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. Considerthe last expression in Section 3.2 of the propertime interval of a particle with positionx and velocity v in a gravitational field with acceleration of gravity g.

dτ = dt

√(1 +

gx

c2)2 − v2

c2

Page 78: LectureNotes On GeneralRelativity

62 Chapter 4. Covariant Differentiation

This expression shows that the proper time of the particle proceeds faster the higherup in the field the particle is, and it proceeds slower the faster the particle moves.Consider figure 4.6. The path a free particle follows between the events O and Pis a compromize between moving as slowly as possible in space, in order to keepthe velocity dependent time dilation small, and moving through regions high up inthe gravitational field, in order to prevent the slow proceeding of proper time fardown. However if the particle moves too high up, its velocity becomes so large thatit proceeds slower again. The compromise between kinematic and gravitational timedilation which gives maximal proper time between O and P is obtained for the thickcurve in fig. 4.6. This is the curve followed by a free particle between the events Oand 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.

τ =

∫ τ1

0

√−gµν xµxνdτ

is maximal for a geodesic curve. However the action

J = −2

∫ τ1

0Ldτ = −

∫ τ1

0gµν x

µxνdτ

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

L =1

2

(−(

1 +gx

c2

)2t2 +

x2

c2

)(4.57)

Using the Euler-Lagrange equations now gives

x+ (1 +gx

c2)gt2 = 0

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 givesstraight paths in 3-space, in accordance with Newton’s 1st law. However projectingit into an accelerated frame where the particle also has a horizontal motion, andtaking 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 P1

and P2. The dashed line is the geodesic for the non-rotating disc. The other curveis a geodesic for the 3-space of a rotating reference frame. We can see that thegeodesic is curved inward when the disc is rotating. The curve has to curve inwardsince the measuring rods are longer there (because of Lorentz-contraction). Thus,the minimum distance between P1 and P2 will be achieved by an inwardly bentcurve.

Page 79: LectureNotes On GeneralRelativity

4.5 Application of the Lagrangian formalism to free particles 63

The path of the particle

O P

g

Figure 4.6: The particle moves between two events O and P at fixed points intime. The path chosen by the particle between O and P is such that the propertime 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 standardclocks goes as fast as possible. If the particle follows the horizontal line betweenO and P it goes as slowly as possible and the kinematical time dilation is assmall as possible. Then the slowing down of its comoving standard clocks due tothe kinematical time dilation is as small as possible, but the particle is far downin the gravitational field and its proper time goes slowly for that reason. Pathsfurther up leads to a greater rate of proper time. But above the curve drawn asa thick line, the kinematical time dilation will dominate, and the proper timeproceeds more slowly.

Page 80: LectureNotes On GeneralRelativity

64 Chapter 4. Covariant Differentiation

We will show this mathematically, using the Lagrangian equations. The lineelement for the space dt = dz = 0 of the rotating reference frame is

2

ω

P1

P

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

dl2 = dr2 +r2dθ2

1− r2ω2

c2

Lagrangian function:

L =1

2r2 +

1

2

r2θ2

1− r2ω2

c2

We will also use the identity:

r2 +r2θ2

1− r2ω2

c2

= 1 (4.58)

(We got this from using ~u · ~u = 1) We see that θ is cyclic ( ∂L∂θ = 0), implying:

pθ =∂L

∂θ=

r2θ

1− r2ω2

c2

= constant

This gives:

θ =(1− r2ω2

c2)pθr2

=pθr2− ω2pθ

c2(4.59)

Inserting 4.59 into 4.58:

r2 = 1 +ω2p2

θ

c2− p2

θ

r2(4.60)

This gives us the equation of the geodesic curve between P1 and P2:

r

θ= ±dr

dθ=r2

√1 +

ω2p2θ

c2− p2

θr2

pθ(1− r2ω2

c2

) (4.61)

Page 81: LectureNotes On GeneralRelativity

4.5 Application of the Lagrangian formalism to free particles 65

Boundary conditions:

2

θ

r

geodesic

r0

P1

P

Figure 4.8: Geodesic curves on a rotating disc,coordinates

r = 0, r = r0, for θ = 0

Inserting this into 4.60 gives:

pθr0

=

1 +p2θω

2

c2(4.62)

Rearranging 4.61,using 4.62 gives:

dr

r√r2 − r2

0

− ω2

c2rdr√r2 − r2

0

=dθ

r0

Integrating this yields:

θ = ±r0ω2

c2

√r2 − r2

0 ∓ arccosr0

r

Example 4.5.3 (Christoffel symbols in a hyperbolically accelerated reference frame)The Christoffel symbols were defined in Equation (4.53).

Γαµν ≡1

2gαβ(g βµ,ν + g βν,µ − g µν,β).

In this example

gtt = −(

1 +gx

c2

)2c2, gxx = gyy = gzz = 1

Page 82: LectureNotes On GeneralRelativity

66 Chapter 4. Covariant Differentiation

x

ct

Pg

O

non-geodetic curve between O and Pgeodetic curve

2

1

HIGHLOW

Figure 4.9: Vertical throw in the accelerated referenceframe.

and only the term ∂gtt∂x contributes to Γαµν . Thus the only non-vanishing Christoffel

symbols are

Γtxt = Γttx =1

2gtt(∂gtt∂x

)

=1

2gtt

∂gtt∂x

=2(1 + gx

c2

)g

2(1 + gx

c2

)2c2

=1(

1 + gxc2

) gc2

Γxtt = −1

2gxx

(∂gtt∂x

)

= −1

2

−2(

1 +gx

c2

) g

c2c2

=(

1 +gx

c2

)g

Page 83: LectureNotes On GeneralRelativity

4.5 Application of the Lagrangian formalism to free particles 67

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

ds2 = −(

1 +gx

c2

)2c2dt2 + dx2 + dy2 + dz2 (4.63)

Vertical motion implies that dy = dz = 0 and the Lagrange function becomes

L =1

2g µν x

µxν

= −1

2

(1 +

gx

c2

)2c2 t2 +

1

2x2

where the dots imply differentiation w.r.t the particle’s proper time, τ . And theinitial conditions are:

x(0) = 0, x(0) = (u0, ux, 0, 0)

= γ(c, v, 0, 0),

where, γ =(1− v2/c2

)−1/2.

What is the maximum height, h reached by the particle?Newtonian description: 1

2mv2 = mgh⇒ h = v2

2g .Relativistic description: t is a cyclic coordinate ⇒ x0 = ct is cyclic and p0 =constant.

p0 =∂L

∂x0=

1

c

∂L

∂t= −c

(1 +

gx

c2

)2t (4.64)

Now the 4-velocity identity is

~u · ~u = g µν xµxν = −c2 (4.65)

so

−1

2

(1 +

gx

c2

)2c2 t2 +

1

2x2 = −1

2c2 (4.66)

and given that the maximum height h is reached when x = 0 we get(

1 +gh

c2

)2

t2x=h = 1. (4.67)

Now, since p0 is a constant of the motion, it preserves its initial value throughoutthe flight (i.e. p0 = −ct(0) = −γc) and particularly at x = h,

(4.64)⇒ p0 = −γc = −c(

1 +gh

c2

)2

tx=h (4.68)

Finally, dividing equation (4.67) by equation (4.68) and substituting back in equation(4.67) gives

h =c2

g(γ − 1) (4.69)

Page 84: LectureNotes On GeneralRelativity

68 Chapter 4. Covariant Differentiation

In the Newtonian limit (4.69) becomes

h =c2

g

(1

(1− v2/c2)1/2− 1

)uc2

g

(1 +

1

2

v2

c2− 1

)⇒ h u

v2

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(γ = 1/0.6). The trip takes 5 years out and 5 years back. This means that Eli, whoremains at Earth is 10 years older when she meets Eva at the end of her journey.Eva, on the other hand, is 10(1 − v2/c2)1/2 = 10(0.6) = 6 years older.

T

EvaEli

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 stationaryand Eli as the one whom undertakes the long journey. In this picture it seems thatEva 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 referenceframes and review the twin “paradox” in this light.

Eva’s 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/γ = 2.4light years. The Earth and Eli travel withv = 0.8c. Her travel time in one direction is then 2.4light years

0.8c = 3 years. So theround trip takes 6 years according to Eva. That is Eva is 6 years older when theymeet again. This is in accordance with the result arrived at by Eli. According to

Page 85: LectureNotes On GeneralRelativity

4.5 Application of the Lagrangian formalism to free particles 69

Eva, Eli ages by only 6 years ×1/γ = 3.6 years during the round trip, not 10 yearsas 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 gravitationalforce directed away from the Earth. Eli is higher up in this gravitational field andages faster than Eva, because of the gravitational time dilation. We assume thatEva has constant proper acceleration and is stationary in a hyperbolically acceleratedframe as she turns about.

The canonical momentum pt for Eli is then(see Equation (4.64))

pt = −(

1 +gx

c2

)2ct

Inserting this into the 4-velocity identity gives

p2t − c2

(1 +

gx

c2

)2=(

1 +gx

c2

)2x2, (4.70)

or

dτ =1 + gx

c2√p2t − c2

(1 + gx

c2

)2 dx

Now, since x = 0 for x = x2 (x2 is Eli’s turning point according to Eva), wehave that

pt = c(

1 +gx2

c2

)

Let x1 be Eli’s position according to Eva just as Eva begins to notice the gravitationalfield. That is when Eli begins to slow down in Eva’s frame.

Integration from x1 to x2 and inserting the value of pt gives

τ1−2 =c

g

√(1 +

gx2

c2

)2−(

1 +gx1

c2

)2

⇒ limg→∞

τ1−2 =1

c

√x2

2 − x21.

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

limg→∞

τ1−2 = 3.2 years

Eli’s aging as she turns about is, according to Eva,

∆τEli = 2 limg→∞

τ1−2 = 6.4 years.

So Eli’s has aged by a total of τEli = 3.6 + 6.4 = 10 years, according to Eva, whichis just what Eli herself found.

Page 86: LectureNotes On GeneralRelativity

70 Chapter 4. Covariant Differentiation

4.5.3 Gravitational Doppler effect

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

~P = E(1, ~w) (c = 1) (4.71)

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

~U · ~P = −E (4.72)

The energy of a particle with 4-momentum ~P measured by an observer with4-velocity ~U is

E = −~U · ~P (4.73)

Let ES = −(~U · ~P )S and Ea = −(~U · ~P )a be the energy of a photon, measuredlocally by observers in rest in the transmitter and receiver positions, respectively.This gives1

ES

(~U · ~P )S=

Ea

(~U · ~P )a(4.74)

Let the angular frequency of the light, measured by the transmitter and receiver,be ws and wa, respectively. We then have

ws =ESh, wa =

Eah, (4.75)

which gives:

wa =(~U · ~P )a

(~U · ~P )sws (4.76)

For an observer at rest in a time-independent orthogonal metric we have

~U · ~P = U tPt =dt

dτPt (4.77)

where Pt is a constant of motion (since t is a cyclic coordinate) for photons andhence has the same value in transmitter and receiver positions. The line-elementis

ds2 = gttdt2 + gii(dx

i)2 (4.78)

Using the physical interpretation (4.55) of the line-element for a time like inter-val, we obtain for the proper time of an observer at rest

dτ2 = −gttdt2 ⇒ dτ =√−gttdt (4.79)

1 ~A · ~B = A0B0 + A1B

1 + ... = g00A0B0 + g11A

1B1 + ..., an orthonormal basis gives~A · ~B = −A0B0 +A1B1 + ...

Page 87: LectureNotes On GeneralRelativity

4.6 The Koszul connection 71

Hencedt

dτ=

1√−gtt, (4.80)

which gives~U · ~P =

1√−gttPt. (4.81)

Inserting this into the expression for angular frequency (4.76) gives

wa =

√(gtt)s(gtt)a

ws (4.82)

Note: we have assumed an orthogonal and time-independent metric, i.e. Pt1 =Pt2 . Inserting the metric of a hyperbolically accelerated reference system with

gtt = −(1 +gx

c2)2 (4.83)

gives

wa =1 + gxs

c2

1 + gxac2ws, (4.84)

orwa −wsws

=1 + gxs

c2

1 + gxac2− 1 =

gc2

(xs − xa)1 + gxa

c2≈ g

c2H, (4.85)

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

w=

200

9× 1016= 2.2 × 10−15.

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:

∇~uf ≡ ~u(f) (4.86)

Here the vector ~u should be taken as a differensial operator. (In coordinatebasis, ~u = uµ ∂

∂xµ )The directional derivative along a basis vector ~eν is written as:

∇ν ≡ ∇~eν (4.87)

Hence ∇µ( ) = ∇~eµ( ) = ~eµ( )

Page 88: LectureNotes On GeneralRelativity

72 Chapter 4. Covariant Differentiation

Definition 4.6.1 (Koszul’s connection coeffecients in an arbitrary basis)In an aribitrary basis the Koszul connection coefficients are defined by

∇ν~eµ ≡ Γαµν~eα (4.88)

which may also be written ~eν(~eµ) = Γαµν~eα. In coordinate basis , Γαµν is reducedto Christoffel symbols and one often writes ~eµ,ν = Γαµν~eα. In an arbitrary basis, Γαµν has no symmetry.

Example 4.6.1 (The connection coefficients in a rotating reference frame.)Coordinate transformation: (T,R,Θ are coordinates in the non-rotating referenceframe, t, r, θ in the rotating.) Corresponding Cartesian coordinates:X,Y and x, y.

t = T, r = R, θ = Θ− ωTX = R cos Θ, Y = R sin Θ

X = r cos(θ + ωt), Y = r sin(θ + ωt)

X

Y

t

y

xP (x,y)(X,Y)

ωtθ+ω

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

~et =∂

∂t=∂X

∂t

∂X+∂Y

∂t

∂Y+∂T

∂t

∂T

Page 89: LectureNotes On GeneralRelativity

4.6 The Koszul connection 73

gives:

~et = −rω sin(θ + ωt)~eX + rω cos(θ + ωt)~eY + ~eT

~er =∂X

∂r

∂X+∂Y

∂r

∂Y= cos(θ + ωt)~eX + sin(θ + ωt)~eY

~eθ =∂X

∂θ

∂X+∂Y

∂θ

∂Y= −r sin(θ + ωt)~eX + r cos(θ + ωt)~eY

We are going to find the Christoffel symbols, which involves differentiation of ba-sis vectors. This coordinate transformation makes this easy, since ~eX , ~eY , ~eT areconstant. Differentiation gives

∇t~et = −rω2 cos(θ + ωt)~eX − rω2 sin(θ + ωt)~eY (4.89)

The connection coefficients are (see eq. 4.88)

∇ν~eµ ≡ Γαµν~eα (4.90)

So, to calculate Γαµν , the right hand side of eq.4.89 has to by expressed by thebasis that we are differentiating.By inspection, the right hand side is −rω2~er.That is ∇t~et = −rω2~er giving Γrtt = −rω2.The other nonzero Christoffel symbols are

Γθrt = Γθtr =ω

r,Γθθr = Γθrθ =

1

rΓrθt = Γrtθ = rω,Γrθθ = −r

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

~r = ~v = (vi + Γijkvjvk)~ei,

where · ≡ ddt . i, j, and k are space indices. Inserting the Christoffel symbols for

plane polar coordinates (see example 4.2.1), gives:

~ainert = (r − rθ2)~er + (θ +2

rrθ)~eθ

Example 4.6.3 (The acceleration of a particle, relative to the rotating reference frame)Inserting the Christoffel symbols from example 4.6.1:

~arot = (r − rθ2 − rω2 + Γrθtθt+ Γrtθ tθ)~er + (θ +2

rrθ + Γθrtrt+ Γθtr tr)~eθ

= (r − rθ2 − rω2 − 2rωθ)~er + (rθ + 2rθ + 2rω)~eθ

= ~ainert − (rω2 − 2rωθ)~er + 2rω~eθ

Page 90: LectureNotes On GeneralRelativity

74 Chapter 4. Covariant Differentiation

The angular velocity of the reference frame, is ~ω = ω~ez. We also introduce ~r = r~er.The velocity relative to the rotating reference frame is then:

~r = r~er + r~er

Furthermore

~er =d~erdt

=∂~er∂xi

dxi

dt= vi~er,i

Using definition 4.6.1 in a coordinate basis, this may be written

~er = viΓjri~ej

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

~er = vθΓθrθ~eθ = θ1

r~eθ = θ~eθ

Hence~v = ~r = r~er + rθ~eθ

Inserting this into the expression for the acceleration, gives:

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

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

4.7 Connection coefficients Γαµν and structure coeffi-cients cαµν in a Riemannian (torsion free) space

The commutator of two vectors, ~u and ~v, expressed by covariant directionalderivatives is given by:

[~u,~v] = ∇~u~v −∇~v~u (4.91)

Let ~u = ~eµ and ~v = ~eν . We then have:

[ ~eµ, ~eν ] = ∇µ ~eν −∇ν ~eµ. (4.92)

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

cαµν ~eα = (Γανµ − Γαµν) ~eα (4.93)

Thus in a torsion free space

cαµν = Γανµ − Γαµν (4.94)

In coordinate basis we have

~eµ =∂

∂xµ, ~eν =

∂xν(4.95)

Page 91: LectureNotes On GeneralRelativity

4.8 Covariant differentiation of vectors, forms and tensors 75

And therefore:

[ ~eµ, ~eν ] = [∂

∂xµ,∂

∂xν]

=∂

∂xµ(∂

∂xν)− ∂

∂xν(∂

∂xµ)

=∂2

∂xµ∂xν− ∂2

∂xν∂xµ= 0

(4.96)

Equation (4.96) shows that cαµν = 0, and that the connection coefficients inEquation (4.94) therefore are symmetrical in a coordinate basis:

Γανµ = Γαµν (4.97)

4.8 Covariant differentiation of vectors, forms and ten-sors

4.8.1 Covariant differentiation of a vector in an arbitrary basis

∇ν ~A = ∇ν(Aµ ~eµ)

= ∇νAµ ~eµ +Aα∇ν ~eα(4.98)

∇νAµ = ~eν(Aµ) , ~eν = Mµν

∂xµ, (4.99)

where Mµν are the elements of a transformation matrix between a coordinate

basis ∂∂xµ and an arbitrary basis ~eν. ( If ~eν had been a coordinate basis

vector, we would have gotten ~eν(Aµ) = ∂∂xν (Aµ) = Aµ,ν).

∇ν ~A = [~eν(Aµ) +AαΓµαν ] ~eµ (4.100)

Definition 4.8.1 (Covariant derivative of a vector)The covariant derivative of a vector in an arbitrary basis is defined by:

∇ν ~A ≡ Aµ;ν~eµ (4.101)

So:Aµ;ν = ~eν(Aµ) +AαΓµαν

where ∇ν~eα ≡ Γµαν~eµ(4.102)

4.8.2 Covariant differentiation of formsDefinition 4.8.2 (Covariant directional derivative of a one-form field)Given a vector field ~A and a one-form field α, the covariant directional derivative ofα in the direction of the vector ~u is defined by:

(∇~uα)( ~A ) ≡ ∇~u[α( ~A )︸ ︷︷ ︸αµAµ

]− α(∇~u ~A) (4.103)

Page 92: LectureNotes On GeneralRelativity

76 Chapter 4. Covariant Differentiation

Let α = ωµ (basis form), ωµ(~eν) ≡ δµν and let ~A = ~eν and ~u = ~eλ. We thenhave:

(∇λωµ)(~eν) = ∇λ[ωµ(~eν)︸ ︷︷ ︸δµν

]− ωµ(∇λ ~eν) (4.104)

The covariant directional derivative ∇λ of a constant scalar field is zero, ∇λδµν =0. We therefore get:

(∇λωµ)(~eν) = −ωµ(∇λ ~eν)

= −ωµ(Γανλ ~eα)

= −Γανλωµ( ~eα)

= −Γανλδµα

= −Γµνλ

(4.105)

The contraction between a one-form and a basis vector gives the componentsof the one-form, α(~eν) = αν . Equation (4.105) tells us that the ν-component of∇λωµ is equal to −Γµνλ, and that we therefore have

∇λωµ = −Γµνλων (4.106)

Equation (4.106) gives the directional derivatives of the basis forms. Using theproduct of differentiation gives

∇λα = ∇λ(αµωµ)

= ∇λ(αµ)ωµ + αµ∇λωµ= ~eλ(αµ)ωµ − αµΓµνλω

ν

(4.107)

Definition 4.8.3 (Covariant derivative of a one-form)The covariant derivative of a one-form α = αµω

µ is therefore given by Equation(4.108) below, when we let µ→ ν in the first term on the right hand side in (4.107):

∇λα = [ ~eλ(αν)− αµΓµνλ]ων (4.108)

The covariant derivative of the one-form components αµ are denoted by αν;λ andare defined by

∇λα ≡ αν;λων (4.109)

It then follows that

αν;λ = ~eλ(αν)− αµΓµνλ (4.110)

It is worth to note that Γµνλ in Equation (4.110) are not Christoffel symbols. Incoordinate basis we get:

αν;λ = αν,λ − ανΓµλν (4.111)

where Γµλν = Γµνλ are Christoffel symbols.

Page 93: LectureNotes On GeneralRelativity

4.9 The Cartan connection 77

4.8.3 Generalization for tensors of higher rankDefinition 4.8.4 (Covariant derivative of a tensor)Let A and B be two tensors of arbitrary rank. The covariant directional derivativealong a basis vector ~eα of a tensor A⊗B of arbitrary rank is defined by:

∇λ(A⊗B) ≡ (∇λA)⊗B +A⊗ (∇λB) (4.112)

We will use (4.112) to find the formula for the covariant derivative of the com-ponents of a tensor of rank 2:

∇αS = ∇α(Sµνωµ ⊗ ων)

= (∇αSµν)ωµ ⊗ ων + Sµν(∇αωµ)⊗ ων + Sµνωµ ⊗ (∇αων)

= (Sµν,α − SβνΓβµα − SµβΓβνα)ωµ ⊗ ων(4.113)

where Sµν,α = ~eα(Sµν). Defining the covariant derivative Sµν;α by

∇αS = Sµν;αωµ ⊗ ων (4.114)

we get

Sµν;α = Sµν,α − SβνΓβµα − SµβΓβνα (4.115)

For the metric tensor we get

gµν;α = gµν,α − gβνΓβµα − gµβΓβνα (4.116)

From

gµν = ~eµ · ~eν (4.117)

we get:

gµν,α = (∇α ~eµ) · ~eν + ~eµ(∇α ~eν)

= Γβµα ~eβ · ~eν + ~eµ · Γβνα ~eβ= gβνΓβµα + gµβΓβνα

(4.118)

This means thatgµν;α = 0 (4.119)

So the metric tensor is a (covariant) constant tensor.

4.9 The Cartan connectionDefinition 4.9.1 (Exterior derivative of a basis vector)

d~eµ ≡ Γνµα~eν ⊗ ωα (4.120)

Page 94: LectureNotes On GeneralRelativity

78 Chapter 4. Covariant Differentiation

Exterior derivative of a vector field:

d ~A = d(~eµAµ) = ~eν ⊗ dAν +Aµd~eµ (4.121)

In arbitrary basis:dAν = ~eλ(Aν)ωλ (4.122)

(In coordinate basis, ~eλ(Aν) = ∂∂xλ

(Aν) = Aν,λ)giving:

d ~A = ~eν ⊗ [~eλ(Aν)ωλ] +AµΓνµλ~eν ⊗ ωλ

= (~eλ(Aν) +AµΓνµλ)~eν ⊗ ωλ(4.123)

d ~A = Aν;λ~eν ⊗ ωλ (4.124)

Definition 4.9.2 (Connection forms Ωνµ)The connection forms Ων

µ are 1-forms, defined by:

d~eµ ≡ ~eν ⊗ Ωνµ

Γνµα~eν ⊗ ωα = ~eν ⊗ Γνµαωα = ~eν ⊗ Ων

µ

(4.125)

Ωνµ = Γνµαω

α (4.126)

The exterior derivatives of the components of the metric tensor:

dgµν = d(~eµ · ~eν) = ~eµ · d~eν + ~eν · d~eµ (4.127)

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 = Aµ

ν~eµ ⊗ ωνis defined by:

~u · A ≡ uαAµν(~eα · ~eµ)ων (4.128)

Using this definition, we get:

dgµν = (~eµ · ~eλ)Ωλν + (~eν · ~eγ)Ωγ

µ

= gµλΩλν + gνγΩγ

µ

(4.129)

Page 95: LectureNotes On GeneralRelativity

4.9 The Cartan connection 79

Lowering an index gives

dgµν = Ωµν + Ωνµ (4.130)

In an orthonormal basis field there is Minkowski-metric:

gµν = ηµν (4.131)

which is constant. Then we have :

dgµν = 0⇒ Ωνµ = −Ωµν (4.132)

where we write Ωνµ = Γνµαωα. It follows that Γνµα = −Γµνα.

It also follows that

Γtij

= −Γtij = Γitj = Γitj

Γijk

= −Γjik

(4.133)

Cartans 1st structure equation (without proof):

dωρ = −1

2cρµνω

µ ∧ ων

= −1

2(Γρνµ − Γρµν)ωµ ∧ ων

= −Γρνµωµ ∧ ων

= −Ωρν ∧ ων

(4.134)

dωρ = −Ωρν ∧ ων and dωρ = Γρµνω

µ ∧ ων (4.135)

In coordinate basis, we have ωρ = dxρ.Thus, dωρ = d2xρ = 0.We also have cρµν = 0 , and C1 is reduced to an identity.This formalism cannot be used in coordinate basis!

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

ds2 = dr2 + r2dθ2

Introducing basis forms in an orthonormal basis field (where the metric is g rr =gθθ = 1):

ds2 = grrωr ⊗ ωr + gθθω

θ ⊗ ωθ = ωr ⊗ ωr + ωθ ⊗ ωθ

⇒ ωr = dr, ωθ = rdθ

Exterior differentiation gives:

dωr = d2r = 0, dωθ = dr ∧ dθ =1

rωr ∧ ωθ

Page 96: LectureNotes On GeneralRelativity

80 Chapter 4. Covariant Differentiation

C1:

dωµ = −Ωµν ∧ ων

= −Ωµr ∧ ωr − Ωµ

θ∧ ωθ

We have that dωr = 0 , which gives:

Ωrθ

= Γrθθωθ (4.136)

since ωθ ∧ ωθ = 0. (Ωrr=0 because of the antisymmetry Ωνµ = −Ωµν .)

We also have: dωθ = −1rω

θ ∧ ωr. C1:

dωθ = −Ωθr ∧ ωr − Ωθ

θ︸︷︷︸=0

∧ωθ

Ωθr = Γθ

rθωθ + Γθrrω

r (4.137)

giving Γθrθ

= 1r .

We have: Ωrθ

= −Ωθr. Using equations 4.136 and 4.137 we get:

Γθrr = 0

⇒ Γrθθ

= −1

r

giving Ωrθ

= −Ωθr = −1

rωθ.

Page 97: LectureNotes On GeneralRelativity

Chapter 5

Curvature

5.1 The Riemann curvature tensor

*

M

1

P

Q

λ

λ+ ∆λ

~A(λ)

~AQP (λ+ ∆λ)

∇~v ~A∆λ

~v

~A(λ+ ∆λ)

Figure 5.1: Parallel transport of ~A

The covariant directional derivative of a vector field ~A along a vector ~u wasdefined and interpreted geometrically in section 4.2, as follows

∇~v ~A =d ~A

dλ= Aµ;νv

ν~eµ

= lim∆λ→0

~AQP (λ+ ∆λ)− ~A(λ)

∆λ

(5.1)

Let ~AQP be the parallel transported of ~A from Q to P. Then to first order in∆λ we have: ~AQP = ~AP + (∇~v ~A)P∆λ and

81

Page 98: LectureNotes On GeneralRelativity

82 Chapter 5. Curvature

-~A

6

~A‖

:

3

Figure 5.2: Parallel transport of a vector along a triangle of angles 90 is rotated90

~APQ = ~AQ − (∇~v ~A)Q∆λ (5.2)

To second order in ∆λ we have:

~APQ = (1−∇~v∆λ+1

2∇~v∇~v(∆λ)2) ~AQ (5.3)

If ~APQ is parallel transported further on to R we get

~APQR = (1−∇~u∆λ+1

2∇~u∇~u(∆λ)2)

· (1−∇~v∆λ+1

2∇~v∇~v(∆λ)2) ~AR

(5.4)

where ~AQ is replaced by ~AR because the differential operator always shall beapplied to the vector in the first position. If we parallel transport ~A around thewhole polygon in figure 5.3 we get:

~APQRSTP = (1 +∇~u∆λ+1

2∇~u∇~u(∆λ)2)

· (1 +∇~v∆λ+1

2∇~v∇~v(∆λ)2)

· (1−∇[~u,~v](∆λ)2) · (1−∇~u∆λ+1

2∇~u∇~u(∆λ)2)

· (1−∇~v∆λ+1

2∇~v∇~v(∆λ)2) ~AP

(5.5)

Page 99: LectureNotes On GeneralRelativity

5.1 The Riemann curvature tensor 83

-

3

]1

-

6

j

I IO1

Qλ+ ∆λ

R

S

T

(∇~u∆λ~v −∇~v∆λ~u)∆λ= [~u,~v]∆λ2

∇~v∆λ~u∆λ

∇~u∆λ~v∆λ

~u(P )∆λ

~AP

~uPQ∆λ

~APQ

~AQ∇~v ~A

~v(P )∆λ

~vPT∆λ

~v(T )∆λ

~u(Q)∆λ

Figure 5.3: Geometrically implied curvature from non-zero differences betweenvectors along a curve (parameterized by λ) and their parallel transported equiv-alents

Calculating to 2. order in ∆λ gives:

~APQRSTP = ~AP + ([∇~u,∇~v]−∇[~u,~v])(∆λ)2 ~AP (5.6)

There is a variation of the vector under parallel transport around the closedpolygon:

δ ~A = ~APQRSTP − ~AP = ([∇~u,∇~v ]−∇[~u,~v]) ~AP (∆λ)2 (5.7)

We now introduce the Riemann’s curvature tensor as:

R( , ~A, ~u,~v) ≡ ([∇~u,∇~v]−∇[~u,~v])( ~A) (5.8)

The components of the Riemann curvature tensor is defined by applying thetensor on basis vectors,

Rµναβ~eµ ≡ ([∇α,∇β ]−∇[~eα,~eβ ])(~eν) (5.9)

Anti-symmetry follows from the definition:

Rµνβα = −Rµναβ (5.10)

The expression for the variation of ~A under parallel transport around the poly-

Page 100: LectureNotes On GeneralRelativity

84 Chapter 5. Curvature

gon, Eq. (5.7), can now be written as:

δ ~A = R( , ~A, ~u,~v)(∆λ)2

= R( , Aν~eν , uα~eα, v

β~eβ)(∆λ)2

= ~eµRµναβA

νuαvβ · (∆λ)2

=1

2~eµR

µναβA

ν(uαvβ − uβvα)(∆λ)2

(5.11)

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

-

6

~v∆λ

~u∆λ

∆~S = ~u× ~v(∆λ)2

∆S

∆~S = ~u× ~v(∆λ)2.

Using that(~u× ~v)αβ = uαvβ − uβvα .

we can write Eq. (5.11) as:

δ ~A =1

2AνRµναβ∆Sαβ~eµ . (5.12)

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

~eµRµναβ = [∇α,∇β]~eν −∇[~eα,~eβ ]~eν

= (∇α∇β −∇β∇α − cραβ∇ρ)~eν= ∇α∇β~eν −∇β∇α~eν − cραβ∇ρ~eν

(Kozul-connection) = ∇αΓµνβ~eµ −∇βΓµνα~eµ − cραβΓµνρ~eµ

= (∇αΓµνβ)~eµ + Γµνβ∇α~eµ− (∇βΓµνα)~eµ − Γµνα∇β~eµ − cραβΓµνρ~eµ

= ~eα(Γµνβ)~eµ + ΓρνβΓµρα~eµ

− ~eβ(Γµνα)~eµ − ΓρναΓµρβ~eµ − cραβΓµνρ~eµ .

(5.13)

This gives (in arbitrary basis):

Rµναβ = ~eα(Γµνβ)− ~eβ(Γµνα)

+ ΓρνβΓµρα − ΓρναΓµρβ − cραβΓµνρ .

(5.14)

Page 101: LectureNotes On GeneralRelativity

5.1 The Riemann curvature tensor 85

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

Rµναβ = Γµνβ,α − Γµνα,β + ΓρνβΓµρα − ΓρναΓµρβ , (5.15)

where Γµνβ = Γµβν are the Christoffel symbols.

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

Rµν =1

2Rµναβω

α ∧ ωβ (5.16)

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

Rµν = (~eα(Γµνβ) + ΓρνβΓµρα −1

2cραβΓµνρ)ω

α ∧ ωβ (5.17)

The connection forms:

Ωµν = Γµναω

α (5.18)

Exterior derivatives of basis forms:

dωρ = −1

2cραβω

α ∧ ωβ (5.19)

Exterior derivatives of connection forms (C1: dωρ = −Ωρα ∧ ωα) :

dΩµν = dΓµνβ ∧ ωβ + Γµνρdω

ρ

= ~eα(Γµνβωα ∧ ωβ − 1

2cραβΓµνρω

α ∧ ωβ(5.20)

The curvature forms can now be written as:

Rµν = dΩµν + Ωµ

λ ∧ Ωλν (5.21)

This is Cartans 2nd structure equation.

Page 102: LectureNotes On GeneralRelativity

86 Chapter 5. Curvature

5.2 Differential geometry of surfaces

λ

µe

N

u

Figure 5.4: The geometry of a surface. We see the normal vector and the unitvectors 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 =∂

∂u,~ev =

∂v(5.22)

where u and v are coordinates on the surface.Line element on the surface:

ds2 = gµνdxµdxν (5.23)

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

~eµ,ν = Γαµν~eα +Kµν~N, α = 1, 2 (5.24)

Greek indices run through the surface coordinates, ~N is a unit vector orthogonalto the surface.

Page 103: LectureNotes On GeneralRelativity

5.2 Differential geometry of surfaces 87

The equation above is called Gauss’ equation. We have: Kµν = ~eµ,ν · ~N . Incoordinate basis, we have ~eµ,ν = ∂2

∂xµ∂xν = ∂2

∂xν∂xµ = ~eν,µ. It follows that

Kµν = Kνµ (5.25)

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

d~u

dλ= uµ;νu

ν~eµ + Kµνuµuν︸ ︷︷ ︸

2nd fundamental form

~N (5.26)

We define κg and κN by:d~u

dλ= κg~e+ κN ~N (5.27)

κg is called geodesic curvature. κN is called normal curvature (external curva-ture). κg = 0 for geodesic curves on the surface.

κg~e = uµ;νuν~eµ = ∇~u~u

κN = Kµνuµuν

And :κN =d~u

dλ· ~N

(5.28)

We also have that ~u · ~N = 0 along the whole curve. Differentiation gives:

d~u

dλ· ~N + ~u · d

~N

dλ= 0 (5.29)

gives:

κN = −~u · d~N

dλ(5.30)

which is called Weingarten’s equation.κg and κN together give a complete description of the geometry of a surface

in a flat 3 dimensional space. We are now going to consider geodesic curvesthrough a point on the surface. Tangent vector ~u = uµ~eµ with ~u ·~u = gµνu

µuν =1. Directions with maximum and minimum values for the normal curvatures arefound, by extremalizing κN under the condition gµνuµuν = 1. We then solve thevariation problem δF = 0 for arbitrary uµ, where F = Kµνu

µuν−k(gµνuµuν−1).

Here k is the Lagrange multiplicator. Variation with respect to uµ gives:

δF = 2(Kµν − kgµν)uνδuµ

δF = 0 for arbitrary δuµ demands:

(Kµν − kgµν)uν = 0 (5.31)

For this system of equations to have nonzero solutions, we must have:

det(Kµν − kgµν) = 0 (5.32)

Page 104: LectureNotes On GeneralRelativity

88 Chapter 5. Curvature

∣∣∣∣K11 − kg11 K12 − kg12

K21 − kg21 K22 − kg22

∣∣∣∣ = 0 (5.33)

This gives the following quadratic equation for k:

k2det(gµν)− (g11K22 − 2g12K12 + g22K11)k + det(Kµν) = 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 uµ:

0 = (Kµν − kgµν)uµuν

= Kµνuµuν − kgµνuµuν

= κN − k ⇒ k = κN

(5.35)

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

Kµνuν = kgµνu

ν (5.36)

We then get:

Kµνuνvµ = k1gµνu

νvµ

= k1uµvµ = k1(~u · ~v)

Kµνvνuµ = k2gµνv

νuµ = k2(~u · ~v)

gives (k1 − k2)(~u · ~v) = Kµν(uνvµ − vνuµ)

= 2Kµνu[νvµ]

(5.37)

Kµν is symmetric in µ and ν. So we get:(k1 − k2)(~u · ~v) = 0. For k1 6= k2 wehave to demand ~u ·~v = 0. So the geodesics with extremal normal curvature, areorthogonal to each other.

The Gaussian curvature (at a point) is defined as:

K = κN1 · κN2 (5.38)

Since κN1 and κN2 are solutions of the quadratic equation above, we get:

K =det(Kµν)

det(gµν)(5.39)

Page 105: LectureNotes On GeneralRelativity

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. Greekindices run through the surface coordinates (two dimensional) and Latin indicesthrough the space coordinates (three dimensional):

~ea = (~e1, ~e2, ~N) , ~eµ = ~e1, ~e2 (5.40)

Using the exterior derivative and form formalism, we find how the unit vectorson the surface change:

d~eν = ~ea ⊗ Ωaν

= ~eα ⊗ Ωαν + ~N ⊗ Ω3

ν ,(5.41)

where Ωµν = Γµναωα are the connection forms on the surface, i.e. the intrinsic

connection forms. The extrinsic connection forms are

Ω3ν = Kναω

α ,Ωµ3 = Kµ

αωα (5.42)

We let the surface be embedded in an Euclidean (flat) 3-dimensional space. Thismeans that the curvature forms of the 3-dimensional space are zero:

Ra3b = 0 = d Ωab + Ωa

k ∧ Ωkb (5.43)

which gives:

Rµ3ν = 0 = d Ωµν + Ωµ

α ∧ Ωαν + Ωµ

3 ∧ Ω3ν

= Rµν + Ωµ3 ∧ Ω3

ν ,(5.44)

where Rµν are the curvature forms of the surface. We then have:1

2Rµναβω

α ∧ ωβ = −Ωµ3 ∧ Ω3

ν (5.45)

Inserting the components of the extrinsic connection forms, we get: (when weremember the anti symmetry of α and β in Rµ

ναβ)

Rµναβ = KµαKνβ −Kµ

βKνα (5.46)

We now lower the first index:

Rµναβ = KµαKνβ −KµβKνα (5.47)

Rµναβ are the components of a curvature tensor which only refer to the dimen-sions of the surface. In particular we have:

R1212 = K11K22 −K12K21 = detK (5.48)

We then have the following connection between this component of the Riemanncurvature tensor of the surface and the Gaussian curvature of the surface:

K = κN1 · κN2 =detKµν

det gµν=

R1212

det gµν(5.49)

Since the right hand side refers to the intrinsic curvature and the metric on thesurface, we have proved that the Gaussian curvature of a surface is an intrinsicquantity. It can be measured by observers on the surface without embedding thesurface in a three-dimensional space. This is the contents of Gauss’ theoremaegregium.

Page 106: LectureNotes On GeneralRelativity

90 Chapter 5. Curvature

5.3 The Ricci identity

~eµRµναβA

ν = (∇α∇β −∇β∇α −∇[ ~eα, ~eβ ])(~A) (5.50)

In coordinate basis this is reduced to

~eµRµναβA

ν = (Aµ;βα −Aµ;αβ) ~eµ , (5.51)

where

Aµ;αβ ≡ (Aµ;β);α (5.52)

The Ricci identity on component form is:

AνRµναβ = Aµ;βα −Aµ;αβ (5.53)

We can write this as:

d2 ~A =1

2RµναβA

ν ~eµ ⊗ ωα ∧ ωβ (5.54)

This shows us that the 2nd exterior derivative of a vector is equal to zero onlyin a flat space. Equations (5.53) and (5.54) both represents the Ricci identity.

5.4 Bianchi’s 1st identity

Cartan’s 1st structure equation:

d ωµ = −Ωµν ∧ ων (5.55)

Cartan’s 2nd structure equation:

Rµν = d Ωµν + Ωµ

λ ∧ Ωλν (5.56)

Exterior differentiation of (5.55) and use of Poincare′s lemma (4.16) gives:(d2 ωµ = 0)

0 = d Ωµν ∧ ων − Ωµ

λ ∧ d ωλ (5.57)

Use of (5.55) gives:

d Ωµν ∧ ων + Ωµ

λ ∧ Ωλν ∧ ων = 0 (5.58)

From this we see that

(d Ωµν + Ωµ

λ ∧ Ωλν) ∧ ων = 0 (5.59)

We now get Bianchi’s 1st identity:

Rµν ∧ ων = 0 (5.60)

Page 107: LectureNotes On GeneralRelativity

5.5 Bianchi’s 2nd identity 91

On component form Bianchi’s 1st identity is

1

2Rµναβω

α ∧ ωβ︸ ︷︷ ︸

Rµν

∧ων = 0 (5.61)

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

Rµ[ναβ] = 0 (5.62)

or

Rµναβ +Rµαβν +Rµβνα+ = 0 (5.63)

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

5.5 Bianchi’s 2nd identity

Exterior differentiation of (5.56) ⇒

d Rµν = Rµλ ∧Ωλν − Ωµ

ρ ∧ Ωρλ ∧ Ωλ

ν − Ωµλ ∧Rλν + Ωµ

λ ∧Ωλρ ∧ Ωρ

ν

= Rµλ ∧Ωλν − Ωµ

λ ∧Rλν(5.64)

We now have Bianchi’s 2nd identity as a form equation:

d Rµν + Ωµλ ∧Rλν −R

µλ ∧Ωλ

ν = 0 (5.65)

As a component equation Bianchi’s 2nd identity is given by

Rµν[αβ;γ] = 0 (5.66)

Definition 5.5.1 (Contraction)‘Contraction’ is a tensor operation defined by

Rνβ ≡ Rµνµβ (5.67)

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

The tensor with components Rνβ is called the Ricci curvature tensor.Another contraction gives the Ricci curvature scalar, R = Rµ

µ.

Page 108: LectureNotes On GeneralRelativity

92 Chapter 5. Curvature

Riemann curvature tensor has four symmetries. The definition of the Rie-mann tensor implies that Rµναβ = −RµνβαBianchi’s 1st identity: Rµ[ναβ] = 0

From Cartan’s 2nd structure equation follows

R µν = dΩ µν + Ωµλ ∧ Ωλν

⇒ R µναβ = −R νµαβ

(5.68)

By choosing a locally Cartesian coordinate system in an inertial frame we getthe following expression for the components of the Riemann curvature tensor:

R µναβ =1

2(gµβ,να − gµα,νβ + gνα,µβ − gνβ,µα) (5.69)

from which it follows that Rµναβ = Rαβµν . Contraction of µ and α leads to:

Rαναβ = Rαβαν

⇒ Rνβ = Rβν(5.70)

i.e. the Ricci tensor is symmetric. In 4-D the Ricci tensor has 10 independentcomponents.

Page 109: LectureNotes On GeneralRelativity

Chapter 6

Einstein’s Field Equations

6.1 Energy-momentum conservation

6.1.1 Newtonian fluid

Energy-momentum conservation for a Newtonian fluid in terms of the divergenceof the energy momentum tensor can be shown as follows. The total derivativeof a velocity field is

D~v

Dt≡ ∂~v

∂t+ (~v · ~∇)~v (6.1)

∂~v∂t is the local derivative which gives the change in ~v as a function of time

at a given point in space. (~v · ~∇)~v is called the convective derivative of ~v. Itrepresents the change of ~v for a moving fluid particle due to the inhomogeneityof the fluid velocity field. In component notation the above become

Dvi

Dt≡ ∂vi

∂t+ vj

∂vi

∂xj(6.2)

The continuity equation

∂ρ

∂t+∇ · (ρ~v) = 0 or

∂ρ

∂t+∂(ρvi)

∂xi= 0 (6.3)

Euler’s equation of motion (ignoring gravity)

ρD~v

Dt= −~∇p or ρ

(∂vi

∂t+ vj

∂vi

∂xj

)= − ∂p

∂xi(6.4)

The energy momentum tensor is a symmetric tensor of rank 2 thatdescribes material characteristics.

T µν =

T 00 T 01 T 02 T 03

T 10 T 11 T 12 T 13

T 20 T 21 T 22 T 23

T 30 T 31 T 32 T 33

(6.5)

c ≡ 1

93

Page 110: LectureNotes On GeneralRelativity

94 Chapter 6. Einstein’s 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 = ρ T i0 = ρvi

T ij = ρvivj + pδij(6.6)

where p is pressure, assumed isotropic here. We choose a locally Cartesian coordinatesystem in an inertial frame such that the covariant derivatives are reduced to partialderivatives. The divergence of the momentum energy tensor, T µν;ν has 4 components,one for each value of µ.

The zeroth component is

T 0ν;ν = T 0ν

,ν = T 00,0 + T 0i

,i

=∂ρ

∂t+∂(ρvi)

∂xi

(6.7)

which by comparison to Newtonian hydrodynamics implies that T 0ν;ν = 0 is the

continuity equation. This equation represents the conservation of energy.The ith component of the divergence is

T iν,ν = T i0,0 + T ij,j

=∂(ρvi)

∂t+∂(ρvivj + pδij)

∂xj

= ρ∂vi

∂t+ vi

∂ρ

∂t+ vi

∂ρvj

∂xj+ ρvj

∂vi

∂xj+

∂p

∂xi

(6.8)

now, according to the continuity equation

∂(ρvi)

∂xi= −∂ρ

∂t

⇒ T iν,ν = ρ∂vi

∂t+ vi

∂ρ

∂t− vi∂ρ

∂t+ ρvj

∂vi

∂xj+

∂p

∂xi

= ρDvi

Dt+

∂p

∂xi

∴ T iν;ν = 0⇒ ρDvi

Dt= − ∂p

∂xi

(6.9)

which is Euler’s equation of motion. It expresses the conservation of momentum.

Page 111: LectureNotes On GeneralRelativity

6.2 Einstein’s curvature tensor 95

The equations T µν;ν = 0 are general expressions for energy and momentumconservation.

6.1.2 Perfect fluids

A perfect fluid is a fluid with no viscosity and is given by the energy-momentumtensor

Tµν = (ρ+p

c2)uµuν + pgµν (6.10)

where ρ and p are the mass density and the stress, respectively, measured in thefluids rest frame, uµ are the components of the 4-velocity of the fluid.

In a comoving orthonormal basis the components of the 4-velocity are uµ =(c, 0, 0, 0). Then the energy-momentum tensor is given by

Tµν =

ρc2 0 0 00 p 0 00 0 p 00 0 0 p

(6.11)

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-momemtumtensor Tµν = ρuµuν .

2. radiation or ultra-relativistic gas is given by a traceless energy-momemtumtensor, i.e. T µµ = 0. It follows that p = 1

3ρc2.

3. vacuum energy: If we assume that no velocity can be measured relativelyto vacuum, then all the components of the energy-momentum tensor mustbe Lorentz-invariant. It follows that Tµν ∝ gµν . If vacuum is defined as aperfect fluid we get p = −ρc2 so that Tµν = pgµν = −ρc2gµν .

6.2 Einstein’s 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 asa consequence of the field equation. This puts the following constraints on thecurvature tensor: It must be a symmetric, divergence free tensor of rank 2.

Bianchi’s 2nd identity:

Rµναβ;σ +Rµνσα;β +Rµνβσ;α = 0 (6.12)

contraction of µ and α ⇒Rµνµβ;σ −R

µνµσ;β +Rµνβσ;µ = 0

R νβ;σ −R νσ;β +Rµνβσ;µ = 0(6.13)

Page 112: LectureNotes On GeneralRelativity

96 Chapter 6. Einstein’s Field Equations

further contraction of ν and σ gives

Rσβ;σ −Rσσ;β +Rσµσβ;µ = 0

Rσβ;σ −R ;β +Rσβ;σ = 0

∴ 2Rσβ;σ = R ;β

(6.14)

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

Rσβ;σ =1

2R ;β (6.15)

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.

Rσβ;σ −1

2R ;β = 0 (6.16)

Keeping in mind that (gσβR);σ = gσβR;σ we multiply (6.16) by gβα to get

gβαRσβ;σ − gβα

1

2R ;β = 0

(gβαR

σβ

);σ− 1

2

(gβαR

);β

= 0(6.17)

interchanging σ and β in the first term of the last equation:(gσαR

βσ

);β− 1

2

(gβαR

);β

= 0

⇒(Rβα −

1

2δβαR

)

= 0(6.18)

since gσαRβσ=δσαR

βσ=Rβα. So that Rβα − 1

2δβαR is the divergence free curvature

tensor desired.This tensor is called the Einstein tensor and its covariant components are

denoted by Eαβ . That is

Eαβ = R αβ −1

2g αβR (6.19)

NOTE THAT: Eµν;ν = 0 → 4 equations, giving only 6 equations from E µν ,

which secures a free choice of coordinate system.

6.3 Einstein’s field equations

Einstein’s field equations:

Eµν = κTµν (6.20)

Page 113: LectureNotes On GeneralRelativity

6.3 Einstein’s field equations 97

or

Rµν −1

2gµνR = κTµν (6.21)

Contraction gives:

R− 1

24R = κT , where T ≡ T µµR = −κT

(6.22)

Rµν =1

2gµν(−κT ) + κTµν , (6.23)

Thus the field equations may be written in the form

Rµν = κ(Tµν −1

2gµνT ) (6.24)

In the Newtonian limit the metric may be written

ds2 = −(

1 +2φ

c2

)dt2 + (1 + hii)(dx

2 + dy2 + dz2) (6.25)

where the Newtonian potential |φ| c2. We also have T00 Tkk and T ≈ −T00.Then the 00-component of the field equations becomes

R00 ≈κ

2T00 (6.26)

Furthermore we have

R00 = Rµ0µ0 = Ri0i0

= Γi00,i − Γi0i,0

=∂Γk00

∂xk=

1

c2∇2φ (6.27)

Since T00 ≈ ρc2 eq.(6.26) can be written ∇2φ = 12κc

4ρ. Comparing this equationwith the Newtonian law of gravitation on local form: ∇2φ = 4πGρ, we see thatκ = 8πG

c4.

In classical vacuum we have : Tµν = 0, which gives

Eµν = 0 or Rµν = 0 . (6.28)

These are the “vacuum field equations”. Note that Rµν = 0 does not implyRµναβ = 0.

Page 114: LectureNotes On GeneralRelativity

98 Chapter 6. Einstein’s Field Equations

Digression 6.3.1 (Lagrange (variation principle))It was shown by Hilbert that the field equations may be deduced from a variationprinciple with action ∫

R√−gd4x , (6.29)

where R√−g is the Lagrange density. One may also include a so-called cosmological

constant Λ: ∫(R + 2Λ)

√−gd4x (6.30)

The field equations with cosmological constant are

Rµν −1

2gµνR+ Λgµν = κTµν (6.31)

6.4 The “geodesic postulate” as a consequence of thefield equations

The principle that free particles follow geodesic curves has been called the“geodesic postulate”. We shall now show that the “geodesic postulate” followsas a consequence of the field equations.

Consider a system of free particles in curved space-time. This system canbe regarded as a pressure-free gas. Such a gas is called dust. It is described byan energy-momentum tensor

T µν = ρuµuν (6.32)

where ρ is the rest density of the dust as measured by an observer at rest in thedust and uµ are the components of the four-velocity of the dust particles.

Einstein’s field equations as applied to space-time filled with dust, take theform

Rµν − 1

2gµνR = κρuµuν (6.33)

Because the divergence of the left hand side is zero, the divergence of the righthand side must be zero, too

(ρuµuν);ν = 0 (6.34)

or

(ρuνuµ);ν = 0 (6.35)

we now regard the quantity in the parenthesis as a product of ρuν and uµ. Bythe rule for differentiating a product we get

(ρuν);νuµ + ρuνuµ;ν = 0 (6.36)

Page 115: LectureNotes On GeneralRelativity

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 oflight we have

uµuµ = −c2 (6.37)

Differentiation gives

(uµuµ);ν = 0 (6.38)

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

uµ;νuµ + uµu

µ;ν = 0 (6.39)

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

u µ;νuµ = uµuµ;ν = gµαuαuµ;ν = uαg

µαuµ;ν = uαuα;ν = uµu

µ;ν (6.40)

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

uµuµ;ν = 0 (6.41)

Multiplying eq.(6.36) by uµ, we get

(ρuν);νuµuµ + ρuνuµu

µ;ν = 0 (6.42)

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

(ρuν);ν = 0 (6.43)

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

ρuνuµ;ν = 0 (6.44)

Since ρ 6= 0 we must have

uνuµ;ν = 0 (6.45)

This is just the geodesic equation. Conclusion: It follows from Einstein’sfield equations that free particles move along paths corresponding to geodesiccurves of space-time.

Page 116: LectureNotes On GeneralRelativity

Chapter 7

The Schwarzschild spacetime

7.1 Schwarzschild’s exterior solution

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

ds2 = −e2α(r)dt2 + e2β(r)dr2 + r2dΩ2

dΩ2 = dθ2 + sin2 θdφ2(7.1)

These coordinates are chosen so that the area of a sphere with radius r is 4πr2.Physical distance in radial direction, corresponding to a coordinate distance

dr, is dlr =√grrdr = eβ(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 Eµν = 0) we find

ωt = eα(r)dt , ωr = eβ(r)dr , ωθ = rdθ , ωφ = r sin θdφ (7.2)

2. Computing the connection forms by applying Cartan’s 1. structure equa-tions

dωµ = −Ωµν ∧ ων (7.3)

dωt = eαα′dr ∧ dt= eαα′e−βωr ∧ e−αωt

= −e−βα′ωt ∧ ωr

= −Ωtr ∧ ωr

(7.4)

∴ Ωtr = e−βα′ωt + f1ω

r (7.5)

100

Page 117: LectureNotes On GeneralRelativity

7.1 Schwarzschild’s exterior solution 101

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

Ω µν = −Ω νµ (7.6)

This gives:

Ωrφ

= −Ωφr = −1

re−βωφ

Ωθφ

= −Ωφ

θ= −1

rcot θωφ

Ωtr = +Ωr

t= e−βα′ωt

Ωrθ

= −Ωθr = −1

re−βωθ

(7.7)

4. We then proceed to determine the curvature forms by applying Cartan’s2nd structure equations

Rµν = dΩµν + Ωµ

α ∧ Ωαν (7.8)

which gives:

Rtr = −e−2β(α′′ + α′2 − α′β′)ωt ∧ ωr

Rtθ

= −1

re−2βα′ωt ∧ ωθ

Rtφ

= −1

re−2βα′ωt ∧ ωφ

Rrθ

=1

re−2ββ′ωr ∧ ωθ

Rrφ

=1

re−2ββ′ωr ∧ ωφ

Rθφ

=1

r2(1− e−2β)ωθ ∧ ωφ

(7.9)

5. By applying the following relation

Rµν =1

2Rµναβ

ωα ∧ ωβ (7.10)

we find the components of Riemann’s curvature tensor.

6. Contraction gives the components of Ricci’s curvature tensor

Rµν ≡ Rαµαν (7.11)

7. A new contraction gives Ricci’s curvature scalar

R ≡ Rµµ (7.12)

Page 118: LectureNotes On GeneralRelativity

102 Chapter 7. The Schwarzschild spacetime

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

Eµν = Rµν −1

2ηµνR , (7.13)

where ηµν = diag(−1, 1, 1, 1). We then have:

Ett =2

re−2ββ′ +

1

r2(1− e−2β)

Err =2

re−2βα′ − 1

r2(1− e−2β)

Eθθ = Eφφ = e−2β(α′′ + α′2 − α′β′ + α′

r− β′

r)

(7.14)

We want to solve the equations Eµν = 0. We get only 2 independentequations, and choose to solve those:

Ett = 0 and Err = 0 (7.15)

By adding the 2 equations we get:

Ett +Err = 0

⇒ 2

re−2β(β′ + α′) = 0

⇒ (α+ β)′ = 0⇒ α+ β = K1 (const) (7.16)

We now have:ds2 = −e2αdt2 + e2βdr2 + r2dΩ2 (7.17)

By choosing a suitable coordinate time, we can achieve

K1 = 0⇒ α = −β

Since we have ds2 = −e2αdt2 + e−2αdr2 + r2dΩ2, this means that grr =− 1gtt

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

2

re−2ββ′ +

1

r2(1− e−2β) = 0

This equation can be written:

1

r2

d

dr[r(1− e−2β)] = 0

∴ r(1− e−2β) = K2 (const)(7.18)

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

ds2 = −dt2 + dr2 + r2dΩ2 , (7.19)

Page 119: LectureNotes On GeneralRelativity

7.1 Schwarzschild’s exterior solution 103

which is the Minkowski space-time described in spherical coordinates. Ingeneral, K2 6= 0 and 1− e−2β = K2

r ≡ Kr , giving

e2α = e−2β = 1− K

r

and

ds2 = −(1− K

r)dt2 +

dr2

1− Kr

+ r2dΩ2 (7.20)

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 atrest ) in the limit of a weak field of a particle at a distance r from a sphericalmass M . Newtonian:

g =d2r

dt2= −GM

r2(7.21)

We anticipate that r >> K. Then the proper time τ of a particle will beapproximately equal to the coordinate time, since dτ =

√1− K

r dt

The acceleration of a particle in 3-space, is given by the geodesic equation:

d2xµ

dτ2+ Γµαβu

αuβ = 0

uα =dxα

(7.22)

For a particle instantanously at rest in a weak field, we have dτ ≈ dt. Usinguµ = (1, 0, 0, 0), we get:

g =d2r

dt2= −Γrtt (7.23)

This equation gives a physical interpretation of Γrtt as the gravitational acceler-ation. This is a mathematical way to express the principle of equivalence: Thegravitational acceleration can be transformed to 0, since the Christoffel symbolsalways can be transformed to 0 locally, in a freely falling non-rotating frame,i.e. a local inertial frame.

Γrtt =1

2grα︸︷︷︸

1grα

(∂gαt∂t︸ ︷︷ ︸=0

+∂gαt∂t︸ ︷︷ ︸=0

−∂gtt∂xα

)

= − 1

2grr

∂gtt∂r

gtt = −(1− K

r) ,

∂gtt∂r

= −Kr2

g = −Γrtt = − K

2r2= −GM

r2

gives K = 2GM

or with c: K =2GM

c2

(7.24)

Page 120: LectureNotes On GeneralRelativity

104 Chapter 7. The Schwarzschild spacetime

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

ds2 = −(1− 2GM

c2r)c2dt2 +

dr2

1− 2GMc2r

+ r2dΩ2 (7.25)

RS ≡ 2GMc2

is the Schwarzschild radius of a mass M.Weak field: r >> RS .For the earth: RS ∼ 0.9cmFor the sun: RS ∼ 3kmA standard clock at rest in the Schwarzschild spacetime shows a proper time τ :

dτ =

√1− RS

rdt (7.26)

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 matterwhere they are. If they hadn’t, the spatial distance between simultanous eventswith given spatial coordinates, would depend on the time of the measuring ofthe distance. Then the metric would be time dependent. Time is not runningat 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 componentsof the metric tensor is infinitely large, but where the curvature of spacetime is finite.

Kretschmann’s curvature scalar is RµναβRµναβ . From the Schwarzschild metric,we get:

RµναβRµναβ =

48G2M2

r8(7.27)

which diverges only at the origin. Since there is no physical singularity atr = 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 descriptionof Schwarzschild spacetime to include the area inside RS).

7.2 Radial free fall in Schwarzschild spacetime

The Lagrangian function of a particle moving radially in Schwarzschild space-time

L = −1

2(1− RS

r)c2 t2 +

1

2

r2

(1− RSr )

, · ≡ d

dτ(7.28)

Page 121: LectureNotes On GeneralRelativity

7.2 Radial free fall in Schwarzschild spacetime 105

where τ 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.

pt =∂L

∂t= −(1− RS

r)c2 t (7.29)

4-velocity identity: uµuµ = −c2:

−(1− RSr

)c2 t2 +r2

1− RSr

= −c2 (7.30)

Inserting the expression for t gives:

r2 − p2t

c2= −(1− RS

r)c2 (7.31)

Boundary conditions: the particle is falling from rest at r = r0.

pt = −(1− RSr0

)c2√

1− RSr0︸ ︷︷ ︸

t(r=r0)

= −√

1− RSr0c2 (7.32)

giving

r =dr

dτ= −c

√RSr0

√r0 − rr

(7.33)∫

dr√r0−rr

= −c√RSr0τ (7.34)

Integration with τ = 0 for r = 0 gives:

τ = −r0

c

√r0

RS(arcsin

√r

r0−√

r

r0

√1− r

r0) (7.35)

τ is the proper time that the particle spends on the part of the fall which isfrom r to r=0. To the lowest order in r

r0we get:

τ = −2

3

√r

RS

r

c(7.36)

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

|τ(RS)| = 2

3

RSc

= 4× 10−6s (7.37)

Page 122: LectureNotes On GeneralRelativity

106 Chapter 7. The Schwarzschild spacetime

7.3 Light cones in Schwarzschild spacetime

The Schwarzschild line-element (with c = 1) is

ds2 = −(1− RSr

)dt2 +dr2

(1− RSr )

+ r2dΩ2 (7.38)

We will look at radially moving photons (ds2 = dΩ2 = 0). We then get

dr√1− RS

r

= ±√

1− RSrdt⇔ r

12 dr√r −RS

= ±√r −RSr

12

dt

rdr

r −RS= ±dt

(7.39)

with + for outward motion and − for inward motion. For inwardly movingphotons, integration yields

r + t+RS ln | rRS− 1| = k = constant (7.40)

We now introduce a new time coordinate t′ such that the equation of motionfor photons moving inwards takes the following form

r + t′ = k ⇒ dr

dt′= −1

∴ t′ = t+RS ln | rRS− 1|

(7.41)

The coordinate t′ is called an ingoing Eddington-Finkelstein coordinate. Thephotons here always move with the local velocity of light, c. For photons movingoutwards we have

r +RS ln | rRS− 1| = t+ k (7.42)

Making use of t = t′ −RS ln | rRS − 1| we get

r + 2RS ln | rRS− 1| = t′ + k

⇒ dr

dt′+

2RSr −RS

dr

dt′= 1⇔ r +RS

r −RSdr

dt′= 1

⇔ dr

dt′=r −RSr +RS

(7.43)

Making use of ordinary Schwarzschild coordinates we would have gotten thefollowing coordinate velocities for inn- and outwardly moving photons:

dr

dt= ±(1− RS

r) (7.44)

which shows us how light is decelerated in a gravitational field. Figure 7.1 showshow this is viewed by a non-moving observer located far away from the mass. InFigure 7.2 we have instead used the alternative time coordinate t′. The specialtheory of relativity is valid locally, and all material particles thus have to remaininside the light cone.

Page 123: LectureNotes On GeneralRelativity

7.3 Light cones in Schwarzschild spacetime 107

S

hori

zon

RS

r

Trajectory of transmitter

t

?

Light cone

Collaps of light cone

at the horizon, r = R

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

t’

hori

zon

RS

r

Trajectory of transmitter

Figure 7.2: Using the ingoing Eddingto n-Finkelstein time coordinate there is nocollapse of the light cone at r = RS . Instead we get a collapse at the singularityat r = 0. The angle between the left part of the light cone and the t′-axis isalways 45 degrees. We also see that once the transmitter gets inside the horizonat r = RS , no particles can escape.

Page 124: LectureNotes On GeneralRelativity

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 holethere exist a horizon at the Schwarzschild radius. As we have seen in section7.3 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 atthe Schwarzschild radius. The curvature of spacetime is finite here.

Kruskal and Szekeres have introduced new coordinates that are well definedinside as well as outside the Schwarzschild radius, and with the property thatthe metric tensor is non-singular for all r > 0.

In order to arrive at these coordinates we start by considering a photonmoving radially inwards. From eq. (7.40) we then have

t = −r −RS ln

∣∣∣∣r

RS− 1

∣∣∣∣+ v (7.45)

where v is a constant along the world line of the photon. We introduce a newradial coordinate

r∗ ≡ r +RS ln

∣∣∣∣r

RS− 1

∣∣∣∣ (7.46)

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 thephoton 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 coordinatefor outgoing photons. The coordinates u and v are the generalization of the lightcone coordinates of Minkowski spacetime to the Schwarzchild spacetime.

From eqs. (7.47) and (7.48) we get

dt =1

2(dv + du) (7.49)

dr∗ =1

2(dv − du) (7.50)

and from eq. (7.46)

dr =

(1− Rs

r

)dr∗ (7.51)

Inserting these differentials into eq. (7.38) we arrive at a new form of theSchwarzschild line-element,

ds2 = −(

1− Rsr

)du dv + r2dΩ2 (7.52)

Page 125: LectureNotes On GeneralRelativity

7.4 Analytical extension of the Schwarzschild spacetime 109

The metric is still not well behaved at the horizon. Introducing the coordinates

U = −e−u

2Rs (7.53)

V = ev

2Rs (7.54)

gives

UV = −ev−u2Rs = −e

r∗Rs = −

∣∣∣∣Rsr− 1

∣∣∣∣ erRs (7.55)

anddu dv = −4R 2

s

dUdV

UV(7.56)

The line-element (7.52) then takes the form

ds2 = −4R 3s

re−

rRs dUdV + r2dΩ2 (7.57)

This is the first form of the Kruskal-Szekeres line-element. Here is no coordinatesingularity, only a physical singularity at r = 0.

We may furthermore introduce two new coordinates

T =1

2(V + U) =

(r

Rs− 1

)e

r2Rs sinh

t

2Rs(7.58)

Z =1

2(V − U) =

(r

Rs− 1

)e

r2Rs cosh

t

2Rs(7.59)

Hence

V = T + Z (7.60)U = T − Z (7.61)

givingdUdV = dT 2 − dZ2 (7.62)

Inserting this into eq. (7.57) we arrive at the second form of the Kruskal-Szekeresline-element

ds2 = −4R 3s

re−

rRs

(dT 2 − dZ2

)+ r2dΩ2 (7.63)

The inverse transformations of eqs. (7.58) and (7.59) is(

1− r

Rs

)erRs = T 2 − Z2 (7.64)

tanht

2Rs=T

Z(7.65)

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

Z = ±T + constant (7.66)

Page 126: LectureNotes On GeneralRelativity

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 simultaneityplane (dt = 0) through the equatorial plane (dθ = 0) has the line element

ds2 = grrdr2 + r2dφ2 (7.67)

with a radial coordinate such that a circle with radius r has a circumference oflength 2πr.

We now embed this surface in a flat 3-dimensional space with cylinder co-ordinates (z, r, φ) and line element

ds2 = dz2 + dr2 + r2dφ2 (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

ds2 = [1 + (dz

dr)2]dr2 + r2dφ2 (7.69)

Demanding that (7.69) is in agreement with (7.67) we get

grr = 1 + (dz

dr)2 ⇔ dz

dr= ±

√grr − 1 (7.70)

Choosing the positive solution gives

dz =√grr − 1dr (7.71)

In the Schwarzschild spacetime we have

grr =1

1− RSr

(7.72)

Making use of this we find z:

z =

∫ r

RS

dr√1− RS

r

=√

4RS(r −RS) (7.73)

This is shown in Figure 7.3.

7.6 Deceleration of light

The radial speed of light in Schwarzschild coordinates found by putting ds2 =dΩ2 = 0 in eq. (7.38) is

c = 1− RSr

(7.74)

To measure this effect one can look at how long it takes for light to get fromMercury to the Earth. This is illustrated in Figure 7.4. The travel time from

Page 127: LectureNotes On GeneralRelativity

7.6 Deceleration of light 111

Figure 7.3: Embedding of the Schwarzschild metric.

z

Sunx

b

Mercury

z1 <0

z2 >0r2

r1

Earth

Figure 7.4: General relativity predicts that light traveling from Mercury to theEarth will be delayed due to the effect of the Suns gravity field on the speed oflight. This effect has been measured.

Page 128: LectureNotes On GeneralRelativity

112 Chapter 7. The Schwarzschild spacetime

z1 to z2 is

∆t =

∫ z2

z1

dz

1− RSr

≈∫ z2

z1

(1 +RSr

)dz =

∫ z2

z1

(1 +RS√b2 + z2

)dz

= z2 + |z1|+RS ln

√z2

2 + b2 + z2√z1

2 + b2 − |z1|

(7.75)

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 isthen small. A series expansion to the lowest order of b/z gives

∆t = z2 + |z1|+RS ln4|z1|z2

b2(7.76)

The last term represents the extra traveling time due to the effect of the Sunsgravity field on the speed of light. The journey takes longer time:

RS = the Schwarzschild radius of the Sun ∼ 2km|z1|= the radius of Earth’s orbit = 15× 1010mz2 = the radius of Mercury’s orbit = 5.8× 1010mb = R = 7× 108m

give a delay of 1.1 × 10−4s. In addition to this one must also, of course, takeinto account among other things the effects of the curvature of spacetime nearthe Sun and atmospheric effects on Earth.

7.7 Particle trajectories in Schwarzschild 3-space

L =1

2g µνX

µXν

= −1

2

(1− Rs

r

)t2 +

12 r

2

1− Rsr

+1

2r2θ2 +

1

2r2 sin2 θφ2

(7.77)

Since t is a cyclic coordinate

−pt = −∂L∂t

=

(1− Rs

r

)t = constant = E (7.78)

where E is the particle’s energy as measured by an observer "far away" (r Rs).Also φ is a cyclic coordinate so that

pφ =∂L

∂φ= r2 sin2 θφ = constant (7.79)

where pφ is the particle’s orbital angular momentum.

Page 129: LectureNotes On GeneralRelativity

7.7 Particle trajectories in Schwarzschild 3-space 113

Making use of the 4-velocity identity ~U2 = g µνXµXν = −1 we transform

the above to get

−(

1− Rsr

)t2 +

r2

1− Rsr

+ r2θ2 + r2 sin2 θφ2 = −1 (7.80)

which on substitution for t = E1−Rs

r

and φ =pφ

r2 sin2 θbecomes

− E2

1− Rsr

+r2

1− Rsr

+ r2θ2 +p2φ

r2 sin2 θ= −1 (7.81)

Now, refering back to the Lagrange equation

d

(∂L

∂Xµ

)− ∂L

∂Xµ= 0 (7.82)

we get, for θ

(r2θ)• = r2 sin θ cos θφ2

=p2φ cos θ

r2 sin3 θ

(7.83)

Multiplying this by r2θ we get

(r2θ)(r2θ)• =cos θθ

sin3 θp2φ (7.84)

which, on integration, gives

(r2θ)2 = k −( pφ

sin θ

)2(7.85)

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 agiven time t = 0. That is θ = π

2 and θ = 0 at time t = 0. This determines theconstant of integration and k = p2

φ such that

(r2θ)2 = p2φ

(1− 1

sin2 θ

)(7.86)

The RHS is negative for all θ 6= π2 . It follows that the particle cannot deviate

from its original (equatorial) trajectory. Also, since this particular choice oftrajectory was arbitrary we can conclude, quite generally, that any motion offree particles in a spherically symmetric gravitational field is planar motion.

Page 130: LectureNotes On GeneralRelativity

114 Chapter 7. The Schwarzschild spacetime

7.7.1 Motion in the equatorial plane

− E2

1− Rsr

+r2

1− Rsr

+p2φ

r2= −1 (7.87)

that is

r2 = E2 −(

1− Rsr

)(1 +

p2φ

r2

)(7.88)

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

V 2(r) =

(1− Rs

r

)(1 +

p2φ

r2

)

r2 + V 2(r) = E2

⇒ V =

1− rsr

+p2φ

r2−Rsp

r3

u 1− 1

2

Rsr

+1

2

p2φ

r2

(7.89)

Newtonian potential VN is defined by using the last expression in

VN = V − 1⇒ VN = −GMr

+p2φ

2r2(7.90)

The possible trajectories of particles in the Schwarzschild 3-space are shownschematically in Figure 7.5 as functions of position and energy of the particlein the Newtonian limit.

To take into account the relativistic effects the above picture must be mod-ified. We introduce dimensionless variables

X =r

GMand k =

pφGMm

(7.91)

The potential V 2(r) now take the form

V =

(1− 2

X+k2

X2− 2k2

X3

)1/2

(7.92)

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

V (2) =

√1− 1 +

k2

4− 2k2

8= 0 (7.93)

For k2 < 12 particles will fall in towards r = 0.

Page 131: LectureNotes On GeneralRelativity

7.7 Particle trajectories in Schwarzschild 3-space 115

Figure 7.5: Newtonian particle trajectories are functions of the position andenergy of the particle. Note the centrifugal barrier. Due to this particleswith pφ 6= 0 cannot arrive at r = 0.

An orbit equation is one which connects r and φ. So for motion in theequatorial plane for weaks fields we have

dt=

pφmr2

• ≡ d

dτ=

pφmr2

d

dφ(7.94)

Introducing the new radial coordinate u ≡ 1r our equations transform to

du

dφ= − 1

r2

dr

dφ= − 1

r2

mr2

dr

dt= −m

pφr

⇒ r = −pφm

du

(7.95)

Substitution from above for r in the energy equation yields the orbit equation,(du

)2

+ (1− 2GMu)

(u2 +

m2

p2φ

)=E2

p2φ

. (7.96)

Differentiating this, we find

d2u

dφ2+ u =

GMm2

p2φ

+ 3GMu2 (7.97)

The last term on the RHS is a relativistic correction term.

Page 132: LectureNotes On GeneralRelativity

116 Chapter 7. The Schwarzschild spacetime

20 40 60 80 100

0.9

0.95

1.05

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

7.8 Classical tests of Einstein’s 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 aclock on the ground.

The proper time interval measured on a clock moving with a velocity v i = dxi

dtin an arbitrary coordinate system with metric tensor gµν is given by

dτ = (−gµνc2dxµdxν)

12 , dx0 = cdt

= (−g00 − 2gi0vi

c− v2

c2)

12 dt

v2 ≡ gijvivj

(7.98)

Page 133: LectureNotes On GeneralRelativity

7.8 Classical tests of Einstein’s general theory of relativity 117

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

dτ = (−g00 −v2

c2)

12dt , v2 = gii(v

i)2 (7.99)

We now look at an idealized situation where a plane flies at constant altitudeand with constant speed along the equator.

dτ = (1− RSr− v2

c2)

12 dt , r = R+ h (7.100)

To the lowest order in RSr and v2

c2we get

dτ = (1− RS2r− 1

2

v2

c2)dt (7.101)

The speed of the moving clock is

v = (R+ h)Ω + u (7.102)

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

∆τ = (1− GM

Rc2− 1

2

R2Ω2

c2+gh

c2− 2RΩu+ u2

2c2)∆t , g =

GM

R2−RΩ2

(7.103)

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

∆τ0 = (1− GM

Rc2− 1

2

R2Ω2

c2)∆t (7.104)

To the lowest order the relative difference in travel time is

k =∆τ −∆τ0

∆τ0

∼= gh

c2− 2RΩu+ u2

2c2(7.105)

Measurements:Travel time: ∆τ0 = 1.2 × 105s (a little over 24h)Traveling eastwards: ke = −1.0× 10−12

Traveling westwards: kw = 2.1× 10−12

(∆τ −∆τ0)e = −1.2× 10−7s ≈ −120ns(∆τ −∆τ0)w = 2.5× 10−7s ≈ 250ns

Page 134: LectureNotes On GeneralRelativity

118 Chapter 7. The Schwarzschild spacetime

7.8.2 Mercury’s perihelion precession

The orbit equation for a planet orbiting a star of mass M is given by equation(7.97),

d2u

dφ2+ u =

GMm2

pφ2+ ku2 (7.106)

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

u0 =GMm2

pφ2+ ku0

2 (7.107)

With a small perturbation from the circular motion u is changed by u1, whereu1 u0. To lowest order in u1 we have

d2u1

dφ2+ u0 + u1 =

GMm2

pφ2+ ku0

2 + 2ku0u1 (7.108)

ord2u1

dφ2+ u1 = 2ku0u1 ⇔ d2u1

dφ2+ (1− 2ku0)u1 = 0 (7.109)

For ku0 1 the equilibrium orbit is stable and we get a periodic solution:

u1 = εu0 cos[√

1− 2ku0(φ− φ0)] (7.110)

where ε and φ0 are integration constants. ε is the eccentricity of the orbit. Wecan choose φ0 = 0 and then have

1

r= u = u0 + u1 = u0[1 + ε cos(

√1− 2ku0φ) (7.111)

Let f ≡√

1− 2ku0 ⇒1

r=

1

r0(1 + ε cos fφ) (7.112)

For f = 1 (k = 0, no relativistic term) this expression describes a non-precessingelliptic orbit (a Kepler-orbit).

For f < 1 (k > 0) the ellipse is not closed. To give the same value for r ason a given starting point, φ has to increase by 2π

f > 2π. The extra angle perrotation is 2π( 1

f − 1) = ∆φ1.

∆φ1 = 2π(1√

1− 2ku0− 1) ≈ 2πku0 (7.113)

Using general relativity we get for Mercury

k = 3GM ⇒ ∆φ = 6πGMu0 ≈ 6πGMGMm2

pφ2(7.114)

∆φ = 6π(GMm

pφ)2per orbit. (7.115)

which in Mercury’s case amounts to (∆φ)century = 43′′

Page 135: LectureNotes On GeneralRelativity

7.8 Classical tests of Einstein’s general theory of relativity 119

7.8.3 Deflection of light

The orbit equation for a free particle with mass m = 0 is

d2u

dφ2+ u = ku2 (7.116)

If light is not deflected it will follow the straight line

cosφ =b

r= b u0 (7.117)

where b is the impact parameter of the path. This is the horizontal dashed linein Figure 7.7. The 0’th order solution (7.117) fullfills

d2u0

dφ2+ u0 = 0 (7.118)

Hence it is a solution of (7.116) with k = 0.

photon

φ

2

SunM, GM<<b

br

∆θ

Figure 7.7: Light traveling close to a massive object is deflected.

The perturbed solution is

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

Inserting this into the orbit equation gives

d2u0

dφ2+d2u1

dφ2+ u0 + u1 = ku 2

0 + 2ku0u1 + ku 21 (7.120)

The first and third term at the left hand side cancel each other due to eq. (7.118]and the last term at the right hand side is small to second order in u1 and willbe neglected. Hence we get

d2u1

dφ2+ u1 = ku 2

0 + 2ku0u1 (7.121)

Page 136: LectureNotes On GeneralRelativity

120 Chapter 7. The Schwarzschild spacetime

The last term at the right hand side is much smaller then the first, and will alsobe neglected. Inserting for u0 from (7.117) we then get

d2u1

dφ2+ u1 =

k

b2cos2 φ (7.122)

This equation has a particular solution of the form

u1p = A+B cos2 φ (7.123)

Inserting this into (7.122) we find

A =2k

3b2, B = − k

3b2(7.124)

Hence

u1p =k

3b2(2− cos2 φ

)(7.125)

giving

1

r= u = u0 + u1 =

cosφ

b+

k

3b2(2− cos2 φ

)(7.126)

The deflection of the light ∆θ is assumed to be small. We therefore put φ =n2 + ∆θ

2 where ∆θ π, (see Figure 7.7). Hence

cosφ = cos

(n

2+

∆θ

2

)= − sin

∆θ

2≈ ∆θ

2(7.127)

Thus, the term cos2 φ in (7.126) can be neglected. Furthermore, the deflectionof the light is found by letting r→∞, i. e. u→ 0. Then we get

∆θ =4k

3b(7.128)

For motion in the Schwarzschild spacetime outside the Sun, k = 32RS where RS

is the Schwarzschild radius of the Sun. And for light passing the surface of theSun b = R where R is the actual radius of the Sun. The deflection is then

∆θ = 2RSR

= 1.75′′ (7.129)

Page 137: LectureNotes On GeneralRelativity

Chapter 8

Black Holes

8.1 ’Surface gravity’:gravitational acceleration on thehorizon of a black hole

Surface gravity is denoted by κ1 and is defined by

κ = limr→r+

a

uta =

√aµaµ (8.1)

where r+ is the horizon radius, r+ = RS for the Schwarzschild spacetime, ut isthe time component of the 4-velocity.

The 4-velocity of a free particle instantanously at rest in the Schwarzschildspacetime:

~u = ut~et =dt

dτ~et =

1√−gtt~et =

~et√1− RS

r

(8.2)

The only component of the 4-acceleration different from zero, is ar. The4-acceleration:~a = ∇~u~u = uµ;νuν~eµ = (uµ,ν + Γµανuα)uν~eµ.

ar = (ur,ν + Γrανuα)uν

= ur,νuν

︸ ︷︷ ︸=0

+Γrtt(ut)2

=Γrtt

1− RSr

Γrtt = −1

2

∂gtt∂r

= −RS2r2

ar =RS2r2

1− RSr

ar = grrar =argrr

= (1− RSr

)ar =RS2r2

(8.3)

The acceleration scalar: a =√arar =

RS2r2q1−RS

r

(measured with standard instru-

121

Page 138: LectureNotes On GeneralRelativity

122 Chapter 8. Black Holes

ments: at the horizon, time is not running).

a

ut=RS2r2

(8.4)

With c:a

ut=c2RS2r2

=GM

r2(8.5)

κ = limr→RS

a

ut=

1

2RS=

1

4GM(8.6)

Including c the expression is κ = c2

4GM . On the horizon of a black hole with onesolar mass, we get κ = 2× 1013 m

s2 .

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 spectrumhas an intensity maximum at a wavelength given by Wien’s displacement law.

Λ =N~ckT

where k is the Boltzmann constant, and N=0.2014

For radiation emitted from a black hole, Hawking derived the following expres-sion for the wavelength at a maximum intensity

Λ = 4πNRS =8πNGM

c2(8.7)

Inserting Λ from Wien’s displacement law, gives:

T =~c3

8πGkM=~c

2πkκ (8.8)

Inserting values for ~, c and k gives:

T ≈ 2× 10−4m

RSK (8.9)

For a black hole with one solar mass,we have T ≈ 10−7. When the mass isdecreasing because of the radiation, the temperature is increasing.So a blackhole has a negative heat capacity. The energy loss of a black hole because ofradiation, is given by the Stefan-Boltzmann law:

−dMdt

= σT 4 A

c2(8.10)

where A is the surface of the horizon.

A = 4πR2S =

16πG2M2

c4(8.11)

Page 139: LectureNotes On GeneralRelativity

8.3 Rotating Black Holes: The Kerr metric 123

gives:

−dMdt

=1

15360π

~c6

G2M2≡ Q

M2

M(t) = (M 30 − 3Qt)1/3, M0 = M(0)

(8.12)

A black hole with mass M0 early in the history of the universe which is aboutto explode now, had to have a starting mass

M0 = (3Qt0)1/3 ≈ 1012kg (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 metricallows rotation.

Consider a stationary metric which describes a axial-symmetric space

ds2 = −e2νdt2 + e2µdr2 + e2ψ(dφ− ωdt)2 + e2λdθ2 , (8.14)

where ν, µ, ψ, λ and ω are functions of r and θ.By solving the vacuum field equations for this line-element, Kerr found the

solution:

e2ν =ρ2∆

Σ2, e2µ =

ρ2

∆, e2ψ =

Σ2

ρ2sin2 θ , e2λ = ρ2 ,

ω =2Mar

Σ2, where ρ2 = r2 + a2 cos2 θ

∆ = r2 + a2 − 2Mr

Σ2 = (r2 + a2)2 − a2∆ sin2 θ

(8.15)

This is the Kerr solution expressed in Boyer-Lindquist coordinates. The functionω is the angular-velocity. The Kerr-solution is the metric for space-time outsidea rotating mass-distribution. The constant a is spin per mass-unit for the mass-distribution and M is its mass.

Line-element:

ds2 = −(1− 2Mr

ρ2)dt2 +

ρ2

∆dr2 − 4Mar

ρ2sin2 θdtdφ+ ρ2dθ2

+ (r2 + a2 +2Ma2r

ρ2sin2 θ) sin2 θdφ2

(8.16)

(Here M is a measure of the mass so that M = G ·mass, ie. G = 1)

Page 140: LectureNotes On GeneralRelativity

124 Chapter 8. Black Holes

Light emitted from the surface, r = r0, where g tt = 0 is infinitely redshiftedfurther out. Observed from the outside time stands still.

ρ2 = 2Mr0 ⇒ r20 + a2 cos2 θ = 2Mr0

r0 = M ±√M2 − a2 cos2 θ

(8.17)

This is the equation for the surface which represents infinite redshift.

8.3.1 Zero-angular-momentum-observers (ZAMO’s)

The Lagrange function of a free particle in the equator plane, θ = π2

L = −1

2(e2ν − ω2e2ψ)t2 +

1

2e2µr2 +

1

2e2ψ φ2 +

1

2e2λθ2 − ωe2ψ tφ (8.18)

Here θ = 0. The momentum pφ of the cyclic coordinates φ:

pφ ≡∂L

∂φ= e2ψ(φ− ωt) , t =

dt

dτ, φ =

dτ(8.19)

The angular speed of the particle relative to the coordinate system:

Ω =dφ

dt=φ

t, φ = Ωt

⇒ pφ = e2ψ t(Ω− ω)

(8.20)

pφ is conserved during the movement.

ω =2Mar

(r2 + a2)2 − a2(r2 + a2 − 2Mr),

ω → 0 when r →∞(8.21)

When studying the Kerr metric one finds that Kerr → Minkowski for larger. The coordinate clocks in the Kerr space-time show the same time as thestandard-clocks at rest in the asymptotic Minkowski space-time.

A ZAMO is per definition a particle or observer with pφ = 0. Consider afar away observer who let a stone fall with vanishing initial velocity. pφ is aconstant of motion, so the stone remains a ZAMO during the movement. Alocal reference frame which coincides with the stone is a local inertial frame.

pφ = 0⇒ Ω =dφ

dt= ω (8.22)

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. Thedistant observer has no motion relative to the BL-system. To this observer theBL-system will appear stiff and non-rotating. The observer will observe thatthe local inertial system of the stone obtains an angular speeda is spin

per massunity andMa is spin

dt= ω =

2Mar

(r2 + a2)2 − a2(r2 + a2 − 2Mr)(8.23)

In other words, inertial systems at finite distances from the rotating mass Mare dragged with it in the same direction. This is known as inertial draggingor the Lense-Thirring effect (about 1920).

Page 141: LectureNotes On GeneralRelativity

8.3 Rotating Black Holes: The Kerr metric 125

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 θ. It’s 4-velocity is:

~u =d~x

dτ=dt

d~x

dt

= (−g tt − 2g tφΩ− g φφΩ2)−12 (1,Ω) , where Ω =

dt

(8.24)

To have stationary orbits the following must be true

g φφΩ2 + 2g tφΩ + g tt < 0 (8.25)

This implies that Ω must be in the interval

Ωmin < Ω < Ωmax , (8.26)

where Ωmin = ω −√ω2 − g tt

g φφ, Ωmax = ω +

√ω2 − g tt

g φφsince g tφ = −ωg φφ.

Outside the surface with infinite redshift g tt < 0. That is Ω can be negative,zero and positive. Inside the surface r = r0 with infinite redshift g tt > 0. HereΩmin > 0 and static particles, Ω = 0, cannot exist. This is due to the inertialdragging effect. The surface r = r0 is therefore known as “the static border”.

The interval of Ω, where stationary orbits are allowed, is reduced to zerowhen Ωmin = Ωmax, that is ω2 =

g ttg φφ⇒ g tt = ω2g φφ (equation for the horizon).

For the Kerr metric we have:

g tt = ω2g φφ − e2ν (8.27)

Therefore the horizon equation becomes

e2ν = 0 ⇒ ∆ = 0 ∴ r2 − 2mr + a2 = 0 (8.28)

The largest solution is r+ = M +√M2 − a2 and this is the equation for a

spherical surface. The static border is r0 = M +√M2 − a2 cos θ.

Page 142: LectureNotes On GeneralRelativity

126 Chapter 8. Black Holes

ergo-sphere

K

stationary paths

Mstatic border

horizon

r+

r0

θ = 0

θ = π2Ω > 0

Ω = 0

2M

Figure 8.1: Static border and horizon of a Kerr black hole

Page 143: LectureNotes On GeneralRelativity

Chapter 9

Schwarzschild’s Interior Solution

9.1 Newtonian incompressible star

∇2φ = 4πGρ, φ = φ(r)

1

r2

d

dr(r2dφ

dr) = 4πGρ

(9.1)

Assuming ρ = constant.

d(r2dφ

dr) = 4πGρr2dr

r2dφ

dr=

3Gρr3 +K

= M(r) +K

(9.2)

Gravitational acceleration: ~g = −∇φ = − dφdr~er

g =M(r)

r2+K1

r2=

3Gρr +

K1

r2(9.3)

Finite g in r = 0 demands K1 = 0.

g =4π

3Gρr,

dr=

3Gρr (9.4)

Assume that the massdistribution has a radius R.

φ =2π

3Gρr2 +K2 (9.5)

Demands continuous potensial at r = R.

3GρR2 +K2 =

M(R)

R= −4π

3GρR2

⇒ K2 = −2πGρR2(9.6)

(with zero level at infinite distance). Gives the potensial inside the mass distri-bution:

φ =2π

3Gρ(r2 − 3R2) (9.7)

127

Page 144: LectureNotes On GeneralRelativity

128 Chapter 9. Schwarzschild’s Interior Solution

The star is in hydrostatic equilibrium, that is, the pressure forces are in equi-librium with the gravitational forces.

4π3

πρ

r ρ3

dm=4 dr2r

Figure 9.1: The shell with thickness dr, is affected by both gravitational andpressure forces.

Consider figure 9.1. The pressure forces on the shell is 4πr2dp. Gravitationalforces on the shell:

Gmass inside shell ·mass of the shell

r2

=G4π3 ρr

3 · 4πρr2dr

r2

(9.8)

Equilibrium:

4πr2dp = −G4π

3ρr3 · 4πρdr

dp = −4π

3Gρ2rdr

p = K3 −2πG

3ρ2r2

P (R) = 0 gives : K3 =2πG

3ρ2R2

p(r) =2πG

3ρ2(R2 − r2)

(9.9)

No matter how massive the star is, it is possible for the pressure forces to keepthe equilibrium with gravity. In Newtonian theory, gravitational collapse is nota necessity.

Page 145: LectureNotes On GeneralRelativity

9.2 The pressure contribution to the gravitational mass of a static, sphericalsymmetric system 129

9.2 The pressure contribution to the gravitational massof a static, spherical symmetric system

We now give a new definition of the gravitational acceleration (not equivalentto (7.23))

g = +a

ut, a =

√aµaµ (9.10)

We have the line element:

ds2 = −e2α(r)dt2 + e2β(r)dr2 + r2dΩ2

gtt = −e2α , grr = e2β(9.11)

gives (because of the gravitational acceleration)

g = +eα−βα′ (9.12)

From the expressions for Ett, E rr, E θθ

, Eφφ

follow (see Section 7.1)

E tt−E rr −E θθ −E

φ

φ= −2e2β(

2α′

r+ α′′ + α′2 − α′β′) . (9.13)

We also have

(r2eα−βα′)′= r2eα−β(

2α′

r+ α′′ + α′2 − α′β′) , (9.14)

which gives

g = +1

2r2

∫(E t

t−E rr −E θθ −E

φ

φ)r2eα+βdr . (9.15)

By applying Einstein’s field equations

Eµν = 8πGT µν (9.16)

we get

g = +4πG

r2

∫(T tt− T rr − T θθ − T

φ

φ)r2eα+βdr . (9.17)

This is the Tolman-Whittaker expression for gravitational acceleration.The corresponding Newtonian expression is :

gN = −4πG

r2

∫ρr2dr (9.18)

The relativistic gravitational mass density is therefore defined as

ρG = −T tt+ T rr + T θ

θ+ T φ

φ(9.19)

Page 146: LectureNotes On GeneralRelativity

130 Chapter 9. Schwarzschild’s Interior Solution

For an isotropic fluid with

T tt

= −ρ , T rr = T θθ

= T φφ

= p (9.20)

we get ρG = ρ+ 3p (with c = 1), which becomes

ρG = ρ+3p

c2(9.21)

It follows that in relativity, pressure has a gravitational effect. Greater pressuregives 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 = −e2α(r)dt2 + e2β(r)dr2 + r2dΩ2

Ett

= 8πGTtt, T µν = diag(−ρ, p, p, p)

(9.22)

From Ettwe get

1

r2

d

dr[r(1− e−2β)] = 8πGρ

r(1− e−2β) = 2G

∫ r

04πρr2dr ,

(9.23)

where m(r) =∫ r

0 4πρr2dr giving

e−2β = 1− 2Gm(r)

r=

1

g rr(9.24)

From E rr we have

E rr = 8πGT rr

2

r

dre−2β − 1

r2(1− e−2β) = 8πGp

(9.25)

We get2

r

dr(1− 2Gm(r)

r)− 2Gm(r)

r3= 8πGp

dr= G

m(r) + 4πr3p(r)

r(r − 2Gm(r))(9.26)

Page 147: LectureNotes On GeneralRelativity

9.3 The Tolman-Oppenheimer-Volkov equation 131

The relativistic generalized equation for hydrostatic equilibrium is T rν;ν = 0,giving

T rν,ν + ΓνανTrα + ΓrανT

αν = 0

T rν,ν = T rr,r = p ,r =1√g rr

∂p

∂r

T rν,ν = e−βdp

dr

ΓνανTrα = Γνrνp = Γt

rtp+ Γαrαp

ΓrανTαν = ΓrννT

νν = Γrttρ+ Γrααp

(9.27)

In orthonormal basis we have

Ω νµ = −Ω µν ⇒ Γ µνα = −Γ νµα

Γαrα = Γ αrα = −Γ rαα = −Γrαα(9.28)

T rν;ν = 0 now takes the form:

e−βdp

dr+ Γt

rtp+ Γr

ttρ (9.29)

We have

Γtrt

= −Γtrt

= Γrtt

= Γrtt

(9.30)

and we also have Γrtt

= e−β dαdr , giving:

dp

dr+ (p+ ρ)

dr= 0 (9.31)

Inserting Equation 9.26 into Equation 9.31 gives

dp

dr= −G(ρ+ p)

m(r) + 4πr3p(r)

r(r − 2Gm(r))(9.32)

This is the Tolman-Oppenheimer-Volkov (TOV) equation. The component gtt =−e2α(r) may now be calculated as follows

dp

ρ+ p= −dα , ρ = constant

ln(ρ+ p) = K − αρ+ p = K1e

−α , p = K1e−α − ρ

(9.33)

Hence

eα = eα(R)(1 +p

ρ)−1 (9.34)

where R is the radius of the mass distribution.

Page 148: LectureNotes On GeneralRelativity

132 Chapter 9. Schwarzschild’s Interior Solution

9.4 An exact solution for incompressible stars - Schwarzschild’sinterior solution

The mass inside a radius r for an incompressiable star is

m(r) =4

3πρr3 (9.35)

e−2β = 1− 2Gm(r)

r≡ 1− r2

a2(9.36)

where

a2 =3

8πGρ, m(r) =

r3

2Ga2, rs = 2Gm =

r2

a2r (9.37)

TOV equation:

dp

dr= −G

43πρr

3 + 4πr3p(r)

r(r − 2G 43πρr

3)(ρ+ p(r))

= −G4

3πρ+ 3p(r)

1−G83πρr

2r(ρ+ p(r))

= − 1

2a2ρ

ρ+ 3p(r)

1− r2

a2

r(ρ+ p(r))

⇒∫ p

0

dp

(ρ+ 3p)(ρ+ p= − 1

2a2ρ

∫ r

R

r

1− r2

a2

dr

p+ ρ

3p+ ρ=

√a2 −R2

a2 − r2

(9.38)

So the relativistic pressure distribution is

p(r) =

√a2 − r2 −

√a2 −R2

3√a2 −R2 −

√a2 − r2

ρ, ∀r ≤ R (9.39)

also

a2 =3

8πGρ,a2

r2=

r

rs> 1⇒ a > r (9.40)

To satisfy the condition for hydrostatic equilibrium we must have p > 0 orp(0) > 0 which gives

p(0) ≡ pc =a−√a2 −R2

3√a2 −R2 − a

> 0 (9.41)

in which the numerator is positive so that

3√a2 −R2 > a

9a2 − 9R2 > a2

R <

√8

9a

R2 <8

9a =

8

9

3

8πGρ=

1

3πGρ

(9.42)

Page 149: LectureNotes On GeneralRelativity

9.4 An exact solution for incompressible stars - Schwarzschild’s interior solution133

Stellar mass:

M =4

3πρR3 <

4

3πρR

1

3πGρ=

4R

9G

M <4

9G

1√3πGρ

(9.43)

For a neutron star we can use ρ ≈ 1017 g /cm3. An upper limit on the mass isthen M < 2.5 M Substitution for p in the expression for eα gives

eα =3

2

√1− Rs

R− 1

2

√1− Rs

R3r2 (9.44)

The line element for the interior Schwarzschild solution is

ds2 = −(

3

2

√1− Rs

R− 1

2

√1− Rs

R3r2

)2

dt2 +dr2

1− RsR3 r2

+ r2dΩ, r ≤ R

(9.45)

Page 150: LectureNotes On GeneralRelativity

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 asreference particles. Then we introduce a ’comoving coordinate system’ in thisframe of reference with spatial coordinates χ, θ, φ. We use time measured onstandard 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[dχ2 + r(χ)2dΩ2] (10.1)

(For standard clocks at rest in the expanding system, dχ = dΩ = 0 and ds2 =−dτ2 = −dt2). The function a(t) is called the expansion factor, and t is calledcosmic time.

The physical distance to a galaxy with coordinate distance dχ from an ob-server at the origin, is:

dlx =√gχχdχ = a(t)dχ (10.2)

Even if the galactic clusters have no coordinate velocity, they do have a radialvelocity expressed by the expansion factor.

The value χ determines which cluster we are observing and a(t) how it ismoving. 4-velocity of a reference particle (galactic cluster):

uµ =dxµ

dτ=dxµ

dt= (1, 0, 0, 0) (10.3)

This applies at an abritrary time, that is duµ

dt = 0. Geodesic equation: duµ

dt +Γµαβu

αuβ = 0 which is reduces to: Γµtt = 0

Γµtt =1

2gµν(

0︷︸︸︷gνt,t +

0︷︸︸︷gtν,t +

0︷︸︸︷gtt,ν ) = 0 (10.4)

We have that gtt = −1. This shows that the reference particles are freely falling.

134

Page 151: LectureNotes On GeneralRelativity

10.2 Curvature isotropy - the Robertson-Walker metric 135

10.2 Curvature isotropy - the Robertson-Walker met-ric

Introduce orthonormal form-basis:

ωt = dt ωχ = a(t)dχ ωθ = a(t)r(χ)dθ

ωφ = a(t)r(χ) sin θdφ(10.5)

Using Cartans 1st equation:

dωµ = −Ωµν ∧ ων (10.6)

to find the connection forms. Then using Cartans 2nd structure equation tocalculate the curvature forms:

Rµν = dΩµν + Ωµ

λ∧ Ωλ

ν (10.7)

Calculations give: (notation: · = ddt ,′ = d

dχ )

Rti

=a

aωt ∧ ωi, ωi = ωχ, ωθ, ωφ

Rχj

=( a2

a2− r′′

ra2

)ωχ ∧ ωj , ωj = ωθ, ωφ

Rθφ

=( a2

a2+

1

r2a2− r′2

r2a2

)ωθ ∧ ωφ

(10.8)

The curvature of 3-space (dt = 0) can be found by putting a = 1. That is:

3Rχ

j= −r

′′

rωχ ∧ ωj

3Rθφ

=( 1

r2− r′2

r2

)ωθ ∧ ωφ

(10.9)

The 3-space is assumed to be isotropic and homogenous. This demands

−r′′

r=

1− r′2r2

= k , (10.10)

where k represents the constant curvature of the 3-space.

∴ r′′ + kr = 0 and r′ =√

1− kr2 (10.11)

Solutions with r(0) = 0, r′(0) = 1 :√−kr = sinh(

√−kχ) (k < 0)

r = χ (k = 0) (10.12)√kr = sin(

√kχ) (k > 0)

Page 152: LectureNotes On GeneralRelativity

136 Chapter 10. Cosmology

The solutions can be characterized by the following 3 cases:

r = sinhχ, dr =√

1 + r2dχ, (k = −1)

r = χ, dr = dχ, (k = 0) (10.13)

r = sinχ, dr =√

1− r2dχ, (k = 1)

In all three cases one may write dr =√

1− kr2dχ, which is just the last equationabove.

We now set dχ2 = dr2

1−kr2 into the line-element :

ds2 = −dt2 + a2(t)(dχ2 + r2(χ)dΩ2

)

= −dt2 + a2(t)

(dr2

1− kr2+ r2dΩ2

) (10.14)

The first expression is known as the standard form of the line-element, thesecond is called the Robertson-Walker line-element.

The 3-space has constant curvature. 3-space is spherical for k = 1, Euclideanfor k = 0 and hyperbolic for k = −1.

Universe models with k = 1 are known as ’closed’ and models with k = −1are known as ’open’. Models with k = 0 are called ’flat’ even though thesemodels also have curved space-time.

10.3 Cosmic dynamics

10.3.1 Hubbles law

The observer is placed in origo of the coordinate-system; χ0 = 0. The properdistance to a galaxy with radial coordinate χe is D = a(t)χe. The galaxy has aradial velocity:

v =dD

dt= aχe =

a

aD = HD whereH =

a

a(10.15)

The expansion velocity v is proportional to the distance D. This is Hubbleslaw.

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 dθ = dφ = 0 we have :

dt = −a(t)dχ (10.16)

Page 153: LectureNotes On GeneralRelativity

10.3 Cosmic dynamics 137

-

6t

χ

t0 + ∆t0

t0

χ0 = 0 χe

te + ∆te

te

Figure 10.1: Schematic representation of cosmological redshift

Integration from transmitter-event to receiver-event :

∫ t0

te

dt

a(t)= −

∫ χ0

χe

dχ = χe

∫ t0+∆t0

te+∆te

dt

a(t)= −

∫ χ0

χe

dχ = χe ,

which gives ∫ t0+∆t0

te+∆te

dt

a−∫ t0

te

dt

a= 0 (10.17)

or∫ t0+∆t0

t0

dt

a−∫ te+∆te

te

dt

a= 0 (10.18)

Under the integration from te to te + ∆te the expansion factor a(t) can beconsidered a constant with value a(te) and under the integration from t0 tot0 + ∆t0 with value a(t0), giving:

∆tea(te)

=∆t0a(t0)

(10.19)

Page 154: LectureNotes On GeneralRelativity

138 Chapter 10. Cosmology

∆t0 and ∆te are intervals of the light at the receiving and transmitting time.Since the wavelength of the light is λ = c∆t we have:

λ0

a(t0)=

λea(te)

(10.20)

This can be interpreted as a “stretching” of the electromagnetic waves due tothe expansion of space. The cosmological redshift is denoted by z and is givenby:

z =λ0 − λeλe

=a(t0)

a(te)− 1 (10.21)

Using a0 ≡ a(t0) we can write this as:

1 + z(t) =a0

a(10.22)

10.3.3 Cosmic fluids

The energy-momentum tensor for a perfect fluid (no viscosity and no thermalconductivity) is

Tµν = (ρ+ p)uµuν + pgµν (10.23)

In an orthonormal basis

Tµν = (ρ+ p)uµuν + pηµν (10.24)

where ηµν is the Minkowski metric. We consider 3 types of cosmic fluid:

1. dust: p = 0,T µν = ρuµuν (10.25)

2. radiation: p = 13ρ,

T µν =4

3ρuµuν + pη µν

3(4uµuν + η µν)

(10.26)

The traceT = T µµ =

ρ

3(4uµuµ + δµµ) = 0 (10.27)

3. vacuum: p = −ρ,T µν = −ρη µν (10.28)

If vacuum can be described as a perfect fluid we have pv = −ρv, whereρ is the energy density. It can be related to Einstein’s cosmological constant,Λ = 8πGρv .

One has also introduced a more general type of vacuum energy given bythe equation of state pφ = wρφ, where φ denotes that the vacuum energy is

Page 155: LectureNotes On GeneralRelativity

10.3 Cosmic dynamics 139

connected to a scalar field φ. In a homogeneous universe the pressure and thedensity are given by

pφ =1

2φ2 − V (φ), ρφ =

1

2φ2 + V (φ) (10.29)

where V (φ) is the potential for the scalar field. Then we have

w =12 φ

2 − V (φ)12 φ

2 + V (φ)(10.30)

The special case φ = 0 gives the Lorentz invariant vacuum with w = −1. Themore general vacuum is called “quintessence”.

10.3.4 Isotropic and homogeneous universe models

We will discuss isotropic and homogenous universe models with perfect fluidand a non-vanishing cosmological constant Λ. Calculating the components ofthe Einstein tensor from the line-ement (10.14) we find in an orthonormal basis

Ett =3a2

a2+

3k

a2(10.31)

Emm = −2a

a− a2

a2− k

a2. (10.32)

The components of the energy-momentum tensor of a perfect fluid in a comovingorthonormal basis are

Ttt = ρ, Tmm = p. (10.33)

Hence the tt component of Einstein’s field equations is

3a2 + k

a2= 8πGρ+ Λ (10.34)

mm components:

−2a

a− a2

a2− k

a2= 8πGp− Λ (10.35)

where ρ is the energy density and p is the pressure. The equations with vanishingcosmological constant are called the Friedmann equations. Inserting eq. (10.34)into eq. (10.35) gives:

a = −4πG

3a(ρ+ 3p) (10.36)

If we interpret ρ as the mass density and use the speed of light c, we get

a = −4πG

3a(ρ+ 3p/c2) (10.37)

Inserting the gravitational mass density ρG from eq.(9.21) this equation takesthe form

a = −4πG

3aρG (10.38)

Page 156: LectureNotes On GeneralRelativity

140 Chapter 10. Cosmology

Inserting p = wρc2 into (9.21) gives

ρG = (1 + 3w)ρ (10.39)

which is negative for w < −1/3, i.e. for φ2 < V (φ). Special cases:

• dust: w = 0, ρG = ρ

• radiation: w = 13 , ρG = 2ρ

• Lorentz-invariant vacuum: w = −1, ρG = −2ρ

In a universe dominated by a Lorentz-invariant vacuum the acceleration of thecosmic expansion is

av =8πG

3aρv > 0, (10.40)

i.e. accelerated expansion. This means that vacuum acts upon itself with repul-sive gravitation.

The field equations can be combined into

H2 ≡(a

a

)2

=8πG

3ρm +

Λ

3− k

a2(10.41)

where ρm is the density of matter, Λ = 8πGρΛ where ρΛ is the vacuum energywith constant density. ρ = ρm + ρΛ is the total mass density. Then we maywrite

H2 =8πG

3ρ− k

a2(10.42)

The critical density ρcr is the density in a universe with euclidean spacelikegeometry, k = 0, which gives

ρcr =3H2

8πG(10.43)

We introduce the relative densities

Ωm =ρmρcr

, ΩΛ =ρΛ

ρcr(10.44)

Furthermore we introduce a dimensionless parameter that describes the curva-ture of 3-space

Ωk = − k

a2H2(10.45)

Eq. (10.42) can now be written

Ωm + ΩΛ + Ωk = 1 (10.46)

From the Bianchi identity and Einstein’s field equations follow that the energy-momentum density tensor is covariant divergence free. The time-componentexpresses the equation of continiuty and takes the form

[(ρ+ p)utuν ];ν + (pηtν);ν = 0 (10.47)

Page 157: LectureNotes On GeneralRelativity

10.3 Cosmic dynamics 141

Since ut = 1, um = 0 and ηtt = −1, ηtm = 0, we get

(ρ+ p). + (ρ+ p)uν;ν − p = 0 (10.48)

orρ+ (ρ+ p)(uν,ν + Γν

tν) = 0 (10.49)

Here uν,ν = 0 and Γttt

= 0. Calculating Γmtm

for dωµ = Γµαβωα ∧ ωβ we get

Γmtm

= Γrtr

+ Γθtθ

+ Γφtφ

= 3a

a(10.50)

Hence

ρ+ 3(ρ+ p)a

a= 0 (10.51)

which may be written(ρa3). + p(a3). = 0 (10.52)

Let V = a3 be a comoving volume in the universe and U = ρV be the energyin the comoving volume. Then we may write

dU + pdV = 0 (10.53)

This is the first law of thermodynamics for an adiabatic expansion. It followsthat the universe expands adiabatically. The adiabatic equation can be written

ρ

ρ+ p= −3

a

a(10.54)

Assuming p = wρ we get

ρ= −3(1 + w)

da

a

lnρ

ρ0= ln

(a

a0

)−3(1+w)

It follows that

ρ = ρ0

(a

a0

)−3(1+w)

(10.55)

This equation tells how the density of different types of matter depends on theexpansion factor

ρa3(1+w) = constant (10.56)

Special cases:

• dust: w = 0 gives ρda3 = constantThus, the mass in a comoving volume is constant.

Page 158: LectureNotes On GeneralRelativity

142 Chapter 10. Cosmology

• radiation: w = 13 gives ρra4 = constant

Thus, the radiation energy density decreases faster than thecase with dust when the universe is expanding. The energyin a comoving volume is decreasing because of the thermo-dynamic work on the surface. In a remote past, the densityof radiation must have exceeded the density of dust:

ρd0a30 =ρda

3

ρr0a40 =ρra

4

ρra4

ρda3=ρr0a

40

ρd0a30

The expansion factor when ρr = ρd:

a(t1) =ρr0ρd0

a0

• Lorentz-invariant vacuum: w = −1 gives ρΛ = constant.The vacuum energy in a comoving volume is increasing ∝ a3.

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

(aa+ kt)· = 0(10.57)

Integration givesaa+ kt = B (10.58)

Another integration gives

1

2a2 +

1

2kt2 = Bt+C (10.59)

The initial condition a(0) = 0 gives C = 0. Hence

a =√

2Bt− kt2 (10.60)

For k = 0 we have

a =√

2Bt , a =

√B

2t(10.61)

The expansion velocity reaches infinity at t = 0, (limt→0 a =∞)

Page 159: LectureNotes On GeneralRelativity

10.4 Some cosmological models 143

Figure 10.2: In a radiation dominated universe the expansion velocity reachesinfinity at t = 0.

ρRa4 = K , a =

√2Bt

4ρRB2t2 = K

(10.62)

According to the Stefan-Boltzmann law we then have

ρR = σT 4 → 4B2σT 4t2 = K ⇒

t =K1

T 2⇔ T =

√K1

t

(10.63)

where T is the temperature of the background radiation.

10.4.2 Dust dominated model

From the first of the Friedmann equations we have

a2 + k =8πG

3ρa2 (10.64)

We now introduce a time parameter η given by

dt

dη= a(η) ⇒ d

dt=

1

a

d

So: a =da

dt=

1

a

da

(10.65)

We also introduce A ≡ 8πG3 ρ0a0

3. The first Friedmann equation then gives

aa2 + ka =8πG

3ρa3 =

8πG

3ρ0a0

3 = A (10.66)

Page 160: LectureNotes On GeneralRelativity

144 Chapter 10. Cosmology

Using η we get

1

a(da

dη)2 = A− ka

1

a2(da

dη)2 =

A

a− k

1

a

da

dη=

√A

a− k =

√A

a

√1− a

Ak

(10.67)

where we chose the positive root. We now introduce u, given by a = Au2, u =√aA . We then get

da

dη= 2Au

du

dη(10.68)

which together with the equation above give

1

Au22Au

du

dη=

1

u

√1− ku2

⇓du√

1− ku2=

1

2dη

(10.69)

This equation will first be integrated for k < 0. Then k = −|k|, so that∫

du√1 + |k|u2

2+K (10.70)

or arcsinh(√−ku) = η

2 +K. The condition u(0) = 0 gives K = 0. Hence

− kAa = sinh2 η

2=

1

2(cosh η − 1) (10.71)

ora = − A

2k(cosh η − 1) (10.72)

From eqs. (10.43), (10.44) and (10.66) we have

A =8πG

3ρm0 = H2

0

ρm0

ρcr0= H2

0Ωm0 (10.73)

From egs. (10.45) and (10.46) we get

k = H20 (Ωm0 − 1) (10.74)

Hence, the scale factor of the negatively curved, dust dominated universe modelis

a(η) =1

2

Ωm0

1− Ωm0(cosh η − 1) (10.75)

Page 161: LectureNotes On GeneralRelativity

10.4 Some cosmological models 145

Inserting this into eq. (10.65) and integrating with t(0) = η(0) leads to

t(η) =Ωm0

2H0(1− Ωm0)3/2(sinh η − η) (10.76)

Integrating eg. (10.69) for k = 0 leads to an Einstein-deSitter universe

a(t) = (t

t0)

23 (10.77)

Finally integrating eg. (10.69) for k > 0 gives, in a similar way as for k < 0

a(η) =1

2

Ωm0

1− Ωm0(1− cos η) (10.78)

t(η) =Ωm0

2H0(Ωm0 − 1)3/2(η − sin η) (10.79)

We see that this is a parametric representation of a cycloid.

In the Einstein-deSitter model the Hubble factor is

H =a

a=

2

3

1

t, t =

2

3

1

H=

2

3tH (10.80)

The critical density in the Einstein-deSitter model is given by the first Fried-mann equation:

H2 =8πG

3ρcr , k = 0

ρcr =3H2

8πG, Ω =

ρ

ρcr

(10.81)

Example 10.4.1 (Age-redshift relation for dust dominated universe with k = 0)

1 + z =a0

a⇒ a =

a0

1 + z

da = − a0

(1 + z)2dz = − a

1 + zdz

(10.82)

Eq. (10.34) gives

( aa

)2=

8πG

3ρ =

8πG

3

ρ0a30

a3

=8πG

3ρ0(1 + z)3

(10.83)

Page 162: LectureNotes On GeneralRelativity

146 Chapter 10. Cosmology

Figure 10.3: For k = 1 the density is larger than the critical density, and theuniverse is closed. For k = 0 we have ρ = ρcr and the expansion velocity ofthe universe will approach zero as t → ∞. For k = −1 we have ρ < ρcr. Theuniverse is then open, and will continue expanding forever.

Using H20 = 8πG

3 ρ0 gives aa = H0(1 + z)

32 . From a = da

dt we get:

dt =da

a=da

a aa= − dz

H0(1 + z)52

(10.84)

Integration gives the age of the universe:

t0 = − 1

H0

∫ 0

dz

(1 + z)52

=2

3

1

H0

[ 1

(1 + z)32

]0∞ (10.85)

t0 = 23 tH where the Hubble-time tH ≡ 1

H0is the age of the universe, if the expansion

rate had been constant. ’Look-back-time’ to a source with redshift z is:

∆t = tH

∫ z

0

dz

(1 + z)52

=2

3tH[1− 1

(1 + z)32

](10.86)

Page 163: LectureNotes On GeneralRelativity

10.4 Some cosmological models 147

tangent

time t

exp.factor a

t0

tH

today

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.

∆t = t0[1− 1

(1 + z)3/2] (10.87)

z =1

(1− ∆tt0

)2/3− 1 (10.88)

∆t

t0= 0, 99⇒ z = 20, 5 (10.89)

10.4.3 Friedmann-Lemaître model

The dynamics of galaxies and clusters of galaxies has made it clear that farstronger gravitational fields are needed to explain the observed motions thanthose produced by visible matter (McGaugh 2001). At the same time it hasbecome clear that the density of this dark matter is only about 30% of the critical

Page 164: LectureNotes On GeneralRelativity

148 Chapter 10. Cosmology

density, although it is a prediction by the usual versions of the inflationaryuniverse models that the density ought to be equal to the critical density (Linde2001). Also the recent observations of the temperature fluctuations of the cosmicmicrowave 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 thatfills up to the critical density must be evenly distributed in order not to affectthe dynamics of the galaxies and the clusters.

Furthermore, about two years ago observations of supernovae of type Ia withhigh 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 ofrepulsive gravitation due to some sort of vacuum energy. Thereby the missingenergy needed to make space flat, was identified as vacuum energy. Hence, itseems that we live in a flat universe with vacuum energy having a density around70% of the critical density and with matter having a density around 30% of thecritical density.

Until the discovery of the accelerated expansion of the universe the standardmodel of the universe was assumed to be the Einstein-DeSitter model, which isa 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 energyhave been discussed (Zlatev, Wang and Steinhardt 1999, Carroll 1998). How-ever, the most simple type of vacuum energy is the Lorentz invariant vacuumenergy (LIVE), which has constant energy density during the expansion of theuniverse (Zeldovich 1968, Grøn 1986). This type of energy can be mathemati-cally represented by including a cosmological constant in Einstein’s gravitationalfield equations. The flat universe model with cold dark matter and this type ofvacuum energy is the Friedmann-Lemaître model.

The field equations for the flat Friedmann-Lemaître is found by puttingk = p = 0 in equation (10.35). This gives

2a

a+a2

a2= Λ (10.90)

Integration leads to

aa2 =Λ

3a3 +K (10.91)

where K is a constant of integration. Since the amount of matter in a volumecomoving with the cosmic expansion is constant, ρMa3 = ρM0a

30, where the

index 0 refers to measured values at the present time. Normalizing the expansionfactor so that a0 = 1 and comparing eqs.(10.42) and(10.91) then gives K =(8πG/3)ρM0. Introducing a new variable x by a3 = x2 and integrating oncemore with the initial condition a(0) = 0 we obtain

a3 =3K

Λsinh2

(t

), tΛ =

2√3Λ

(10.92)

Page 165: LectureNotes On GeneralRelativity

10.4 Some cosmological models 149

The vacuum energy has a constant density ρΛ given by

Λ = 8πGρΛ (10.93)

The critical density, which is the density making the 3-space of the universe flat,is

ρcr =3H2

8πG(10.94)

The relative density, i.e. the density measured in units of the critical density, ofthe matter and the vacuum energy, are respectively

ΩM =ρ

ρcr=

8πGρM3H2

(10.95)

ΩΛ =ρΛ

ρcr=

Λ

3H2(10.96)

Since the present universe model has flat space, the total density is equal tothe critical density, i.e. ΩM + ΩΛ = 1. Eq. (10.91) with the normalizationa(t0) = 1, where t0 is the present age of the universe, gives 3H2

0 = 3K + Λ.Eq. (10.34) with k = 0 gives 8πGρ0 = 3H2

0 − Λ. Hence K = 8πGρ0/3 and3KΛ = 8πGρ0

Λ = ρ0

ρΛ= ΩM0

ΩΛ0. In terms of the values of the relative densities at the

present time the expression for the expansion factor then takes the form

a = A1/3 sinh2/3

(t

), A =

ΩM0

ΩΛ0=

1−ΩΛ0

ΩΛ0(10.97)

Using the identity sinh(x/2) =√

(coshx− 1)/2 this expression may be written

a3 =A

2

[cosh

(2t

)− 1

](10.98)

The age t0 of the universe is found from a(t0) = 1, which by use of the formulaarc tanhx = arc sinh(x/

√1− x2), leads to the expression

t0 = tΛarc tanh√

ΩΛ0 (10.99)

Inserting typical values t0 = 15 · 109years, ΩΛ0 = 0.7 we get A = 0.43, tΛ =12·109years. With these values the expansion factor is a = 0.75 sinh2/3(1.2t/t0).This function is plotted in fig. 10.5. The Hubble parameter as a function oftime is

H = (2/3tΛ) coth(t/tΛ) (10.100)

Inserting t0 = 1.2tΛ we get Ht0 = 0.8 coth(1.2t/t0), which is plotted in fig. 10.6The Hubble parameter decreases all the time and approaches a constant valueH∞ = 2/3tΛ in the infinite future. The present value of the Hubble parameteris

H0 =2

3tΛ√

ΩΛ0(10.101)

The corresponding Hubble age is tH0 = (3/2)tΛ√

ΩΛ0. Inserting our numericalvalues gives H0 = 64km/secMpc−1 and tH0 = 15.7 · 109years. In this universe

Page 166: LectureNotes On GeneralRelativity

150 Chapter 10. Cosmology

a

t/t00.2 0.4 0.6 0.8 1 1.2 1.4

0.25

0.5

0.75

1

1.25

1.5

Figure 10.5: The expansion factor as function of cosmic time in units of the ageof the universe.

Figure 10.6: The Hubble parameter as function of cosmic time.

Page 167: LectureNotes On GeneralRelativity

10.4 Some cosmological models 151

model the age of the universe is nearly as large as the Hubble age, while inthe Einstein-DeSitter model the corresponding age is t0ED = (2/3)tH0 = 10.5 ·109years. The reason for this difference is that in the Einstein-DeSitter modelthe expansion is decelerated all the time, while in the Friedmann-Lemaître modelthe repulsive gravitation due to the vacuum energy have made the expansionaccelerate lately (see below). Hence, for a given value of the Hubble parameterthe previous velocity was larger in the Einstein-DeSitter model than in theFriedmann-Lemaître model.

The ratio of the age of the universe and its Hubble age depends upon thepresent relative density of the vacuum energy as follows,

t0tH0

= H0t0 =2

3

arc tanh√

ΩΛ0√ΩΛ0

(10.102)

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 functionof the present relative density of the vacuum energy.

with increasing density of vacuum energy. In the limit that the density of thevacuum approaches the critical density, thereis no dark matter, and the universemodel approaches the DeSitter model with exponential expansion and no BigBang. This model behaves in the same way as the Steady State cosmologicalmodel 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 timeis

q =1

2[1− 3 tanh2(t/tΛ)] (10.103)

Page 168: LectureNotes On GeneralRelativity

152 Chapter 10. Cosmology

which is shown graphically in fig. 10.8 The inflection point of time t1 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 = tΛarc tanh(1/√

3) (10.104)

or expressed in terms of the age of the universe

t1 =arc tanh(1/

√3)

arc tanh√

ΩΛ0t0 (10.105)

The corresponding cosmic red shift is

z(t1) =a0

a(t1)− 1 =

(2ΩΛ0

1− ΩΛ0

)1/3

− 1 (10.106)

Inserting ΩΛ0 = 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 te = 0.30t0 = 4.5 · 109years, have providedevidence that the universe was decelerated at that time (Riess n.d.). M.Turnerand A.G.Riess (Turner and Riess 2001) have recently argued that the othersupernova data favour a transition from deceleration to acceleration for a redshift around z = 0.5.

Note that the expansion velocity given by Hubble’s law, v = Hd, alwaysdecreases as seen from fig. 10.6. This is the velocity away from the Earth of thecosmic fluid at a fixed physical distance d from the Earth. The quantity a on theother hand, is the velocity of a fixed fluid particle comoving with the expansion

Page 169: LectureNotes On GeneralRelativity

10.4 Some cosmological models 153

of the universe. If such a particle accelerates, the expansion of the universe issaid to accelerate. While H tells how fast the expansion velocity changes at afixed distance from the Earth, the quantity a represents the acceleration of afree particle comoving with the expanding universe. The connection betweenthese two quantities are a = a(H +H2).

The ratio of the inflection point of time and the age of the universe, as givenin eq.(10.105), is depicted graphically as function of the present relative densityof 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 overto acceleration to the age of the universe.

greater the vacuum density is. The change from deceleration to accelerationwould happen at the present time if ΩΛ0 = 1/3.

The red shift of the inflection point given in eq.(10.106) as a function ofvacuum energy density, is plotted in fig. 10.10 Note that the red shift of futurepoints of time is negative, since then a > a0. If ΩΛ0 < 1/3 the transition toacceleration will happen in the future.

The critical density is

ρcr = ρΛ tanh−2(t/tΛ) (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

ΩΛ = tanh2(t/tΛ) (10.108)

which is plotted in fig. 10.12. The density of the vacuum energy approachesthe critical density. Since the density of the vacuum energy is constant, this is

Page 170: LectureNotes On GeneralRelativity

154 Chapter 10. Cosmology

Figure 10.10: The cosmic red shift of light emitted at the turnover time fromdeceleration to acceleration as function of the present relative density of vacuumenergy.

Figure 10.11: The critical density in units of the constant density of the vacuumenergy as function of time.

Page 171: LectureNotes On GeneralRelativity

10.4 Some cosmological models 155

Figure 10.12: The relative density of the vacuum energy density as function oftime.

better expressed by saying that the critical density approaches the density ofthe vacuum energy. Furthermore, since the total energy density is equal to thecritical density all the time, this also means that the density of matter decreasesfaster than the critical density. The density of matter as function of time is

ρM = ρΛ sinh−2(t/tΛ) (10.109)

which is shown graphically in fig. 10.13 The relative density of matter as func-tion of time is

ΩM = cosh−2(t/tΛ) (10.110)

which is depicted in fig. 10.14 Adding the relative densities of fig. 10.13 andfig. 10.14 or the expressions (10.107) and (10.109) we get the total relativedensity ΩTOT = ΩM + ΩΛ = 1.

The universe became vacuum dominated at a point of time t2 when ρΛ(t2) =ρM (t2). From eq.(10.109) follows that this point of time is given by sinh(t2/tΛ) =1. According to eq.(10.99) we get

t2 =arc sinh(1)

arc tanh(√

ΩΛ0)t0 (10.111)

From eq.(10.97) follows that the corresponding red shift is

z(t2) = A−1/3 − 1 (10.112)

Inserting ΩΛ0 = 0.7 gives t2 = 0.73t0 and z(t2) = 0.32. The transition toaccelerated expansion happens before the universe becomes vacuum dominated.

Page 172: LectureNotes On GeneralRelativity

156 Chapter 10. Cosmology

Figure 10.13: The density of matter in units of the density of vacuum energy asfunction of time.

Figure 10.14: The relative density of matter as function of time.

Page 173: LectureNotes On GeneralRelativity

10.5 Inflationary Cosmology 157

Note from eqs.(10.103) and (10.108) that in the case of the flat Friedmann-Lemaître universe model, the deceleration parameter may be expressed in termsof the relative density of vacuum only, q = (1/2)(1 − 3ΩΛ). The supernova Iaobservations have shown that the expansion is now accelerating. Hence if theuniverse is flat, this alone means that ΩΛ0 > 1/3.

As mentioned above, many different observations indicate that we live in auniverse with critical density, where cold matter contributes with about 30%of the density and vacuum energy with about 70%. Such a universe is welldescribed by the Friedmann-Lemaître universe model that have been presentedabove.

However, this model is not quite without problems in explaining the observedproperties of the universe. In particular there is now much research directed atsolving the so called coincidence problem. As we have seen, the density of thevacuum energy is constant during the expansion, while the density of the mat-ter decreases inversely proportional to a volume comoving with the expandingmatter. Yet, one observes that the density of matter and the density of thevacuum energy are of the same order of magnitude at the present time. Thisseems to be a strange and unexplained coincidence in the model. Also just atthe present time the critical density is approaching the density of the vacuumenergy. At earlier times the relative density was close to zero, and now it changesapproaching the constant value 1 in the future. S. M. Carroll (Carroll 2001) hasillustrated this aspect of the coincidence problem by plotting ΩΛ as a functionof ln(t/t0). Differentiating the expression (10.108) we get

tΛ2

dΩΛ

dt=

sinh(t/tΛ)

cosh3(t/tΛ)(10.113)

which is plotted in fig. 10.15Putting ΩΛ = 0 we find that the rate of change of ΩΛ was maximal at

the point of time t1 when the deceleration of the cosmic expansion turned intoacceleration. 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 hopethat 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 thedynamics of a vacuum dominated universe by studying simple and beautifuluniverse models such as the Friedmann-Lemaître 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 Bin opposite directions has the same temperature. This means that it has been

Page 174: LectureNotes On GeneralRelativity

158 Chapter 10. Cosmology

Figure 10.15: Rate of change of ΩΛ as function of ln( tt0

). The value ln( tt0

) = −40corresponds to the cosmic point of time t0 ∼ 1s.

radiated by sources of the same temperature in these points. Thus, the universemust have been in thermic equilibrium at the decoupling time, td = 3 ·105years.This implies that points A and B, “at opposite sides of the universe”, had beenin causal contact already at that time. I.e., a light signal must have had time tomove from A to B during the time from t = 0 to t = 3 · 105 years. The pointsA and B must have been within each other’s 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. thecurve is defined by ds2 = 0. We get

dr =dt

a(t). (10.114)

The coordinate distance the photon has moved during the time t is

∆r =

∫ t

0

dt

a(t). (10.115)

The physical distance the light has moved at the time t is called the horizondistance, and is

lh = a(t)∆t = a(t)

∫ t

0

dt

a(t). (10.116)

To find a quantitative expression for the “horizon problem”, we may considera model with critical mass density (Euclidian spacelike geometry.) Using p = wρand Ω = 1, integration of equation (10.36) gives

a ∝ t 23+3w . (10.117)

Page 175: LectureNotes On GeneralRelativity

10.5 Inflationary Cosmology 159

Inserting this into the expression for lh and integrating gives

lh =3w + 3

3w + 1t. (10.118)

Let us call the volume inside the horizon the “horizon volume” and denote it byVH . From equation (10.118) follows that VH ∝ t3. At the decoupling time, thehorizon volume may therefore be written

(VH)d =

(tdt0

)3

V0, (10.119)

where V0 is the size of the present horizon volume. Events within this volumeare causally connected, and a volume of this size may be in thermal equilibriumat the decoupling time.

Let (V0)d be the size, at the decoupling, of the part of the universe thatcorresponds to the present horizon volume, i.e. the observable universe. For ourEuclidean universe, the equation (10.117) holds, giving

(V0)d =a3(td)

a3(t0)V0 =

(tdt0

) 2w+1

V0. (10.120)

From equations (10.119) and (10.120), we get

(V0)d(VH)d

=

(tdt0

) 3w+1w+1

. (10.121)

Using that td = 10−4t0 and inserting w = 0 for dust, we find V0)d(VH)d

= 104.Thus, there was room for 104 causally connected areas at the decoupling timewithin what presently represents our observable universe. Points at oppositesides of our observable universe were therefore not causally connected at thedecoupling, according to the Friedmann models of the universe. These modelscan therefore not explain that the temperature of the radiation from such pointsis the same.

The Flatness Problem

According to eq. (10.42), the total mass parameter Ω = ρρcr

is given by

Ω− 1 =k

a2. (10.122)

By using the expansion factor (10.117) for a universe near critical massdensity, we get

Ω− 1

Ω0 − 1=

(t

t0

)2( 3w+13w+3 )

. (10.123)

For a radiation dominated universe, we get

Ω− 1

Ω0 − 1=

t

t0. (10.124)

Page 176: LectureNotes On GeneralRelativity

160 Chapter 10. Cosmology

Measurements indicate that Ω0 − 1 is of order of magnitude 1. The age ofthe universe is about t0 = 1017s. When we stipulate initial conditions for theuniverse, it is natural to consider the Planck time, tP = 10−43s, since this isthe limit to the validity of general relativity. At earlier time, quantum effectswill be important, and one can not give a reliable description without usingquantum gravitation. The stipulated initial condition for the mass parameterthen becomes that Ω− 1 is of order 10−60 at the Planck time. Such an extremefine tuning of the initial value of the universe’s mass density can not be explainedwithin standard Big Bang cosmology.

Other Problems

The Friedman models can not explain questions about why the universe is nearlyhomogeneous and has an isotropic expansion, nor say anything about why theuniverse is expanding.

10.5.2 Cosmic Inflation

Spontaneous Symmetry Breaking and the Higgs Mechanism

The particles responsible for the electroweak force, W ± and Z0 are massive(causing the weak force to only have short distance effects). This was originallya problem for the quantum field theory describing this force, since it made itdifficult to create a renormalisable theory1. This was solved by Higgs and Kibblein 1964 by introducing the so-called Higgs mechanism.

The main idea is that the massive bosons W± and Z0 are given a mass byinteracting with a Higgs field φ. The effect causes the mass of the particles to beproportional to the value of the Higgs field in vacuum. It is therefore necessaryfor the mechanism that the Higgs field has a value different from zero in thevacuum (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 fieldmay be temperature dependent. Let us assume that the potential for the Higgsfield is described by the function

V (φ) =1

2µ2φ2 +

1

4λφ4, (10.125)

where the sign of µ2 depends on whether the temperature is above or below acritical temperature Tc. This sign has an important consequence for the shapeof the potential V . The potential is shown in figure 10.16 for two differenttemperatures. For T > Tc, µ2 > 0, and the shape is like in fig. 10.16(a), andthere is a stable minimum for φ = 0. However, for T < Tc, µ2 < 0, and theshape is like in fig. 10.16(b). In this case the potential has stable minima forφ = ±φ0 = ± |µ|√

λand an unstable maximum at φ = 0. For both cases, the

1The problem is that the Lagrangian for the gauge bosons can not include terms likem2W 2µ ,

which are not gauge invariant

Page 177: LectureNotes On GeneralRelativity

10.5 Inflationary Cosmology 161

potential V (φ) is invariant under the symmetry transformation φ 7→ −φ (i.e.V (φ) = V (−φ)).

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 φ = 0. On the otherhand, 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 φ = ±φ0. When thesystem is in one of these states, it is no longer symmetric under the change ofsign of φ. Such a symmetry, which is not reflected in the vacuum state, is calledspontaneously broken. Note that from figure 10.16(b) we see that the energy ofthe false vacuum is larger than for the real vacuum.

Figure 10.16: The shape of the potential depends on the sign of µ2.(a): Higher temperature than the critical, with µ2 > 0.(b): Lower temperature than the critical, with µ2 < 0.

The central idea, which originated the “inflationary cosmology”, was to takeinto consideration the consequences of the unified quantum field theories, thegauge theories, at the construction of relativistic models for the early universe.According to the Friedmann models, the temperature was extremely high in theearly history of the universe. If one considers Higgs fields associated with GUTmodels (grand unified theories), one finds a critical temperature Tc correspond-ing to the energy kTc = 1014GeV , where k is Boltzmann’s constant. Beforethe universe was about t1 = 10−35s old, the temperature was larger than this.Thus, the Higgs field was in the symmetric ground state. According to most ofthe inflation models, the universe was dominated by radiation at this time.

When the temperature decreases, the Higgs potential changes. This couldhappen as shown in figure 10.17. Here, there is a potential barrier at the critcaltemperature, 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.

Page 178: LectureNotes On GeneralRelativity

162 Chapter 10. Cosmology

Figure 10.17: The temperature dependence of a Higgs potential with a firstorder phase transition.

Guth’s Inflation Model

Alan Guth’s original inflation model (Guth 1981) was based on a first orderphase transition.

According to most of the inflationary models, the universe was dominatedby radiation during the time before 10−35s. The universe was then expanding sofast that there was no causal contact between the different parts of the universethat became our observable universe. Probably, the universe was rather homo-geneous, with considerable spacelike variations in temperature. There was alsoareas of false vacuum, with energy densities characteristic of the GUT energyscale, which also controls it’s critical temperature. While the energy densityof the radiation decreased quickly, as a−4, the energy density of vacuum wasconstant. At the time t = 10−35s, the energy density of the radiation becameless than that of the vacuum.

At the same time, the potential started to change, such that the vacuum wentfrom being stable to being an unstable false vacuum. Thus, there was a firstorder phase transition to the real vacuum. Because of the inhomogeneouty ofthe universe’s initial condition, this happened with different speed at differingplaces. The potential barrier slowed down the process, which happened bytunneling, and the universe was at several places considerably undercooled, andthere appeared “bubbles” dominated by the energy of the false vacuum. Theseareas acted on themselves with repulsive gravity.

By integrating the equation of motion for the expansion factor in such avacuum dominated bubble, one gets

a = eHt, H =

√8πGρc

3. (10.126)

Page 179: LectureNotes On GeneralRelativity

10.5 Inflationary Cosmology 163

By inserting the GUT value above, we get H = 6.6 · 1034s−1, i.e. H−1 =1.5 · 10−35s. With reference to field theoretical works by Sindney Coleman andothers, Guth argumented that a realistic duration of the nucleation process hap-pening during the phase transition is 10−33s. During this time, the expansionfactor increases by a factor of 1028. This vacuum dominated epoch is called theinflation era.

Let us look closer at what happens with the energy of the universe in thecourse of it’s development, according to the inflationary models. To understandthis 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 oscillatingfield, the energy of the false vacuum is converted into radiation and particles. Inthis way the equation of state for the energy dominating the development of theexpansion factor changes from p = −ρ, characteristic for vacuum, to p = 1

3ρ,characteristic of radiation.

The energy density and the temperature of the radiation is then increasedenourmously. Before and after this short period around the time t = 10−33s theradiation energy increases adiabatically, such that ρa4 = constant. According toStefan-Boltzmanns law of radiadion, ρ ∝ T 4. Therefore, aT = constant duringadiabatic expansion. This means that during the inflationary era, while theexpansion factor increases exponentially, the energy density and temperatureof radiation decreases exponentially. At the end of the inflationary era, theradiation is reheated so that it returns to the energy it had when the inflationaryera started.

It may be interesting to note that the Newtonian theory of gravitation doesnot allow an inflationary era, since stress has no gravitational effect accordingto it.

The Inflation Models’ Answers to the Problems of the FriedmannModels

The horizon problem will here be investigated in the light of this model. Theproblem was that there was room for about 10000 causally connected areasinside the area spanned by our presently observable universe at the time. Let uscalculate the horizon radius lh and the radius a of the region presently withinthe horizon, lh = 15 · 109ly = 1.5 · 1026cm, at the time t1 = 10−35s when theinflation started. From equation (10.118) for the radiation dominated periodbefore the inflatinary era, one gets

lh = 2t1 = 6 · 10−25cm. (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 = 10−35s tot2 = 10−33s, a ∝ t

12 in the radiation dominated period from t2 to t3 = 1011s,

and a ∝ t 23 in the matter dominated period from t3 until now, t0 = 1017s. This

Page 180: LectureNotes On GeneralRelativity

164 Chapter 10. Cosmology

gives

a1 =eHt1

eHt2

(t2t3

) 12(t3t0

) 23

lh(t0) = 1.5 · 10−28cm. (10.128)

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 observableuniverse. The whole of this region was then causally connected, and thermicequilibrium was established. This equilibrium has been kept since then, andexplains 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 obtainthe closeness of the observed mass density to the critical density. Again, theinflationary models give another result. Inserting the expansion factor (10.126)into equation (10.122), we get

Ω− 1 =k

H2e−2Ht, (10.129)

where H is constant and given in eq. (10.126). The ratio between Ω− 1 at theend of and the beginning of the inflationary era becomes

Ω2 − 1

Ω1 − 1= e−2H(t2−t1) = 10−56. (10.130)

Contrary to in the Friedmann models, where the mass density moves awayfrom the critical density as time is increasing, the density approaches the criticaldensity exponentially during the inflationary era. Within a large range of initialconditions, this means that according to the inflation models the universe shouldstill have almost critical mass density.

Page 181: LectureNotes On GeneralRelativity

Bibliography

Bernadis et.al, P. d.: 2001, Multiple peaks in the angular power spectrum ofthe 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, 3067–3070.

Carroll, S. M.: 2001, Dark energy and the preposterous universe., astro-ph/0107571.

Grøn, Ø.: 1986, Repulsive gravitation and inflationary universe models., 54, 46–52.

Grøn, Ø. and Flø, H. K.: 1984, Differensialgeometri, Departmental. Coursecompendium.

Guth, A. H.: 1981, The inflationary universe: a possible solution to the horizonand flatness problems, Phys. Rev. D23, 347–356.

Linde, A.: 2001, Inflation and string cosmology, astro-ph/0107176.

McGaugh, S.: 2001, Constraints on the radial mass distribution of dark matterhalos 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, 565–586.

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 seasonof observations with dasi., astro-ph/0104490.

Ravndal, F.: 1978, Inføring i generell relativitetsteori, Departmental. Coursecompendium.

Riess, A. G.: n.d., The farthest known supernova: Support for an acceleratinguniverse and a glimpse of the epoch of deceleration., astro-ph/0104455.

165

Page 182: LectureNotes On GeneralRelativity

166 BIBLIOGRAPHY

Riess et al., A. G.: 1998, Observational evidence from supernovae for an accel-erating universe and a cosmological constant., Astron. J. 116, 1009–1038.

Stompor et al., R.: 2001, Cosmological implications of the maxima-i highresolution 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 pastdeceleration of the universe?, astro-ph/0106051.

Zeldovich, Y.: 1968, The cosmological constant and the theory of elementaryparticles, Sov. Phys. Usp. 11, 381–393.

Zlatev, I., Wang, L. and Steinhardt, P. J.: 1999, Quintessence, cosmic coinci-dence, and the cosmological constant, Phys. Rev. Lett. 82, 896–899.

Page 183: LectureNotes On GeneralRelativity

Index

Symbols, ν1ν2 . . . . . . . . . . . . . . . . . . . . . . . . . . . .50δS = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . 59δ(~r) (Dirac delta function) . . . . . . . 4γij . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34⊗ (tensor product) . . . . . . . . . . . . . . 26d2α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51dΩ (solid angle) . . . . . . . . . . . . . . . . . . 5dt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35dl2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34[ ] (antisymmetric combination) . 32Γαµν . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613-space. . . . . . . . . . . . . . . . . . . . . . . . . .344-acceleration . . . . . . . . . . . . . . 16, 1204-momentum. . . . . . . . . . . . . . . . . . . . 154-velocity . . . . . . . . . . . . . . . . . . . . . . . 15

identity . . . . . . . . . . . . . . . . . 16, 59

Aaccelerated frame

hyperbolically . . . . . . . . . . . . . . . 44acceleration of particle

in non-rotating frame . . . . . . . 56in rotating frame. . . . . . . . . . . .57

acceleration scalar . . . . . . . . . . . . . 120action . . . . . . . . . . . . . . . . . . . . . . . . . . . 58Archimedean spiral . . . . . . . . . . . . . .41

Bbasis

1-form . . . . . . . . . . . . . . . . . . . . . . 25coordinate basis vector . . 18, 21orthonormal . . . . . . . . . . . . . 18, 22

Bianchi1st identity . . . . . . . . . . . . . . 90, 92component form . . . . . . . . . . 91

2nd identity . . . . . . . . . . . . . 91, 95black hole

horizon . . . . . . . . . . . . . . . . . . . . 120

mini . . . . . . . . . . . . . . . . . . . . . . . 122radiation. . . . . . . . . . . . . . . . . . .121temperature . . . . . . . . . . . . . . . 121

Born-stiff motion . . . . . . . . . . . . . . . . 44

Ccanonical momentum. . . . . . . . . . . . 59Cartan

formalism. . . . . . . . . . . . . . . . . . 100structure equations1st . . . . . . . . . . . . . . . . . . . 79, 1002nd . . . . . . . . . . . . . . . . . . 85, 101

surface curvature. . . . . . . . . . . .89centrifugal acceleration . . . . . . . . . . 57centrifulgal barrier . . . . . . . . . . . . . 112Christoffel symbols . . . . . . . . . . . . . . 61

in coordinate basis . . . . . . . . . . 53physical interpretation . . . . . 103rotating refernece frame . . . . . 74

comoving orthonormal basis . . . . . 31connection coefficients

in a Riemannian space . . . . . . 74in coordinate basis . . . . . . . . . . 53rotating reference frame . . . . . 73

connection forms . . . . . . . . . . . . . . . 100conservation

energy-momentum . . . . . . . . . . 93constant of motion . . . . . . . . . . . . . . 59continuity equation. . . . . . . . . . . . . .93contravariant tensor . . . . . . . . . . . . . 26convective derivative . . . . . . . . . . . . 93coordinate

cyclic . . . . . . . . . . . . . . . . . . . 59, 60Gaussian. . . . . . . . . . . . . . . . . . . .54transformation . . . . . . . . . . 20, 27

coordinate basis . . . . . . . . . . . . . 18, 21connection coefficients. . . . . . .53

coordinate system . . . . . . . . . . . . . . . 18

167

Page 184: LectureNotes On GeneralRelativity

168 INDEX

coriolis acceleration . . . . . . . . . . . . . 57cosmological redshift . . . . . . . . . . . 135covariance principle. . . . . . . . . .11, 51covariant equations. . . . . . . . . . . . . .28covariant tensor . . . . . . . . . . . . . . . . . 26curvature . . . . . . . . . . . . . . . . . . . . . . . 81

forms . . . . . . . . . . . . . . . . . . 85, 101of surface . . . . . . . . . . . . . . . . . 89

Gaussian . . . . . . . . . . . . . . . . 88, 89geodesic . . . . . . . . . . . . . . . . . . . . 87normal . . . . . . . . . . . . . . . . . . . . . . 87principal . . . . . . . . . . . . . . . . . . . . 88Ricci . . . . . . . . . . . . . . . . . . . . . . . . 91Riemann tensor . . . . . . . . . . . . . 83surface (Cartan formalism) . .89

Ddeflection of light . . . . . . . . . . . . . . 117differential forms

1-form . . . . . . . . . . . . . . . . . . . . . . 252-form . . . . . . . . . . . . . . . . . . . . . . 32components . . . . . . . . . . . . . . . . . 33connection . . . . . . . . . . . . . . . . . 100covariant differentiation of . . 76curvature . . . . . . . . . . . . . . 85, 101of surface . . . . . . . . . . . . . . . . . 89

differentiation of . . . . . . . . . . . . 49directional derivatives of . . . . 76null form. . . . . . . . . . . . . . . . . . . .33p-form . . . . . . . . . . . . . . . . . . . . . . 32transformation . . . . . . . . . . . . . . 27

differentiationconvective . . . . . . . . . . . . . . . . . . 93covariant . . . . . . . . . . . . . . . . 53, 55exterior . . . . . . . . . . . . . . . . . . . . . 49of form . . . . . . . . . . . . . . . . . . . . . 49

Dirac delta function . . . . . . . . . . . . . . 4distance

coordinate . . . . . . . . . . . . . . . . . 100physical. . . . . . . . . . . . . . . . . . . .100

Doppler effect . . . . . . . . . . . . . . . . . . . 23Doppler effect, gravitational . . . . . 71dust dominated universe . . . . . . . 142

EEddington - Finkelstein coordinates

104

Einsteinfield equations . . . . . . . . . . . . . . 96principle of equivalence. . . . . . .9synchronization . . . . . . . . . . . . . 35tensor. . . . . . . . . . . . . . . . . . . . . . .96

Einstein deSitter universe . . . . . . 144embedding

of the Schwarzschild metric 108energy-momentum tensor. . . . . . . .93equivalence, principle of . . . . . . . . . . 9Euler equation of motion . . . . . . . . 93Euler-Lagrange equations . . . . . . . 59exterior differentiation. . . . . . . . . . .49

Ffield equations

Einstein . . . . . . . . . . . . . . . . . . . . 96vacuum . . . . . . . . . . . . . . . . . . . . . 97

first order phase transition . . . . . 160flatness problem. . . . . . . . . . . 158, 163forms . . . . . . . . . see differential formsfree fall . . . . . . . . . . . . . . . . . . . . . . . . 104

GGauss

coordinates . . . . . . . . . . . . . . . . . 54curvature. . . . . . . . . . . . . . . .88, 89integral theorem . . . . . . . . . . . . . 4theorema egregium. . . . . . . . . . 89

general principle of relativity . . . . 10geodesic

curvature . . . . . . . . . . . . . . . . . . . 87curve. . . . . . . . . . . . . . . . . . . . 57, 61spatial . . . . . . . . . . . . . . . . . . . . 65

equation . . . . . . . . . . . . . . . . . . . . 56postulate . . . . . . . . . . . . . . . . . . . 98

gravitational acceleration. . . . . . .103Tolman-Whittaker expression128

gravitational collapse. . . . . . . . . . .112gravitational time dilation. . . . . . . 40Grøn, Øyvind.. . . . . . . . . . . . . . . . . . . .3Guth, Alan . . . . . . . . . . . . . . . . . . . . 161

HHafele-Keating experiment . . . . . 114Hamilton-principle . . . . . . . . . . . . . . 58Hawking radiation . . . . . . . . . . . . . 121

Page 185: LectureNotes On GeneralRelativity

INDEX 169

heat capacity of black hole . . . . . 121Higgs mechanism . . . . . . . . . . . . . . 159horizon problem. . . . . . . . . . . 156, 162Hubbles law. . . . . . . . . . . . . . . . . . . .135hydrostatic equilibrium . . . . . . . . 127hyperbolically accelerated frame. 44

Iincompressible stars . . . . . . . . . . . . 130inertial reference frame . . . . . . . . . . 18inertial dragging . . . . . . . . . . . . . . . 123inertial reference frame . . . . . . . . . . 54inflation . . . . . . . . . . . . . . . . . . . . . . . 161

KKerr metric . . . . . . . . . . . . . . . . . . . . 122Kretschmann’s curvature scalar 104Kruskal - Szekers. . . . . . . . . . . . . . .104

LLagrange

covariant dynamics. . . . . . . . . .57equations . . . . . . . . . . . . . . . . . . . 59free particle . . . . . . . . . . . . . . . 61

function . . . . . . . . . . . . . . . . . . . . 58free particles . . . . . . . . . . . . . . 61

Laplace equation . . . . . . . . . . . . . . . . . 3Levi-Civita connection . . . . . . . . . . 53light

deceleration of . . . . . . . . . . . . . 108deflection of . . . . . . . . . . . . . . . 117

light cones . . . . . . . . . . . . . . . . . . . . . 106line element . . . . . . . . . . . . . . . . . . . . . 31

free particle . . . . . . . . . . . . . . . . . 60in space . . . . . . . . . . . . . . . . . . . . . 34spacetime . . . . . . . . . . . . . . . . . . . 35

Lorentz transformation . . . . . . . . . . 23

MMach’s principle . . . . . . . . . . . . . . . . 12metric

covariant differentiation . . . . . 77Kerr (rotating) . . . . . . . . . . . . 122static,stationary . . . . . . . . . . . . 60tensor. . . . . . . . . . . . . . . . . . . . . . .28

mini black holes. . . . . . . . . . . . . . . .122Minkowski space-time

from Schwarzschild solution 103mixed tensor . . . . . . . . . . . . . . . . . . . . 26momentum-energy tensor. . . . . . . .93

NNewton

fluid . . . . . . . . . . . . . . . . . . . . . . . . 94hydrodynamics. . . . . . . . . . . . . .93law of gravitation . . . . . . . . . . . . 1continuous distribution. . . . . 2local form. . . . . . . . . . . . . . . . . . 2

Newtonian incompressible star . 126

Oorbit equation. . . . . . . . . . . . . . . . . .113orthonormal basis . . . . . . . . . . . . . . . 18

Pparallel transport . . . . . . . . . . . . 55, 81particle trajectory

Schwarzschild 3-space . . . . . . 110Poincaré’s lemma . . . . . . . . . . . . . . . 51precession

of Mercury’s perihelion . . . . 115pressure forces . . . . . . . . . . . . . . . . . 127principle of equivalence . . . . . . 9, 103projectile motion . . . . . . . . . . . . . . . . 68proper distance . . . . . . . . . . . . . . . . 135

Rradiation dominated universe. . .141rank . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26rapidity . . . . . . . . . . . . . . . . . . . . . . . . . 45redshift

cosmological . . . . . . . . . . . . . . . 135Kerr spacetime . . . . . . . . . . . . 123

reference frame. . . . . . . . . . . . . . . . . . 17inertial . . . . . . . . . . . . . . . . . . 18, 54

relativity, general principle of. . . .10Ricci

curvature . . . . . . . . . . . . . . . . . . . 91tensor, divergence . . . . . . . . . 96

identity . . . . . . . . . . . . . . . . . . . . . 90component form . . . . . . . . . . 90

Riemann curvature tensor . . 83, 101components . . . . . . . . . . . . . . . . . 84

rotating black hole . . . . . . . . . . . . . 122

Page 186: LectureNotes On GeneralRelativity

170 INDEX

rotating reference frame . . . . . . . . . 65connection coefficients. . . . . . .73

SSagnac effect . . . . . . . . . . . . . . . . . . . . 42Schwarzschild

3-space . . . . . . . . . . . . . . . . . . . . 110energy . . . . . . . . . . . . . . . . . . . 110momentum . . . . . . . . . . . . . . 110

light cones . . . . . . . . . . . . . . . . . 106metric . . . . . . . . . . . . . . . . . . . . . 104embedding. . . . . . . . . . . . . . . 108

radius . . . . . . . . . . . . . . 2, 104, 120solutioninterior . . . . . . . . . . . . . . . . . . 130outer . . . . . . . . . . . . . . . . . . . . 100

spacetime . . . . . . . . . . . . . 104, 120simultaneity

on Einstein synchronized clocks35

space, geometry of . . . . . . . . . . 35singularity

coordinate . . . . . . . . . . . . . . . . . 104physical. . . . . . . . . . . . . . . . . . . .104

solid angle . . . . . . . . . . . . . . . . . . . . . . . 4spherical geometry . . . . . . . . . . . . . . 38spontaneous symmetry breaking160structure coefficients . . . . . . . . . . . . 23

in a Riemannian space . . . . . . 74in coordinate basis . . . . . . . . . . 25

surface curvature . . . . . . . . . . . . . . . . 89surface gravity . . . . . . . . . . . . . . . . . 120

Ttangent plane . . . . . . . . . . . . . . . . . . . 17tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

antisymmetric . . . . . . . . . . . . . . 31construction example . . . . . . . 27contraction of . . . . . . . . . . . . . . . 91contravariant . . . . . . . . . . . . . . . 26covariant. . . . . . . . . . . . . . . . . . . .26covariant differentiation of . . 77Einstein . . . . . . . . . . . . . . . . . . . . 96metric . . . . . . . . . . . . . . . . . . . . . . 28mixed . . . . . . . . . . . . . . . . . . . 26, 28momentum energy . . . . . . . . . . 93

product . . . . . . . . . . . . . . . . . . . . . 26rank . . . . . . . . . . . . . . . . . . . . . . . . 26Ricci curvature. . . . . . . . . . . . . .91Riemann curvature . . . . . 83, 101transformation properties . . . 27

theorema egregium . . . . . . . . . . . . . . 89tidal forces . . . . . . . . . . . . . . . . . . . . . . . 5time dilation, gravitational . . . . . . 40Tolman-Oppenheimer-Volkov eq.130Tolman-Whittaker expression . . 128travelling time . . . . . . . . . . . . . . . . . 105Twin “paradox” . . . . . . . . . . . . . . . . . 69

Uuniverse model

dust dominated . . . . . . . . . . . . 142inflationary . . . . . . . . . . . . . . . . 161radiation dominated . . . . . . . 141

Vvacuum field equations . . . . . . . . . . 97vector

basis . . . . . . . . . . . . . . . . . . . . . . . . 14covariant differentiation of . . 75dimension. . . . . . . . . . . . . . . . . . . 13linearly independent . . . . . . . . 13

ZZAMO. . . . . . . . . . . . . . . . . . . . . . . . . 123