General Relativity: Matthias Bartelmann
General Relativity: Matthias Bartelmann
Physik LN
General Relativity
MATTHIAS BARTELMANN
HEIDELBERG
UNIVERSITY PUBLISHING
GENER AL RELATIVITY
Lecture Notes Physik LN
GENERAL RELATIVITY
Matthias Bartelmann
Instead of a preface
Before we begin, some words are in order on the purpose and the limits
of these notes, on the notation used, and on some of the many people I
am indebted to.
Each of the chapters of these notes is meant for a week of two lectures
of two hours each. Much more could be said about all of their topics in
all directions, in terms of mathematics, physics, and experimental tests
of general relativity. These notes are meant as an introduction which
can in no way be considered complete. They may serve as a first guide
through the subject, not a comprehensive one. These lectures are part of
a curriculum in which cosmology, gravitational lensing, and theoretical
astrophysics are regularly taught separately. In these areas, they are thus
only meant to lay the foundation.
We use index-free notation where possible and convenient. Then, the
curvature, the curvature tensor, the Ricci tensor and the Ricci scalar,
often denoted with an R with different numbers of indices, need different
symbols. We denote the curvature and the curvature tensor, closely
related as they are, with R̄, the Ricci tensor with R and the Ricci scalar
with R. Since the symbol G is then reserved for the Einstein tensor, we
write Newton’s gravitational constant as G.
Indices refering to coordinates on general, d-dimensional manifolds are
written as Latin characters. On 4-dimensional, spacetime manifolds,
Greek indices run from 0 to 3, while Latin indices refer to spatial coordi-
nates and run from 1 to 3.
These lecture notes grew over several years. Many students were ex-
posed to this lecture and contributed corrections and suggestions that
greatly helped improving it. In particular, Dr. Christian Angrick and
Dr. Francesco Pace were kind and patient enough to meticulously work
through the entire notes and point out many mistakes. Thank you all
very much!
Particular thanks are due to the wonderful and inspiring teachers I myself
had on general relativity. Jürgen Ehlers always impressed me with his
depth and clarity of thinking, and Norbert Straumann introduced me to
the elegance and beauty of the theory.
Contents
1 Introduction 1
1.1 The idea behind general relativity . . . . . . . . . . . 1
1.2 Fundamental properties of gravity . . . . . . . . . . . 4
1.3 Consequences of the equivalence principle . . . . . . . 7
1.4 Futile attempts . . . . . . . . . . . . . . . . . . . . . 9
2 Differential Geometry I 15
2.1 Differentiable manifolds . . . . . . . . . . . . . . . . 15
2.2 The tangent space . . . . . . . . . . . . . . . . . . . . 18
2.3 Dual vectors and tensors . . . . . . . . . . . . . . . . 24
2.4 The metric . . . . . . . . . . . . . . . . . . . . . . . . 27
3 Differential Geometry II 31
3.1 Connections and covariant derivatives . . . . . . . . . 31
3.2 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 Curvature . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4 Riemannian connections . . . . . . . . . . . . . . . . 42
vii
viii CONTENTS
A Electrodynamics 213
A.1 Electromagnetic field tensor . . . . . . . . . . . . . . 213
A.2 Maxwell’s equations . . . . . . . . . . . . . . . . . . 213
A.3 Lagrange density and energy-momentum tensor . . . . 214
Introduction
There was no need for general relativity when Einstein started working
on it. There was no experimental data signalling any failure of the
Newtonian theory of gravity, except perhaps for the minute advance of
the perihelion of Mercury’s orbit by 43 per century, which researchers
at the time tried to explain by perturbations not included yet into the
calculations of celestial mechanics in the Solar System.
Essentially, Einstein found general relativity because he was deeply
dissatisfied with some of the concepts of the Newtonian theory, in par-
ticular the concept of an inertial system, for which no experimental
demonstration could be given.
After special relativity, he was convinced quite quickly that trying to
build a relativistic theory of gravitation led to conclusions which were in
conflict with experiments. Action at a distance is impossible in special
relativity because the absolute meaning of space and time had to be
given up. The most straightforward way to combine special relativity
with Newtonian gravity seemed to start from Poisson’s equation for the
gravitational potential and to add time derivatives to it so as to make it
relativistically invariant.
However, it was then unclear how the law of motion should be modified
because, according to special relativity, energy and mass are equiva-
lent and thus the mass of a body should depend on its position in a
gravitational field.
This led Einstein to a result which raised his suspicion. In Newtonian
theory, the vertical acceleration of a body in a vertical gravitational
field is independent of its horizontal motion. In a special-relativistic
extension of Newton’s theory, this would no longer be the case: the
1
2 1 Introduction
This was in striking conflict with experiment, which says that all bodies
experience the same gravitational acceleration. At this point, the equiva-
lence of inertial and gravitational mass struck Einstein as a law of deep
significance. It became the heuristic guiding principle in the construction
of general relativity.
Freely falling frames of reference
This line of thought leads to the fundamental concept of general rela-
tivity. It says that it must be possible to introduce local, non-rotating,
freely-falling frames of reference in which gravity is locally “trans-
formed away”.
The directions of motion of different freely-falling reference frames will
generally not be parallel: Einstein elevators released at the same height
above the Earth’s surface but over different locations will fall towards
the Earth’s centre and thus approach each other.
Space-time as a manifold
Replacing inertial frames by freely falling, non-rotating frames of ref-
erences leads to the idea that spacetime is a four-dimensional manifold
instead of the “rigid”, four-dimensional Euclidean space.
1.1 The idea behind general relativity 3
g
g
Figure 1.2 Einstein elevators: The left elevator is thought to be placed
outside a gravitational field, but accelerated upwards with an acceleration
−g; the right elevator is placed at rest in a gravitational field with gravitational
acceleration g directed downwards. According to the equivalence principle,
their occupants cannot distinguish these situations from each other.
1.2.1 Scales
Gm2 !
= mc2 . (1.3)
λ
The result is the Planck mass,
c GeV
m = MPl = = 2.2 · 10−5 g = 1.2 · 1019 2 , (1.4)
G c
Gm2p 2
1 mp
= , (1.7)
e2 α MPl
2
which can be cast into the more precise form that in an arbitrary gravi-
tational field, no local non-gravitational experiment can distinguish a
freely falling, non-rotating system from a uniformly moving system in
absence of the gravitational field.
6 1 Introduction
Without any specific form of the theory, the equivalence principle imme-
diately allows us to draw conclusions on some of the consequences any
theory must have which is built upon it. We discuss two here to illustrate
its general power, namely the gravitational redshift and gravitational
light deflection.
Δv = 12 gt2
h g
v=0
w
Figure 1.3 Two Einstein elevators, both outside a gravitational field and
accelerated upwards with acceleration g. When the photon reaches the
top of the elevator (left) or while the light ray crosses it (right), the elevator
is accelerated to the velocity Δv = gt2 /2. This leads to redshift (left) and
aberration (right).
1.4 Futile attempts 9
Δv |∇Φ|w
Δα = = (1.16)
c c2
downward from the horizontal because of the aberration due to the finite
light speed.
Light deflection by gravitational fields
Since the upward accelerated elevator corresponds to an elevator at
rest in a downward gravitational field, this leads to the expectation that
light will be deflected towards gravitational fields.
Although it is possible to construct theories of gravity which obey the
equivalence principle and do not lead to gravitational light deflection,
the bending of light in gravitational fields is by now a well-established
experimental fact.
the proper time measured by observers moving along a world line x μ (λ)
from λ1 to λ2 is
λ2
Δτ = dλ −ημν ẋ μ ẋ ν , (1.17)
λ1
where the minus sign under the square root appears because we choose
the signature of ημν to be (−1, 1, 1, 1).
Now, let a light ray propagate from the floor to the ceiling of the elevator
in which we have measured gravitational redshift before. Specifically, let
the light source shine between coordinates times t1 and t2 . The emitted
photons will propagate to the receiver at the ceiling along world lines
which may be curved, but must be parallel because the metric is constant.
The time interval within which the photons arrive at the receiver must
thus equal the time interval t2 − t1 within which they left the emitter.
Thus there cannot be gravitational redshift in a theory of gravity in flat
spacetime.
Let us now try and construct a scalar theory of gravity starting from the
field equation
φ = −4πGT , (1.18)
since then the time derivatives in d’Alembert’s operator and the pressure
contributions to T can be neglected.
Let us further adopt the Lagrangian
L(x μ , ẋ μ ) = −mc −ημν ẋ μ ẋ ν (1 + φ) , (1.20)
√
−ημν ẋ μ ẋ ν = c2 − v 2 = c 1 − β 2 , (1.21)
1.4 Futile attempts 11
(1 + φ)x¨ = −c2 ∇φ (1.29)
or
¨x = −c2 ∇φ(1 2 φ2
− φ) = −c ∇ φ − . (1.30)
2
Compared to the equation of motion in Newtonian gravity, therefore, the
potential is augmented by a quadratic perturbation.
12 1 Introduction
L2
VL = V + , (1.32)
Caution Recall that the equa- 2mr2
tion of motion (1.31) follows
and L is the (orbital) angular momentum. Thus,
from the conservation laws of
angular-momentum,
dr r2 L2
= 2m(E − V) − 2 . (1.33)
L dϕ L r
ϕ̇ = ,
mr2
The perihelion shift is the change in ϕ upon integrating once around the
and energy,
orbit, or integrating twice from the perihelion radius r0 to the aphelion
2 radius r1 , r1
ṙ2 = (E − VL (r)) . dϕ
m Δϕ = 2 dr . (1.34)
r0 dr
Inverting (1.33), it is easily seen that (1.34) can be written as
r1
∂ L2
Δϕ = −2 dr 2m(E − V) − 2 . (1.35)
? ∂L r0 r
Confirm (1.35) by carrying out
the calculation yourself. Now, we split the potential energy V into the Newtonian contribution V0
and a perturbation δV V and expand the integrand to lowest order in
δV, which yields
r1
∂ mδV
Δϕ ≈ −2 dr A0 1 − (1.36)
∂L r0 A0
Δϕ = − . (1.45)
ac2 (1 − e2 )
For the Sun, M
= 2 · 1033 g, thus
GM
Differential Geometry I
15
16 2 Differential Geometry I
such as
f1+ : U1+ → D23 , f1+ (x1 , x2 , x3 ) = (x2 , x3 ) . (2.5)
Thus, the six charts fi± ,
i ∈ {1, 2, 3}, together form an atlas of the
two-sphere. See Fig. 2.1 for an illustration.
(p1 , p2 ) ∈ D12
p ∈ U3+
Figure 2.1 Example for a chart, explained in the text: the point p on the
half-sphere U3+ the two-sphere is projected into the domain D12 ⊂ R2 .
In other words, pairs of points from the product manifold are mapped to
pairs of points from the two open subsets M ⊂ Rm and N ⊂ Rn .
and a multiplication,
Tangent space
Generally, the tangent space T p M of a differentiable manifold M at a
point p is the set of derivations of F (p). A derivation v is a map from
F (p) into R,
v : F (p) → R , (2.12)
which is linear,
v(λ f + μg) = λv( f ) + μv(g) (2.13)
for f, g ∈ F (p) and λ, μ ∈ R, and satisfies the product rule (or Leibniz
rule)
v( f g) = v( f )g + f v(g) . (2.14)
See Fig. 2.2 for an illustration of the tangent space to a 2-sphere.
Tp M
for v, w ∈ T p M, f ∈ F and λ ∈ R.
Note the equality! This is not a Taylor expansion. This is easily seen
using the identity
1
d
F(x) − F(0) = F(tx1 , . . . , txn )dt
0 dt
n 1
= x i
Di F(tx1 , . . . , txn )dt , (2.18)
i=1 0
where we have used that v applied to the constant f (p) vanishes, that
(xi ◦ h)(p) = 0 and that Hi (0) = ei ( f ) according to (2.22). Thus, setting
vi = v(xi ◦ h), we find that any v ∈ T p M can be written as a linear
combination of the basis vectors ei . This also demonstrates that the
dimension of the tangent space T p M equals that of the manifold itself.
Coordinate basis of T p M
The basis {ei }, which is often simply denoted as {∂/∂xi } or {∂i }, is called
a coordinate basis of T p M. Vectors v ∈ T p M can thus be written as
v = vi ei = vi ∂i . (2.24)
which shows that the two different coordinate bases are related by the
Jacobian matrix of the coordinate change, which has the elements
Ji j = ∂x j /∂xi . Its inverse has the elements J ij = ∂xi /∂x j .
and thus it is called the derivative of f with respect to the vector field v.
Example: Transformation of S 2
In our example on S 2 , we fix a point p on the sphere whose orbit under
the map γt is a part of the “latitude circle” through p. The tangent
vector to this curve in p defines the local “direction of motion” under
the rotation expressed by γt . Applying this to all points p ∈ S 2 defines
a vector field v on S 2 .
[v + w, x] = [v, x] + [w, x]
[v, w] = −[w, v]
[ f v, gw] = f g[v, w] + f v(g)w − gw( f )v
[v, [w, x]] + [x, [v, w]] + [w, [x, v]] = 0 , (2.33)
df : TM → R , d f (v) = v( f ) . (2.35)
which shows that the n-tuple {e∗i } = {dxi } forms a basis of T ∗ M, which
is called the dual basis to the basis {ei } = {∂i } of the tangent space T M.
Dual vectors
Dual vectors map vectors to the real numbers. If {∂i } is a coordinate ba-
sis of T M, the dual basis of T ∗ M is given by the coordinate differentials
{dxi }. Dual vectors can thus be written as
w = wi dxi . (2.38)
Starting the same operation leading from T M to the dual space T ∗ M
with T ∗ M instead, we arrive at the double-dual vector space T ∗∗ M as the
vector space of all linear maps from T ∗ M → R. It can be shown that
T ∗∗ M is isomorphic to T M and can thus be identified with T M.
2.3.2 Tensors
With the obvious rules for adding linear maps and multiplying them with
scalars, the set of tensors T sr of rank (r, s) attains the structure of a vector
space of dimension nr+s .
26 2 Differential Geometry I
Given a tensor t of rank (r, s) and another tensor t of rank (r , s ), we can
construct a tensor of rank (r + r , s + s ) called the outer product t ⊗ t of
t and t by simply multiplying their results on the r + r dual vectors wi
and the s + s vectors v j , thus
(t ⊗ t )(w1 , . . . , wr+r , v1 , . . . , v s+s ) = (2.40)
r+r
t(w , . . . , w , v1 , . . . , v s ) t (w
1 r r+1
,...,w , v s+1 , . . . , v s+s ) .
C ij t : T sr → T s−1
r−1
, C ij t = t(. . . , e∗k , . . . , ek , . . .) , (2.43)
where {ek } and {e∗k } are bases of the tangent and dual spaces, as before,
and the summation over all 1 ≤ k ≤ n is implied. The basis vectors e∗k
and ek are inserted as the i-th and j-th arguments of the tensor t.
Expressing the tensor in a coordinate basis, we can write the tensor in
the form (2.41), and thus its contraction with respect to the ia -th and
jb -th arguments reads
C ijab t = tij11...ir ik
... j s dx (∂ jk )
# $
∂i1 ⊗ . . . ⊗ ∂ia −1 ⊗ ∂ia +1 ⊗ . . . ⊗ ∂ir
dx j1 ⊗ . . . ⊗ dx jb −1 ⊗ dx jb +1 ⊗ . . . ⊗ dx js
= tij11...ia−1 ik ia+1 ...ir
... jb−1 ik jb+1 ... j s
# $
∂i1 ⊗ . . . ⊗ ∂ia −1 ⊗ ∂ia +1 ⊗ . . . ⊗ ∂ir
dx j1 ⊗ . . . ⊗ dx jb −1 ⊗ dx jb +1 ⊗ . . . ⊗ dx js . (2.44)
2.4 The metric 27
We need some way to define and measure the “distance” between two
points on a manifold. A metric is introduced via the infinitesimal squared
distance between two neighbouring points on the manifold.
We have seen above that tangent vectors v ∈ T p M are closely related to
infinitesimal displacements around a point p on the manifold. Moreover,
the infinitesimal squared distance between two neighbouring points p
and q should be quadratic in the displacement leading from one point to
the other. Thus, we construct the metric g as a bi-linear map
g : TM × TM → R , (2.47)
which means that the g is a tensor of rank (0, 2). The metric thus assigns
a number to two elements of a vector field T M on M. The metric g
thus defines to two vectors their scalar product, which is not necessarily
positive. We abbreviate the scalar product of two vectors v, w ∈ T M by
g(v, w) ≡ v, w . (2.48)
Metric
A metric is a rank-(0, 2) tensor field which is symmetric and non-
degenerate.
In a coordinate basis, the metric can be written in components as
g = gi j dxi ⊗ dx j . (2.50)
28 2 Differential Geometry I
Given a coordinate basis {ei }, the metric g can always be chosen such
that
g(ei , e j ) = ei , e j = ±δi j , (2.52)
where the number of positive and negative signs is independent of the
coordinate choice and is called the signature of the metric. Positive-
(semi-) definite metrics, which have only positive signs, are called Rie-
mannian, and pseudo-Riemannian metrics have positive and negative
signs.
Given a tangent vector v, the metric can also be seen as a linear map
from T M into T ∗ M,
Differential Geometry II
The curvature of the two-dimensional sphere S 2 can be described by Caution Whitney’s (strong)
embedding the sphere into a Euclidean space of the next-higher dimen- embedding theorem states that
sion, R3 . However, (as far as we know) there is no natural embedding any smooth n-dimensional mani-
of our four-dimensional curved spacetime into R5 , and thus we need a fold (n > 0) can be smoothly
description of curvature which is intrinsic to the manifold. embedded in the 2n-dimensional
Euclidean space R2n . Embed-
There is a close correspondence between the curvature of a manifold and
dings into lower-dimensional Eu-
the transport of vectors along curves.
clidean spaces may exist, but not
As we have seen before, the structure of a manifold does not trivially necessarily so. An embedding
allow to compare vectors which are elements of tangent spaces at two f : M → N of a manifold M into
different points. We will thus have to introduce an additional structure a manifold N is an injective map
which allows us to meaningfully shift vectors from one point to another such that f (M) is a submanifold
on the manifold. of N and M → f (M) is differen-
tiable.
Even before we do so, it is intuitively clear how vectors can be trans-
ported along closed paths in flat Euclidean space, say R3 . There, the
vector arriving at the starting point after the transport will be identical to
the vector before the transport.
However, this will no longer be so on the two-sphere: starting on the
equator with a vector pointing north, we can shift it along a meridian to
the north pole, then back to the equator along a different meridian, and
finally back to its starting point on the equator. There, it will point into a
different direction than the original vector.
Curvature can thus be defined from this misalignment of vectors after
transport along closed curves. In order to work this out, we thus first
need some way for transporting vectors along curves.
31
32 3 Differential Geometry II
∇ f v y = f ∇v y
∇v ( f y) = f ∇v y + v( f )y , (3.1)
where f ∈ F is a C ∞ function on M.
where the n3 numbers Γki j are called the Christoffel symbols or connec-
tion coefficients of the connection ∇ in the given chart.
Connection
A connection ∇ generalises the directional derivative of objects on a
manifold. The directional derivative of a vector y in the direction of
the vector v is the vector ∇v y. The connection is linear and satisfies the
product rule.
The Christoffel symbols are not the components of a tensor, which is
seen from their transformation under coordinate changes. Let xi and xi
3.1 Connections and covariant derivatives 33
3.2 Geodesics
∇γ̇ v = 0 . (3.12)
and if this is to vanish identically, (3.12) and (3.14) imply the components
∇γ̇ γ̇ = 0 , (3.16)
∇γ̇ v = 0 (3.18)
holds. Geodesics are autoparallel curves,
∇γ̇ γ̇ = 0 . (3.19)
3.2 Geodesics 35
which maps any vector v from U ⊂ T p M into a point along the geodesic
through p into direction v at distance t = 1.
Now, we choose a coordinate basis {ei } of T p M and use the n basis
vectors in the exponential mapping (3.21). Then, the neighbourhood of
p can uniquely be represented by the exponential mapping along the
basis vectors, exp p (xi ei ), and the xi are called normal coordinates.
Since exp p (tv) = γtv (1) = γv (t), the curve γv (t) has the normal coordinates
xi = tvi , with v = vi ei . In these coordinates, xi is linear in t, thus ẍi = 0,
and (3.17) implies
Γki j vi v j = 0 , (3.22)
and thus
Γki j + Γk ji = 0 . (3.23)
t =v⊗w, (3.25)
∇ x (v ⊗ w) = ∇ x v ⊗ w + v ⊗ ∇ x w , (3.26)
C [∇ x (v ⊗ w)] = C (∇ x v ⊗ w) + C (v ⊗ ∇ x w)
= w(∇ x v) + (∇ x w)(v)
= ∇ x [w(v)] = xw(v) , (3.28)
where (3.1) was used in the final step (note that w(v) is a real-valued
function). Thus, we find an expression for the covariant derivative of a
dual vector,
(∇ x w)(v) = xw(v) − w(∇ x v) . (3.29)
(∇t)(w1 , . . . , wr , v1 , . . . , v s , v s+1 ) ≡
(∇vs+1 t)(w1 , . . . , wr , v1 , . . . , v s ) , (3.33)
where the vi are vector fields and the w j dual vector fields.
3.2 Geodesics 37
∇ x (w1 ⊗ . . . ⊗ wr ⊗ v1 ⊗ . . . ⊗ v s ⊗ t)
= (∇ x w1 ) ⊗ . . . ⊗ t + . . . w1 ⊗ . . . ⊗ (∇ x v1 ) ⊗ . . . ⊗ t
+ w1 ⊗ . . . ⊗ (∇ x t) , (3.34)
and then taking the total contraction, using that it commutes with the
covariant derivative, which yields
∇ x [t(w1 , . . . , wr , v1 , . . . , v s )]
= t(∇ x w1 , . . . , wr , v1 , . . . , v s ) + . . . + t(w1 , . . . , wr , v1 , . . . , ∇ x v s )
+ (∇ x t)(w1 , . . . , wr , v1 , . . . , v s ) . (3.35)
(∇ x t)(w1 , . . . , wr , v1 , . . . , v s )
= xt(w1 , . . . , wr , v1 , . . . , v s )
− t(∇ x w1 , . . . , v s ) − . . . − t(w1 , . . . , ∇ x v s ) . (3.36)
We now work out the last expression for the covariant derivative of a
tensor field in a local coordinate basis {∂i } and its dual basis {dx j } for the
special case of a tensor field t of rank (1, 1). The result for tensor fields
of higher rank are then easily found by induction.
We can write the tensor field t as
t = ti j (∂i ⊗ dx j ) , (3.37)
(∇ x t)(dxa , ∂b ) (3.39)
=x c
∂c tab − t j (∇ x dx )(∂i )dx (∂b ) − t j dx (∂i )dx (∇ x ∂b ) .
i a j i a j
showing that the covariant indices are transformed with the negative, the
contravariant indices with the positive Christoffel symbols.
In particular, the covariant derivatives of tensors of rank (0, 1) (dual
vectors w) and of tensors of rank (1, 0) (vectors v) have components
wi;k = wi,k − Γ jki w j ,
? vi;k = vi,k + Γik j v j . (3.43)
Convince yourself of the results
(3.42) and (3.43).
3.3 Curvature
Torsion
The torsion T maps two vector fields x and y into another vector field,
T : TM × TM → TM , (3.44)
such that
T (x, y) = ∇ x y − ∇y x − [x, y] . (3.45)
∇x y
∇x y − ∇ y x
x
∇y x
y
Figure 3.2 Torsion quantifies by how much parallelograms do not close.
with w ∈ T ∗ M and x, y ∈ T M is a tensor of rank (1, 2) called the torsion Caution In alternative, but
tensor . equivalent representations of gen-
eral relativity, the torsion does
According to (3.48), the components of the torsion tensor in the coordi- not vanish, but the curvature
nate basis {∂i } and its dual basis {dxi } are does. This is the teleparallel or
Einstein-Cartan version of gen-
T ki j = dxk T (∂i , ∂ j ) = Γki j − Γk ji . (3.49) eral relativity.
Curvature
The curvature R̄ maps three vector fields x, y and v into a vector field,
R̄ : T M × T M × T M → T M , (3.50)
such that
R̄(x, y)v = ∇ x (∇y v) − ∇y (∇ x v) − ∇[x,y] v . (3.51)
∇v x
R(u, v)x
∇u x
∇u ∇v x
∇ v ∇u x
x
v u
u v
R jl = R̄i jil = Γil j,i − Γii j,l + Γml j Γiim − Γmi j Γilm . (3.57)
The curvature and the torsion together satisfy the two Bianchi identities.
Bianchi identities
The first Bianchi identity is
R̄(x, y)z = {T [T (x, y), z] + (∇ x T )(y, z)} , (3.58)
cyclic cyclic
where the sums extend over all cyclic permutations of the vectors x, y
and z. The second Bianchi identity is
(∇ x R̄)(y, z) + R̄[T (x, y), z] = 0 . (3.59)
cyclic
They are important because they define symmetry relations of the curva-
ture and the curvature tensor. In particular, for a symmetric connection,
T = 0 and the Bianchi identities reduce to
R̄(x, y)z = 0 , (∇ x R̄)(y, z) = 0 . (3.60)
cyclic cyclic
3.3 Curvature 41
∇ x ∇y z − ∇y ∇ x z + ∇y ∇z x − ∇z ∇y x + ∇z ∇ x y − ∇ x ∇z y
− ∇[x,y] z − ∇[y,z] x − ∇[z,x] y
= ∇ x (∇y z − ∇z y) + ∇y (∇z x − ∇ x z) + ∇z (∇ x y − ∇y x)
− ∇[x,y] z − ∇[y,z] x − ∇[z,x] y
= ∇ x [y, z] − ∇[y,z] x + ∇y [z, x] − ∇[z,x] y + ∇z [x, y] − ∇[x,y] z
= [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0 , (3.67)
?
Convince yourself of the result where we have used the relation (3.10) and the Jacobi identity (2.33).
(3.67).
Up to now, the affine connection ∇ has not yet been uniquely defined.
We shall now see that a unique connection can be introduced on each
pseudo-Riemannian manifold (M, g).
A connection is called metric if the parallel transport along any smooth
curve γ in M leaves the inner product of two autoparallel vector fields x
and y unchanged. This is the case if and only if the covariant derivative
∇ of g vanishes,
∇g = 0 . (3.68)
Because of (3.36), this condition is equivalent to the Ricci identity
Caution In a third, equivalent xg(y, z) = g(∇ x y, z) + g(y, ∇ x z) , (3.69)
representation of general relativ-
ity, curvature and torsion both where x, y, z are vector fields.
vanish, but the metricity (3.68)
It can now be shown that a unique connection ∇ can be introduced on
is given up.
each pseudo-Riemannian manifold such that ∇ is symmetric or torsion-
free, and metric, i.e. ∇g = 0. Such a connection is called the Riemannian
or Levi-Civita connection.
Suppose first that such a connection exists, then (3.69) and the symmetry
of ∇ allow us to write
Taking the cyclic permutations of this equation, summing the second and
the third and subtracting the first (3.70), we obtain the Koszul formula
thus
1
gmk Γmi j = gik, j + g jk,i − gi j,k . (3.73)
2
If (gi j ) denotes the matrix inverse to (gi j ), we can write
1
Γli j = glk gik, j + g jk,i − gi j,k . (3.74)
2
Levi-Civita-connection
On a pseudo-Riemannian manifold (M, g) with metric g, a unique
connection exists which is symmetric and metric, ∇g = 0. It is called
Levi-Civita connection.
44 3 Differential Geometry II
The first of these relations is easily seen noting that the antisymmetry is
equivalent to
R̄(x, y)v, v = 0 . (3.76)
From the definition of R̄ and the antisymmetry (3.52), we first have
R̄i jkl = −R̄ jikl = −R̄i jlk , R̄i jkl = R̄kli j , (3.81)
where R̄i jkl ≡ gim R̄m jkl . Of the 44 = 256 components of the Riemann
tensor in four dimensions, the first symmetry relation (3.81) leaves
6 × 6 = 36 independent components, while the second symmetry relation
(3.81) reduces their number to 6 + 5 + 4 + 3 + 2 + 1 = 21.
In a coordinate basis, the Bianchi identities (3.60) for the curvature tensor
of a Riemannian connection read
R̄i jkl = 0 , R̄i jkl;m = 0 , (3.82)
( jkl) (klm)
by multiplying with δki and use the symmetry relations (3.81) to find
which is the contracted Bianchi identity. Moreover, the Ricci tensor can
easily be shown to be symmetric,
Ri j = R ji . (3.89)
R
Gi j ≡ Ri j − gi j , (3.90)
2
which has vanishing divergence because of the contracted Bianchi iden-
tity,
Gi j;i = 0 . (3.91)
46 3 Differential Geometry II
Ri j = R̄aia j . (3.92)
The Ricci scalar is the only contraction (the trace) of the Ricci tensor,
R = Tr R. The Ricci tensor, the Ricci scalar and the metric together
define the Einstein tensor, which has the components
R
Gi j = Ri j − gi j (3.93)
2
and is divergence-free, Gi j;i = 0.
Chapter 4
b b b
S = −mc2 dτ = −mc ds = −mc −ημν dxμ dxν , (4.1)
a a a
47
48 4 Physics in Gravitational Fields
To see what this equation implies, we now carry out the variation of S
and set it to zero,
b
δS = −mc δ −g(u, u) dτ = 0 . (4.5)
a
1
? ẍα + gαλ 2∂μ gλν − ∂λ gμν ẋμ ẋν = 0 . (4.11)
2
Convince yourself that (4.11)
is correct and agrees with the
geodesic equation. Comparing the result (4.11) to (3.17) and recalling the symmetry of
the Christoffel symbols (3.74), we arrive at the following important
conclusion:
Motion of freely falling particles
The trajectories extremising the action (4.4) are geodesic curves. Freely
falling particles thus follow the geodesics of the spacetime.
4.2 Motion of light 49
As an example for how physical laws can be carried from special to gen-
eral relativity, we now formulate the equations of classical electrodynam-
ics in a gravitational field. For a summary of classical electrodynamics,
see Appendix A.
In terms of the field tensor F, Maxwell’s equations read
∂μ Aμ = 0 , (4.14)
holds because of the antisymmetry of the field tensor F and the symmetry
of the connection ∇.
Indices have to be raised with the inverse metric g−1 now,
∇μ Aμ = 0 , (4.22)
but now the inhomogeneous wave equation (4.15) becomes more com-
plicated. We first note that
Fμν = ∇μ Aν − ∇ν Aμ ≡ ∂μ Aν − ∂ν Aμ (4.23)
λ L and λ R. (4.28)
k μ aμ = 0 , (4.31)
∇μ aμ + ikμ bμ = 0 . (4.32)
kν kν = 0 , (4.34)
Recall that the wave vector is the gradient of the scalar phase ψ. The
second covariant derivatives of ψ commute,
∇μ ∇ν ψ = ∇ν ∇μ ψ (4.37)
kν ∇ν kμ = 0 or ∇k k = 0 . (4.39)
1
kν ∇ν aμ + aμ ∇ν kν = 0 . (4.41)
2
1
kν (a∗μ ∇ν aμ + aμ ∇ν a∗μ ) = − ∇ν kν (a∗μ aμ + aμ a∗μ ) = −a2 ∇ν kν . (4.43)
2
Combining (4.43) with (4.42) yields
a
kν ∂ν a = − ∇ν kν , (4.44)
2
which shows how the amplitude is transported along light rays: the
change of the amplitude in the direction of the wave vector is proportional
?
to the negative divergence of the wave vector, which is a very intuitive
Why and in what sense is the re-
result.
sult (4.44) called intuitive here?
Finally, we obtain a law for the propagation of the polarisation. Using What does it mean?
aμ = aeμ in (4.41) gives
1
0 = kν ∇ν (aeμ ) + aeμ ∇ν kν
2) *
a
= akν ∇ν eμ + eμ kν ∂ν a + ∇ν kν = akν ∇ν eμ , (4.45)
2
where (4.44) was used in the last step. This shows that
k ν ∇ ν eμ = 0 or ∇k e = 0 , (4.46)
or in other words:
Transport of polarisation
The polarisation of electromagnetic waves is parallel-transported along
light rays.
54 4 Physics in Gravitational Fields
4.2.3 Redshift
From (3.74), we see that the contracted Christoffel symbol can be written
as
1
Γμμν = gμα gαν,μ + gμα,ν − gμν,α . (4.49)
2
Exchanging the arbitrary dummy indices α and μ and using the symmetry
of the metric, we can simplify this to
1
Γμμν = gμα gμα,ν . (4.50)
2
for any fixed i, where n is the dimension of the (square) matrix A. This
?
so-called Laplace expansion of the determinant follows after multiplying
Using (4.51) and (4.53), calculate
(4.52) with the matrix A jk .
the inverse and the determinant
By definition of the cofactors, any cofactor C ji does not contain the of 2 × 2 and 3 × 3 matrices.
element A ji of the matrix A. Therefore, we can use (4.52) and the
Laplace expansion (4.53) to conclude
∂ det A
= C ji = det A(A−1 )i j . (4.54)
∂A ji
The metric is represented by the matrix gμν , its inverse by gμν . We abbre-
viate its determinant by g here. Cramer’s rule (4.52) then implies that
the cofactors of gμν are C μν = g gμν , and we can immediately conclude
from (4.54) that
∂g
= ggμν (4.55)
∂gμν
and thus
∂g
∂λ g = ∂λ gμν = ggμν ∂λ gμν . (4.56)
∂gμν
1 √ 1 √
∇μ vμ = ∂μ vμ + √ vμ ∂μ −g = √ ∂μ ( −g vμ ) . (4.59)
−g −g
Again, by means of (4.57), we can combine the first and third terms on
the right-hand side to write
1 √
∇ν Aμν = √ ∂ν ( −gAμν ) + Γμαν Aαν . (4.61)
−g
Tensor divergences
If the tensor Aμν is antisymmetric, the second term on the right-hand
side of the divergence (4.61) vanishes because then the symmetric
Christoffel symbol Γμαν is contracted with the antisymmetric tensor Aαν .
If Aμν is symmetric, however, this final term remains, with important
consequences.
1 √ 1 √
∇μ ∇ν F μν = √ ∂μ ( −g∇ν F μν ) = √ ∂μ ∂ν ( −gF μν ) , (4.63)
−g −g
where we have used (4.62) in the final step. But the partial derivatives
commute, so that once more the antisymmetric tensor F μν is contracted
with the symmetric symbol ∂μ ∂ν . Thus, the result must vanish, allowing
us to conclude
∇μ ∇ν F μν = 0 . (4.64)
4π
∇μ ∇ν F μν = ∇μ jμ , (4.65)
c
which implies, by (4.59)
√
∂μ ( −g jμ ) = 0 . (4.66)
Charge conservation
Equation (4.66) is the continuity equation of the electric four-current,
implying charge conservation. We thus see that the antisymmetry of
the electromagnetic field tensor is necessary for charge conservation.
4.3 Energy-momentum (non-)conservation 57
∂ν T μν = 0 . (4.67)
and for μ = 0, the vanishing divergence (4.67) yields the energy con-
servation equation
⎛ ⎞ + ,
2 ⎟⎟
∂ ⎜⎜⎜ E 2 + B · c (E × B)
⎟⎟⎠ + ∇ =0,
⎜⎝ (4.69)
∂t 8π 4π
According to our general rule for moving results from special relativity
to general relativity, we can replace the partial derivative in (4.67) by the
covariant derivative,
∇ν T μν = 0 , (4.71)
and obtain an equation which is covariant and thus valid in all reference
frames. Moreover, we would have to replace the Minkowski metric η in
(4.68) by the metric g if we wanted to consider the energy-momentum
tensor of the electromagnetic field.
From our general result (4.61), we know that we can rephrase (4.71) as
1 √ μ
√ ∂ν ( −gT μν ) + Γ λν T λν = 0 . (4.72)
−g
If the second term on the left-hand side was absent, this equation would
imply a conservation law. It remains there, however, because the energy-
momentum tensor is symmetric. In presence of this term, we cannot
convert (4.72) to a conservation law any more. This result expresses the
following important fact:
58 4 Physics in Gravitational Fields
Energy non-conservation
Energy is not generally conserved in general relativity. This is not
surprising because energy can now be exchanged with the gravitational
field.
Finally, we want to see under which conditions for the metric the Newto-
nian limit for the equation of motion in a gravitational field is reproduced,
which is
x¨ = −∇Φ (4.73)
to very high precision in the Solar System.
We first restrict the gravitational field to be weak and to vary slowly with
time. This implies that the Minkowski metric of flat space is perturbed
by a small amount,
gμν = ημν + hμν , (4.74)
with |hμν | 1.
Moreover, we restrict the consideration to bodies moving much slower
than the speed of light, such that
? dxi dx0
Does it matter with respect to ≈1. (4.75)
dτ dτ
which coordinate frame the ve-
locity is assumed to be much less Under these conditions, the geodesic equation for the i-th spatial coordi-
than the speed of light? nate reduces to
d2 x i d2 xi dxα dxβ
≈ = −Γi
αβ ≈ −Γi00 . (4.76)
c2 dt2 dτ2 dτ dτ
The constant can be set to zero because both the deviation from the
Minkowski metric and the gravitational potential vanish at large distance
from the source of gravity. Therefore, the metric in the Newtonian limit
has the 0-0 element
2Φ
g00 ≈ −1 − 2 . (4.80)
c
∇k k = 0 or kν ∂ν kμ + Γμνλ kν kλ = 0 , (4.82)
Assuming that the gravitational potential Φ does not vary with time,
∂0 Φ = 0, the only non-vanishing Christoffel symbols of the metric (4.81)
are
1
Γ00i ≈ 2 ∂i Φ ≈ Γi00 . (4.84)
c
For μ = 0, (4.82) yields
1 ω + ω e · ∇Φ = 0 ,
∂t + e · ∇ (4.85)
c c2
which shows that the frequency changes with time only because the light
path can run through a spatially varying gravitational potential. Thus, if
the potential is constant in time, the frequencies of the incoming and the
outgoing light must equal.
Using this result, the spatial components of (4.82) read
1 e = de = − 1 ∇
∂t + e · ∇ Φ = − ∇⊥ Φ ;
− e(e · ∇) (4.86)
c cdt c 2 c2
60 4 Physics in Gravitational Fields
in other words, the total time derivative of the unit vector in the direc-
tion of the light ray equals the negative perpendicular gradient of the
gravitational potential.
For calculating the light deflection, we need to know the total change
of e as the light ray passes the Sun. This is obtained by integrating
(4.86) along the actual (curved) light path, which is quite complicated.
Caution Note that this approx- However, due to the weakness of the gravitational field, the deflection
imation is conceptually identi- will be very small, and we can evaluate the integral along the unperturbed
cal to Born’s approximation in (straight) light path.
quantum-mechanical scattering
We choose a coordinate system centred on the Sun and rotated such that
problems.
the light ray propagates parallel to the z axis from −∞ to ∞ at an impact
parameter b. Outside the Sun, its gravitational potential is
Φ GM
GM
=− 2 =− √ . (4.87)
c 2 cr c2 b2 + z2
⊥ Φ = ∂Φ eb =
∇
GM b
eb , (4.88)
∂b c2 (b2 + z2 )3/2
where eb is the radial unit vector in the x-y plane from the Sun to the
light ray.
Light deflection in (incomplete) Newtonian approximation
Thus, under the present assumptions, the deflection angle is
∞
GMb 2GM
δe = −eb dz 2 2 = − 2 eb . (4.89)
−∞ c (b + z )
2 3/2 cb
γt : M → M (5.3)
63
64 5 Differential Geometry III
Pull-back
Let now M and N be two manifolds and φ : M → N a map from M
onto N. A function f defined at a point q ∈ N can be defined at a point
p ∈ M with q = φ(p) by
φ∗ : T p M → T q N , v → φ∗ v = v ◦ φ∗ , (5.6)
such that (φ∗ v)( f ) = v(φ∗ f ) = v( f ◦ φ). This defines a vector from the
tangent space of N in q = φ(p).
Push-forward
The map φ∗ “pushes” vectors from the tangent space of M in p to the
tangent space of N in q and is thus called the push-forward.
In a natural generalisation to dual vectors, we define their pull-back φ∗
by
φ∗ : T q∗ N → T p∗ M , w → φ∗ w = w ◦ φ∗ , (5.7)
φ∗ : T sr (N) → T sr (M) ,
φ∗ : T sr (M) → T sr (N) . (5.11)
Using the latter expression, we can derive coordinate expressions for the
Lie derivative. We introduce the coordinate basis {∂i } and its dual basis
{dxi } and apply (5.26) to dxi ,
Lv dxi = dLv xi = dv(xi ) = dv j ∂ j xi = dvi = ∂ j vi dx j . (5.27)
The Lie derivative of the basis vectors ∂i is
Lv ∂i = [v, ∂i ] = −(∂i v j )∂ j , (5.28)
where (2.32) was used in the second step.
? where we have identified the Christoffel symbols (3.74) in the last step.
Derive the Killing equation (5.34) This is the Killing equation.
yourself.
Let γ be a geodesic, i.e. a curve satisfying
∇γ̇ γ̇ = 0 , (5.35)
changing indices and using the symmetry of the metric, we can also
write it as
γ̇, ∇γ̇ K = g jk γ̇ j γ̇i ∇i K k = γ̇ j γ̇i ∇i K j . (5.39)
Adding the latter two equations and using the Killing equation (5.34)
shows
2γ̇, ∇γ̇ K = γ̇i γ̇ j ∇i K j + ∇ j Ki = 0 , (5.40)
which proves (5.36). More elegantly, we have contracted the symmetric
tensor γ̇i γ̇ j with the tensor ∇i K j which is antisymmetric because of the
Killing equation, thus the result must vanish.
Equation (5.36) has a profound meaning:
5.3 Differential forms 69
5.3.1 Definition
Differential p-forms are totally antisymmetric tensors of rank (0, p). The
most simple example are dual vectors w ∈ T p∗ M since they are tensors
of rank (0, 1). A general tensor t of rank (0, 2) is not antisymmetric, but
can be antisymmetrised defining the two-form
1
τ(v1 , v2 ) ≡ [t(v1 , v2 ) − t(v2 , v1 )] , (5.41)
2
with two vectors v1 , v2 ∈ V.
To generalise this operation for tensors of arbitrary ranks (0, r), we first
define the alternation operator by
1
(At)(v1 , . . . , vr ) := sgn(π)t(vπ(1) , . . . , vπ(r) ) , (5.42)
r! π
where the sum extends over all permutations π of the integer numbers
?
from 1 to r. The sign of a permutation, sgn(π), is negative if the permu-
As an exercise, explicitly apply
tation is odd and positive otherwise.
the alternation operator to a ten-
In components, we briefly write sor field of rank (0, 3).
1
ωi j = ω[i j] = ωi j − ω ji . (5.45)
2
-
The vector space of p-forms is denoted by p . Taking the product of two
-p -q
differential forms ω ∈ and η ∈ yields a tensor of rank (0, p + q)
70 5 Differential Geometry III
(p + q)!
ω∧η≡ A(ω ⊗ η) (5.46)
p!q!
into a (p + q)-form.
The definition of the exterior product implies that it is bilinear, associa-
tive, and satisfies
ω ∧ η = (−1) pq η ∧ ω . (5.47)
-
A basis for the vector space p can be constructed from the basis {dxi },
1 ≤ i ≤ n, of the dual space V ∗ by taking
Given two vector spaces V and W above the same field F, the Cartesian
product V × W of the two spaces can be turned into a vector space by
defining the vector-space operations component-wise. Let v, v1 , v2 ∈ V
and w, w1 , w2 ∈ W, then the operations
-
(i) d is an antiderivation of degree 1 on , i.e. it satisfies
d (ω ∧ η) = dω ∧ η + (−1) p ω ∧ dη (5.58)
-p -
for ω ∈ and η ∈ .
(ii) d ◦ d = 0.
Since ωi2 ...i p+1 is itself antisymmetric, this last expression can be brought
into the form
p+1
(dω)i1 ...i p+1 = (−1)k+1 ∂ik ωi1 ,...,îk ,...i p+1 , (5.62)
k=1
with 1 ≤ i1 < . . . < i p < i p+1 ≤ n. Indices marked with a hat are left out.
The Lie derivative, the interior product and the exterior derivative are
related by Cartan’s equation
? Lv = d ◦ iv + iv ◦ d . (5.63)
Verify the expressions (5.61) and
(5.62).
Cartan’s equation implies the convenient formula for the exterior deriva-
tive of a p-form ω
p+1
dω(v1 , . . . , v p+1 ) = (−1)i+1 vi ω(v1 , . . . , v̂i , . . . , v p+1 ) (5.64)
i=1
where the hat over a symbol means that this object is to be left out.
dωi j = ∂i ω j − ∂ j ωi (5.67)
5.4 Integration
Volume form
An n-dimensional, paracompact manifold M is orientable if and only
if a C ∞ , n-form exists on M which vanishes nowhere. This is called a
volume form.
The canonical volume form on a pseudo-Riemannian manifold (M, g)
is defined by
η ≡ |g| dx1 ∧ . . . ∧ dxn . (5.69)
This definition is independent of the coordinate system because it
transforms proportional to the Jacobian determinant upon coordinate
changes.
Equation (5.69) implies that the components of the canonical volume
form in n dimensions are proportional to the n-dimensional Levi-Civita
symbol,
ηi1 ...in = |g| εi1 ...in , (5.70)
which is defined such that it is +1 for even permutations of the i1 , . . . , in ,
−1 for odd permutations, and vanishes if any two of its indices are equal.
A very useful relation is
∗dx1 = dx2 ∧ dx3 , ∗dx2 = dx3 ∧ dx1 , ∗dx3 = dx1 ∧ dx2 (5.78)
Codifferential
The codifferential is a map
.p . p−1
δ: → , ω → δω (5.81)
defined by
δω ≡ sgn(g)(−1)n(p+1) (∗d∗)ω . (5.82)
d ◦ d = 0 immediately implies δ ◦ δ = 0.
By successive application of (5.71) and (5.62), it can be shown that the
coordinate expression for the codifferential is
1
(δω)i1 ...i p−1 =
∂k |g|ωki1 ...i p−1 . (5.83)
|g|
Comparing this with (4.59), we see that this generalises the divergence
of ω. To see this more explicitly, let us work out the codifferential of a
1-form in R3 by first taking the exterior derivative of ∗ω from (5.79),
d∗ω = (∂1 ω1 + ∂2 ω2 + ∂3 ω3 ) dx1 ∧ dx2 ∧ dx3 , (5.84)
5.4 Integration 75
δω = ∂1 ω1 + ∂2 ω2 + ∂3 ω3 . (5.85)
dF = 0 . (5.88)
Integral theorems
Stokes’ theorem can now be formulated as follows: let M be an n-
Caution Like lowers a note dimensional manifold and the region D ⊂ M have a smooth boundary
by a semitone in music, the op- ∂D such that D̄ ≡ D ∪ ∂D is compact. Then, for every n − 1-form ω,
erator lowers the index of vec- we have
tor components and thus turns dω = ω. (5.94)
them into dual-vector compo- D ∂D
nents. Analogously, raises Likewise, Gauss’ theorem can be brought into the form
notes by semitones in music, and
indices of dual-vector compo- δx η = ∗x , (5.95)
nents. D ∂D
Musical operators
Generally, the musical operators and are isomorphisms between the
tangent spaces of a manifold and their dual spaces given by the metric,
77
78 6 Einstein’s Field Equations
v = γ̇ = ∂λ . (6.3)
Since the partial derivatives with respect to the curve parameters τ and
λ commute, so do the vectors u and v, and thus v is Lie-transported (or
Lie-invariant) along u,
0 = [u, v] = Lu v . (6.4)
γ
v
Figure 6.1 Geodesic bundle with tangent vector u of the fiducial geodesic,
the curve γ towards a neighbouring geodesic, and tangent vector v = γ̇.
which does indeed satisfy n, u = 0 because of u, u = −1. This vector
is also Lie-transported along u, as we shall verify now.
6.1 The physical meaning of curvature 79
First, we have
where (6.4) was used in the first step. Since u, u = −1, we have
where the Ricci identity was used in the second step, the geodesic prop-
erty (6.2) in the third, and (6.7) in the last. Returning to (6.6), this proves
that n is Lie-transported,
Lu n = 0 . (6.9)
The perpendicular separation vector between neighbouring geodesics of
the congruence is thus Lie-invariant along the congruence.
where we have used again that u and v commute and that u is a geodesic.
With [u, v] = 0, the curvature (3.51) applied to u and v reads
Jacobi equation
In this way, we see that the second derivative of v along u is determined
by the curvature tensor through the Jacobi equation
We then use
n = ni ei (6.17)
and thus
dni
∇u n = (uni )ei + (ni ∇u )ei = ei . (6.18)
dτ
d2 ni
ei = R̄(e0 , n j e j )e0 = n j R̄(e0 , e j )e0 = n j R̄i00 j ei . (6.19)
dτ2
Thus, defining a matrix K by
d2 ni
= R̄i00 j n j ≡ K ij n j , (6.20)
dτ2
we can write (6.19) in matrix form
d2n
= Kn . (6.21)
dτ2
Note that K is symmetric because of the symmetries (3.81) of the curva-
ture tensor.
Moreover, the trace of K is
where we have inserted R̄0000 = 0 and the definition of the Ricci tensor
(3.57).
6.1 The physical meaning of curvature 81
Let us now compare this result to the motion of test bodies in Newtonian
theory. At two neighbouring points x and x + n, we have the equations
of motion
ẍi = − (∂i Φ)|x (6.23)
and, to first order in a Taylor expansion,
ẍi + n̈i = − (∂i Φ)|x+n ≈ − (∂i Φ)|x − (∂i ∂ j Φ)x n j . (6.24)
Subtracting (6.23) from (6.24) yields the evolution equation for the
separation vector.
Relative acceleration in Newtonian gravity
In Newtonian gravity, the separation vector between any two particle
trajectories changes due to the tidal field according to
∂i ∂ j Φ
j = −
Ki(N) , (6.27)
c2
and its trace is
2Φ
∇ ΔΦ
Tr K (N) = − 2
=− 2 , (6.28)
c c
i.e. the negative Laplacian of the Newtonian potential, scaled by the
squared light speed.
Tidal field and curvature
The essential results of this discussion are the correspondences
∂i ∂ j Φ
R̄i0 j0 ↔ (6.29)
c2
and
2Φ
∇
Rμν uμ uν ↔ . (6.30)
c2
These confirm the assertion that the curvature represents the gravita-
tional tidal field, describing the relative accelerations of freely-falling
test bodies; (6.29) and (6.30) will provide useful guidance in guessing
the field equations.
82 6 Einstein’s Field Equations
We start from the field equation from Newtonian gravity, i.e. the Poisson
equation
4πGρ = ∇ 2 Φ = −c2 Tr K (N) . (6.31)
and insert this into the field equation (6.31), using (6.22) for the trace of
K. We thus obtain
4πG
R(u, u) = 4
λT + (1 − λ) Tr T g (u, u) . (6.36)
c
Since this equation should hold for any observer and thus for arbitrary
four-velocities u, we find
4πG
R= 4
λT + (1 − λ) Tr T g , (6.37)
c
where λ ∈ R remains to be determined.
We take the trace of (6.37), obtain
4πG
Tr R = R = [λ + 4(1 − λ)] Tr T (6.38)
c4
and combine this with (6.37) to assemble the Einstein tensor (3.90),
R
G =R− g
2
4πG 2−λ
= 4 λT − Tr T g . (6.39)
c 2
6.2 Einstein’s field equations 83
We have seen in (3.91) that the Einstein tensor G satisfies the contracted
Bianchi identity
∇·G = 0 . (6.40)
6.2.2 Uniqueness
D[g] = T , (6.44)
Lovelock’s theorem
If D[g] depends on g and its derivatives only up to second order, then
it must be a linear combination of the Einstein and metric tensors,
D[g] = αG + βg , (6.45)
?
Express the coefficients α and β with α, β ∈ R. This absolutely remarkable theorem says that G must be
in terms of κ and Λ. of the form
G = κT − Λg , (6.46)
with κ and Λ are constants. The correct Newtonian limit then requires
that κ = 8πGc−4 , and Λ is the “cosmological constant” introduced by
Einstein for reasons which will become clear later.
Working out the variation of this action with respect to the metric com-
ponents gμν , we write explicitly
R = gμν Rμν (6.49)
6.3 Lagrangian formulation 85
and thus
√
δS GR = δ gμν Rμν −g d4 x (6.50)
D √
√
= δRμν gμν −g d4 x + Rμν δ gμν −g d4 x .
D D
We evaluate the variation of the Ricci tensor first, using its expression
(3.57) in terms of the Christoffel symbols. Matters simplify considerably
if we introduce normal coordinates, which allow us to ignore the terms
in (3.57) which are quadratic in the Christoffel symbols. Then, the Ricci
tensor specialises to
which is a tensor identity, called the Palatini identity, and thus holds in
all coordinate systems everywhere. It implies
gμν δRμν = ∇α gμν δΓαμν − gμα δΓνμν , (6.54)
where the indices α and ν were swapped in the last term. Thus, the
variation of the Ricci tensor, contracted with the metric, can be expressed
by the divergence of a vector W,
√
Using these expressions, we obtain for the variation of −g
√
√ δg ggμν δgμν −g
δ −g = − √ = √ =− gμν δgμν , (6.58)
2 −g 2 −g 2
86 6 Einstein’s Field Equations
1 1
δη = − gμν δgμν η = gμν δgμν η . (6.59)
2 2
Now, we put (6.58) and (6.55) back into (6.50) and obtain
α R
δS GR = ∇α W η + Rμν − gμν δgμν η
2
D
D
Gμν δgμν η + ∇α W α η = 0 .
!
= (6.60)
D D
Gμν = 0 . (6.61)
where the Lagrangian is varied with respect to the fields ψ and their
derivatives ∇ψ. Thus,
∂L ∂L
δ Lη = δψ + δ∇ψ η = 0 . (6.65)
D D ∂ψ ∂∇ψ
1 m2 2 1 m2 2
L = − ∇ψ, ∇ψ − ψ = − ∇μ ψ∇μ ψ − ψ , (6.68)
2 2 2 2
where mψ2 /2 is a mass term with constant parameter m. The Euler-
Lagrange equations then imply the field equations
− + m2 ψ = −∇μ ∇μ + m2 ψ = 0 , (6.69)
Similarly, we can vary the action with respect to the metric, which
requires care because the Lagrangian may depend on the metric explicitly
and implicitly through the covariant derivatives ∇ψ of the fields, and
the canonical volume form η depends on the metric as well because of
(6.59). Thus,
1
δ Lη = (δL)η + Lδη = δL − gμν Lδgμν η . (6.70)
D D D 2
If the Lagrangian does not implicitly depend on the metric, we can write
∂L μν
δL = δg . (6.71)
∂gμν
88 6 Einstein’s Field Equations
1
δΓαμν = gασ ∇ν δgσμ + ∇μ δgσν − ∇σ δgμν , (6.72)
2
Λ Λ Λc4
T μν → T μν + T μν , T μν ≡− gμν . (6.81)
8πG
T = ρc2 u ⊗ u . (6.82)
This expression just says that, if the domain D is mapped along the flow
?
lines of the fluid flow, it will encompass a constant amount of material
Convince yourself recalling the
independent of time t.
definition of the pull-back that
Now, we can use the pull-back to write (6.84) is correct.
ρη = φ∗t (ρη) , (6.84)
φt (D) D
and take the limit t → 0 to see the equivalence of (6.83) and (6.84) with
the vanishing Lie derivative of ρη along u,
Lu (ρη) = 0 . (6.85)
90 6 Einstein’s Field Equations
(Lu ρ)η = (uρ)η = (ui ∂i ρ)η = (ui ∇i ρ)η = (∇u ρ)η . (6.87)
For the second term, we can apply equation (5.30) for the components
of the Lie derivative of a rank-(0, 4) tensor, and use the antisymmetry of
η to see that
Lu η = (∇μ uμ )η = (∇ · u)η . (6.88)
Accordingly, (6.86) can be written as
? or
Give a physical interpretation of ∇μ (ρuμ ) = 0 . (6.90)
equation (6.90). What does it
mean? At the same time, the divergence of T must vanish, hence
The first term vanishes because of (6.90), and the second implies
uν ∇ ν u μ = 0 ⇔ ∇u u = 0 . (6.92)
In other words, the flow lines have to be geodesics. For an ideal fluid, the
equation of motion thus follows directly from the vanishing divergence
of the energy-momentum tensor, which is required in general relativity
by the contracted Bianchi identity (3.91).
Chapter 7
We begin our study of solutions for the field equations with situations in
which the metric is almost Minkowskian, writing
gμν = ημν + hμν , (7.1)
where hμν is considered as a perturbation of the Minkowski metric ημν
such that
|hμν | 1 . (7.2)
This condition is excellently satisfied e.g. in the Solar System, where
Φ ?
|hμν | ≈ ≈ 10−6 . (7.3) How can you most easily confirm
c2
the estimate (7.3) for the Solar
Note that small perturbations of the metric do not necessarily imply System and other astronomical
small perturbations of the matter density, as the Solar System illustrates. objects?
Also, the metric perturbations may change rapidly in time.
First, we write down the Christoffel symbols for this kind of metric.
Starting from (3.74) and ignoring quadratic terms in hμν , we can write
1
Γαμν = ηαβ ∂ν hμβ + ∂μ hβν − ∂β hμν
2
1 α
= ∂ν hμ + ∂μ hαν − ∂α hμν . (7.4)
2
Next, we can ignore the terms quadratic in the Christoffel symbols in the
components of the Ricci tensor (3.56) and find
Rμν = ∂λ Γλμν − ∂ν Γλλμ . (7.5)
91
92 7 Weak Gravitational Fields
1
Rμν = ∂λ ∂ν hλμ + ∂λ ∂μ hλν − ∂λ ∂λ hμν − ∂μ ∂ν hλλ
2
1
= ∂λ ∂ν hλμ + ∂λ ∂μ hλν − hμν − ∂μ ∂ν h , (7.6)
2
where we have introduced the d’Alembert operator and abbreviated the
trace of the metric perturbation,
Caution Note that the
d’Alembert operator is the
= ∂λ ∂λ , h ≡ hλλ . (7.7)
square of the ordinary partial
derivative here, not the covariant
derivative. Is this appropriate, The Ricci scalar is the contraction of Rμν ,
and if so, why?
R = ∂λ ∂μ hλμ − h , (7.8)
∂ν T μν = 0 . (7.11)
One could now insert the Minkowski metric in T μν , search for a first
μν of the linearised field equations and iterate replacing ημν
solution h(0)
μν
by ημν + h(0)
μν in T to find a corrected solution h(1)
μν , and so forth. This
procedure is useful as long as the back-reaction of the metric on the
energy-momentum tensor is small.
If we specialise (7.11) for pressure-less dust and insert (6.82), we find
the equation of motion
? uν ∂ν uμ = 0 , (7.12)
Compare (7.12) to the geodesic
which means that the fluid elements follow straight lines.
equation for the motion of fluid
particles.
7.1.2 Wave equation for metric fluctuations
1
γμν ≡ hμν − ημν h (7.13)
2
7.2 Gauge transformations 93
for hμν . Since γ ≡ γμμ = −h, we can solve (7.13) for hμν and insert
1
hμν = γμν − ημν γ (7.14)
2
into (7.9) to obtain the linearised field equations
16πG
∂λ ∂ν γλμ + ∂λ ∂μ γλν − ημν ∂λ ∂σ γλσ − γμν = T μν . (7.15)
c4
Diffeomorphism invariance
Let φ be a diffeomorphism of M, such that φ : M → N in diffeomor- ?
phic way. Since φ is then bijective and smoothly differentiable and In particular, the diffeomorphism
has a smoothly differentiable inverse, M and N can be considered as invariance of general relativity
indistinguishable abstract manifolds. The manifolds M and N then implies that coordinate systems
represent the same physical spacetime. In particular, the metric g on can have no physical significance.
M is then physically equivalent to the pulled-back metric φ∗ g. This dif- Use your own words to explain
feomorphism invariance is a fundamental property of general relativity. why this is so.
Now, set g = η + h and define the infinitesimal vector ξ ≡ tv. Then, the
transformation (7.16) implies
h → φ∗ h = h + tLv η + tLv h = h + Lξ η + Lξ h . (7.17)
For weak fields, the third term on the right-hand side can be neglected.
Using (5.31), we see that
(Lξ η)μν = ηλν ∂μ ξλ + ημλ ∂ν ξλ = ∂μ ξν + ∂ν ξμ . (7.18)
We thus find the following important result:
Gauge transformations of weak metric perturbations
The weak metric perturbation hμν admits the gauge transformation
∂ν γμν = 0 . (7.21)
such that, if (7.21) is not satisfied yet, it can be achieved by choosing for
ξμ a solution of the inhomogeneous wave equation
1
G(x, x ) = δD x0 − x 0 − |x − x | . (7.26)
|x − x |
T x0 − |x − x |, x
4G μν
γμν (x) = 4 d3 x (7.27)
c |x − x |
⊥ ln n = − 2 ∇
e˙ = ∇ ⊥Φ (7.43)
c2
7.3 Nearly Newtonian gravity 97
Shapiro delay
Compared to light propagation in vacuum, there is thus a time delay
?
dl 2Φ How could the Shapiro delay be
Δ(dt) = dt − = − 3 dl , (7.45) measured?
c c
which is called the Shapiro delay.
Gravitomagnetic potential
Matter currents create a magnetic gravitational potential similar to the
electromagnetic vector potential.
The most direct approach to the equations of motion starts from the
variational principle (4.5), or
1/2
δ −gμν ẋμ ẋν dt = 0 , (7.50)
where the dot now denotes the derivative with respect to the coordinate
time t. The radicand is
· v − v 2 ,
c2 − 2c2 A0 − 8cA (7.51)
where we have neglected terms of order Φv 2 since the velocities are
assumed to be small compared to the speed of light.
Using (7.51), we can reduce the least-action principle (7.51) to the
Euler-Lagrange equations with the effective Lagrangian
v 2 · v .
L= + A0 c2 + 4cA (7.52)
2
We first find
∂L d ∂L dv A.
= v + 4cA ⇒ = + 4c(v · ∇) (7.53)
∂v dt ∂v dt
?
Convince yourself of the vector Using the vector identity
identity (7.54).
a · b) = (a · ∇)
∇( b + (b · ∇)
a + a × ∇
× b + b × ∇
× a , (7.54)
we further obtain
∂L
0 + 4c (v · ∇)
= c2 ∇A A × A)
+ v × (∇ , (7.55)
∂x
7.3 Nearly Newtonian gravity 99
dv 0 + 4cv × (∇
× A)
,
≡ f = c2 ∇A (7.56)
dt
= Be3 , i.e. B1 = 0 =
Let us now orient the coordinate frame such that B
B2 . Then,
ṡ1 = 2cBs2 , ṡ2 = −2cs1 B . (7.60)
σ̇ = −2cBiσ , (7.61)
= −2c B
ω ×A
= −2c∇ , (7.62)
γμν = 0 , (7.69)
∂ν γμν = 0 . (7.70)
Within the Hilbert gauge class, we can further require that the trace of
γμν vanish,
γ = γμμ = 0 . (7.71)
To see this, we return to the gauge transformation (7.20), which implies
2∂μ ξμ = γ . (7.73)
Moreover, (7.22) shows that the Hilbert gauge condition remains pre-
served if ξμ satisfies the d’Alembert equation
ξμ = 0 (7.74)
at the same time. It can be generally shown that vector fields ξμ can be
?
constructed which indeed satisfy (7.74) and (7.74) at the same time. If
How would you construct a solu-
we arrange things in this way, (7.14) shows that then hμν = γμν .
tion to both equations (7.73) and
(7.73)? All functions propagating with the light speed satisfy the d’Alembert
equation (7.69). In particular, we can describe them as superpositions of
plane waves
γμν = hμν = Re εμν eik,x (7.75)
7.4 Gravitational waves 101
as used in (7.58).
102 7 Weak Gravitational Fields
k2 = k, k = kμ kμ = 0 . (7.76)
εμμ = 0 . (7.78)
The five conditions (7.78) and (7.78) imposed on the originally ten inde-
pendent components of εμν leave five independent components. Without
loss of generality, suppose the wave propagates into the positive z direc-
tion, then
kμ = (k, 0, 0, k) , (7.79)
and (7.77) implies
ε0μ = ε3μ ; (7.80)
specifically,
ε00 = ε30 = ε03 = ε33 and ε01 = ε31 , ε02 = ε32 , (7.81)
∂μ ξμ = 0 , (7.86)
1 00
εμ = −ε , −2ε01 − 2ε02 , −ε00 (7.89)
2k
fixes the gauge transformation, and the polarisation tensor is reduced to
⎛ ⎞
⎜⎜⎜ 0 0 0 0 ⎟⎟⎟
⎜⎜⎜ 0 ε11 ε12 ⎟⎟⎟
ε = ⎜⎜⎜⎜
μν ⎜ 0 ⎟⎟⎟ . (7.90)
⎜⎜⎝ 0 ε12 −ε11 0 ⎟⎟⎟
⎟⎠
0 0 0 0 ?
Carry out all steps yourself lead-
ing from (7.77) to (7.90).
Gauge-invariant polarisation states
As for electromagnetic waves, there are only two gauge-invariant, lin- ?
early independent polarisation states for gravitational waves. How are polarisation states of
electromagnetic waves being de-
Under rotations about the z axis by an arbitrary angle φ, the polarisation scribed? Recall the Stokes pa-
tensor changes according to rameters and their meaning.
which shows that the two polarisation states ε± have helicity ±2, and
thus that they correspond to left and right-handed circular polarisation.
104 7 Weak Gravitational Fields
because “far away” means that the source is small compared to its
distance. Moreover, we can approximate the retarded time coordinate x0
as follows:
√
x0 − |x − x | = x0 − (x − x )2 = x0 − x 2 + x 2 − 2x · x
≈ x0 − r + x · er , (7.96)
From Gauß’ theorem, we infer that the volume integral over the diver-
gence of a vector field equals the integral of the vector field over the
boundary of the volume and must vanish if the field disappears on the
surface,
∂ j T j0 xl xk d3 x = 0 . (7.101)
7.4 Gravitational waves 105
?
Does the expression (7.109) have showing one of the rare cases of a third time derivative in physics.
the correct units?
The transversal quadrupole tensor is
Q22 Q23
QT = (7.110)
Q32 Q33
M ≈ 30 M
. (7.115)
less than a thousandth of the Solar radius. No ordinary stars could ever
come as close. Objects of mass m closer than R must be black holes.
The merging black-hole binary became known as GW150914.
Inserting the quadrupole tensor (7.122) into (7.105) leads to
jk GM Gμ
γ ≤ 4 (7.118)
rc2 Rc2
108 7 Weak Gravitational Fields
Ṙ 2 ω̇
=− (7.126)
R 3ω
and thus
1 GμM ω̇
|Epot | = . (7.127)
3 R ω
Equating this to (7.124), using (7.121) to eliminate the radius via the
angular frequency, taking into account that the frequency f of the
gravitational waves emitted by the binary is f = ω/π, and sorting terms
leads to the chirp mass
3/5 3/5
c3 5 f˙
M := M 2/3
μ = . (7.128)
8G 3π8/3 f 11/3
111
112 8 The Schwarzschild Solution
-
Let now α ∈ 1 be a one-form such that α = αi θi with arbitrary functions
αi . Then, the equations we have derived so far imply
or
? ∇x = ei ⊗ (dxi + ωik xk ) (8.10)
Derive the expressions (8.9) and
for the covariant derivative of the vector x.
(8.10) yourself, beginning with
(8.1).
8.1.2 Torsion and curvature forms
We are now in a position to use the connection forms for defining the
torsion and curvature forms.
Torsion and curvature forms
By definition, the torsion T (x, y) is a vector, which can be written in
terms of the torsion forms Θi as
The next important step is now to realise that the torsion and curvature
2-forms satisfy Cartan’s structure equations:
8.1 Cartan’s structure equations 113
Θi = dθi + ωij ∧ θ j
Ωij = dωij + ωik ∧ ωkj . (8.13)
Θi (x, y) = ∇ x y − ∇y x − [x, y]
= ∇ x (θi (y)ei ) − ∇y (θi (x)ei ) − θi ([x, y])ei , (8.14)
where we have expanded the vectors x, y and [x, y] in the basis {ei }
according to x = θi , xei = θi (x)ei . Then, we continue by using the
connection forms,
1
Θi ≡ T ijk θ j ∧ θk (8.21)
2
are the elements of the torsion tensor in the basis {ei }, since
ωi j + ω ji = dgi j , (8.23)
were used, i.e. the gi j are the components of the metric in the arbitrary
basis {ei }.
ω ∧ dω = 0 (8.28)
− φ2 = K, K . (8.33)
4πr2 = A (8.34)
The full metric g is thus characterised by two radial functions a(r) and
b(r) which we need to determine. The exponential functions in (8.35)
and (8.36) are chosen to ensure that the prefactors ea and eb are always
positive.
The spatial sections Σ are now foliated according to
Σ = I × S2 , I ⊂ R+ , (8.37)
The functions a(r) and b(r) are constrained by the requirement that the
metric should asymptotically turn flat, which means
They must be determined by inserting the metric (8.38) into the vacuum
field equations, G = 0.
Given the dual tetrad {θμ }, we must take their exterior derivatives. For
this purpose, we apply the expression (??) and find, for dθ0 ,
dθ1 = 0 (8.44)
8.3 The Schwarzschild solution 119
because dr ∧ dr = 0, further
dθ2 = dr ∧ dϑ (8.45)
and
dθ3 = sin ϑ dr ∧ dϕ + r cos ϑ dϑ ∧ dϕ . (8.46)
Using (8.40), we can also express the coordinate differentials by the dual
tetrad,
θ2 θ3
cdt = e−a θ0 , dr = e−b θ1 , , dϕ =
dϑ = , (8.47)
r r sin ϑ
so that we can write the exterior derivatives of the dual tetrad as
e−b 1
dθ0 = a e−b θ1 ∧ θ0 , dθ1 = 0 , dθ2 = θ ∧ θ2 ,
r
e−b 1 cot ϑ 2 ?
dθ3 = θ ∧ θ3 + θ ∧ θ3 . (8.48)
r r Test by independent calculation
whether you can confirm the dif-
Since the torsion must vanish, Θi = 0, Cartan’s first structure equation ferentials (8.48).
from (8.13) implies
dθ μ = −ωμν ∧ θν . (8.49)
Connection forms
With (8.48), this suggests that the connection forms of a static,
spherically-symmetric metric are
a θ 0
ω01 = ω10 = , ω02 = ω20 = 0 , ω03 = ω30 = 0 , ?
eb
θ2 θ3 Why can none of the connection
ω21 = −ω12 = b , ω31 = −ω13 = b , forms in (8.50) depend on ϕ?
re re
cot ϑ θ3
ω2 = −ω3 =
3 2
. (8.50)
r
They satisfy the antisymmetry condition (8.42) and Cartan’s first struc-
ture equation (8.49) for a torsion-free connection.
For evaluating the curvature forms Ωμν , we first recall that the exterior
derivative of a one-form ω multiplied by a function f is
d( f ω) = d f ∧ ω + f dω
= (∂i f )dxi ∧ ω + f dω (8.51)
according to the (anti-)Leibniz rule (??).
Thus, we have for dω01
dω01 = (a e−b ) dr ∧ θ0 + a e−b dθ0
= (a e−b − a b e−b )e−b θ1 ∧ θ0 + (a e−b )2 θ1 ∧ θ0
=: A θ0 ∧ θ1 (8.52)
120 8 The Schwarzschild Solution
The components of the curvature tensor are given by (8.20), and thus the
components of the Ricci tensor in the tetrad {eα } are
Rμν = R̄λμλν = Ωλμ (eλ , eν ) . (8.56)
Thus, the components of the Ricci tensor in the Schwarzschild tetrad are
2a E
R00 = Ω10 (e1 , e0 ) + Ω20 (e2 , e0 ) + Ω30 (e3 , e0 ) = A + ,
r
2b E
R11 = Ω01 (e0 , e1 ) + Ω21 (e2 , e1 ) + Ω31 (e3 , e1 ) = −A + (8.57)
r
8.4 Solution of the field equations 121
1−E
R22 = Ω02 (e0 , e2 ) + Ω12 (e1 , e2 ) + Ω32 (e3 , e2 ) =: B +
r2
R33 = Ω03 (e0 , e3 ) + Ω13 (e1 , e3 ) + Ω23 (e2 , e3 ) = R22 (8.58)
The vacuum field equations now require that all components of the
Einstein tensor vanish. In particular, then,
2E
0 = G00 + G11 = (a + b ) (8.61)
r
shows that a + b = 0. Since a + b → 0 asymptotically for r → ∞,
integrating a + b from r → ∞ indicates that a + b = 0 everywhere, or
b = −a.
After multiplying with r2 , equation G00 = 0 itself implies that
C
rE = r + C ⇔ E =1+ , (8.63)
r
with an integration constant C to be determined.
122 8 The Schwarzschild Solution
Schwarzschild metric
We thus obtain the Schwarzschild solution for the metric,
2m 2 2 dr2
ds2 = − 1 − c dt + + r2 dϑ2 + sin2 ϑdϕ2 . (8.67)
r 2m
1−
r
The Schwarzschild metric (8.67) has an (apparent) singularity at r = 2m
or
2GM
r = Rs ≡ 2 , (8.68)
c
8.4 Solution of the field equations 123
We can now try and identify the two induced line elements from (8.69)
and (8.70) and find that this is possible if
1/2
1 2m
z = −1 = , (8.71)
1 − 2m/r r − 2m
Figure 8.3 Surface of rotation illustrating the spatial part of the Schwarz-
schild metric.
the components of the Einstein tensor following from them, had resulted
in the new components Ḡμν
a = −b + f (t) , (8.74)
Birkhoff’s theorem
This is Birkhoff’s theorem, which states that a spherically symmetric
solution of Einstein’s vacuum equations is necessarily static for r > 2m.
Chapter 9
Note that we have to differentiate and integrate with respect to the proper
?
time τ rather than the coordinate time t because the latter has no invariant
Recall the essential arguments for
physical meaning. In the Newtonian limit, τ = t.
the Lagrange function (9.1) and
The constant value of u, u allows that, instead of varying the action its physical interpretation.
b
S = −mc −u, u dτ , (9.4)
a
127
128 9 The Schwarzschild Spacetime
r2 ϑ̇ = const. = 0 . (9.9)
Effective Lagrangian
? Without loss of generality, we can thus restrict the discussion to motion
Derive the Lagrangian (9.10) in the equatorial plane, which simplifies the Lagrangian to
yourself and convince yourself of & '
all steps taken. 1 ṙ2
L= −(1 − 2m/r)c t +
2 ˙2
+ r ϕ̇ .
2 2
(9.10)
2 1 − 2m/r
Note that the effective potential has (and must have) the dimension of a
squared velocity.
Since it is our primary goal to find the orbit r(ϕ), we use r = dr/dϕ =
ṙ/ϕ̇ to transform (9.15) to
L2 2
ṙ2 + V(r) = ϕ̇2 r2 + V(r) = r + V(r) = E 2 . (9.17)
r4
Now, we substitute u ≡ 1/r and u = −r /r2 = −u2 r and find
u2
L 2 u4 4
+ V(1/u) = L2 u2 + (1 − 2mu) c2 + L2 u2 = E 2 (9.18)
u
or, after dividing by L2 and rearranging terms,
E 2 − c2 2mc2
u2 + u2 = + 2 u + 2mu3 . (9.19)
L2 L
mc2
u + u = + 3mu2 . (9.21)
L2
Note that this is the equation of a driven harmonic oscillator.
130 9 The Schwarzschild Spacetime
The fact that t and ϕ are cyclic coordinates in the Schwarzschild space-
time can be studied from a more general point of view. Let γ(τ) be a
geodesic curve with tangent vector γ̇(τ), and let further ξ be a Killing
vector field of the metric. Then, we know from (5.36) that the projection
of the Killing vector field on the geodesic is constant along the geodesic,
L2 dΦ
− L2 u2 u − r + =0 (9.30)
r4 dr
or, after dividing by −u2 L2 ,
1 dΦ
u + u = . (9.31)
L 2 u2
dr ?
Can you agree with the result
(9.31)?
Orbital equation in Newtonian gravity
In the Newtonian limit of the Schwarzschild solution, the potential and
its radial derivative are
GM dΦ GM
Φ=− , = 2 = GMu2 = mc2 u2 , (9.32)
r dr r
so that the orbital equation becomes
mc2
u + u = . (9.33)
L2
Compared to this, the equation of motion in the Schwarzschild case (9.21)
has the additional term 3mu2 . We have seen in (8.66) that m ≈ 1.5 km
in the Solar System. There, the ratio of the two terms on the right-hand
side of (9.21) is
1 dΦ(r) mc2
= 2 + 3mu2 (9.35)
L 2 u2
dr L
or
dΦ(r) mc2 3mL2
= mc2 u2 + 3mL2 u4 = 2 + 4 , (9.36)
dr r r
132 9 The Schwarzschild Spacetime
e = 0.9
e = 0.2 3
1
2
0.5
1
y in units of p
y in units of p
0 0
0 0.5 1 0 1 2 3
-1
-0.5
-2
-1 -3
-1 -0.5 0 0.5 1 -3 -2 -1 0 1 2 3
x in units of p
x in units of p
Figure 9.1 Numerical solutions of the orbital equation (9.21) for test parti-
cles, for different values of the orbital eccentricity e. All lengths, including the
mass m = 0.025, are scaled by the orbital parameter p. The orbits shown
begin at u = 1 + e with u = 0. For e = 0.2 (left), two orbits are shown, and
twelve orbits for e = 0.9 (right).
which leads to
mc2 mL2
Φ(r) = − − 3 (9.37)
r r
if we set the integration constant such that Φ(r) → 0 for r → ∞.
As a function of x ≡ r/Rs = r/2m, the effective potential V(r) from
(9.16) depends in an interesting way on L/(cRs ) = L/(2mc ≡ λ). The
dimensionless function
V(x) 1 λ2
v(x) := = 1 − 1 + (9.38)
c2 x x2
3
=1
= 3
2.5
=2
=3
effective potential v(x) := V(x)/c2
2 =4
1.5
0.5
-0.5
-1
0 1 2 3 4 5
radius x = r/Rs
Figure 9.2 Dimensionless effective potential v(x) for a test particle in the
Schwarzschild spacetime for various scaled angular momenta λ.
mc2 3m
u + u = + 2 (1 + e cos ϕ)2 . (9.44)
L2 p
u = u0 + u1 + u2 + u3 (9.47)
& '
1 3m e2 1
= (1 + e cos ϕ) + 2 1 + eϕ sin ϕ + 1 − cos 2ϕ .
p p 2 3
This solution of (9.44) has its perihelion at ϕ = 0 because the unperturbed
solution u0 was chosen to have it there. This can be seen by taking the
derivative with respect to ϕ,
e 3me + e ,
u = − sin ϕ + 2 sin ϕ + ϕ cos ϕ + sin 2ϕ (9.48)
p p 3
and verifying that u = 0 at ϕ = 0, i.e. the orbital radius r = 1/u still has
an extremum at ϕ = 0.
We now use equation (9.48) in the following way. Starting at the perihe-
lion at ϕ = 0, we wait for approximately one revolution at ϕ = 2π + δϕ
and see what δϕ has to be for u to vanish again. Thus, the condition for
the next perihelion is
3m + e ,
0 = − sin δϕ + sin δϕ + (2π + δϕ) cos δϕ + sin 2δϕ (9.49)
p 3
or, to first order in the small angle δϕ,
& '
3m 2e
δϕ ≈ 2δϕ + 2π + δϕ . (9.50)
p 3
9.3 Perihelion shift and light deflection 135
δϕ ≈ 43 (9.53)
For light rays, the condition 2L = −c2 that we had before for material
particles is replaced by 2L = 0. Then, (9.14) changes to
ṙ2 − E 2 L2
+ 2 =0 (9.54)
1 − 2m/r r
or
L2 2m
ṙ + 2 1 −
2
= E2 . (9.55)
r r
Changing again the independent variable to ϕ and substituting u = 1/r
leads to the equation of motion for light rays in the Schwarzschild
spacetime
E2
u2 + u2 = 2 + 2mu3 , (9.56)
L
which should be compared to the equation of motion for material parti-
cles, (9.19). Differentiation finally yields the orbital equation for light ?
rays in the Schwarzschild spacetime. Derive the orbital equation (9.56)
yourself.
Light rays in the Schwarzschild spacetime
Light rays (null geodesics) in the Schwarzschild spacetime follow the
orbital equation
u + u = 3mu2 . (9.57)
Compared to u on the left-hand side, the term 3mu2 is very small. In the
Solar System,
3mu2 3Rs Rs
= 3mu = ≤ ≈ 10−6 . (9.58)
u 2r R
Keplerian
4 m = 0.025
m = 0.050
m = 0.075
2
y in units of p
0
0 5 10
-2
-4
-10 -5 0 5 10
x in units of p
Figure 9.3 Numerical solutions of the orbital equation (9.57) for light rays,
compared to the Keplerian straight line, for different values of m. All lengths,
including the mass m, are scaled by the orbital parameter p. The orbits
shown begin at u = 1 with u = 0.
Thus, the light ray is almost given by the homogeneous solution of the
harmonic-oscillator equation u + u = 0, which is u0 = A sin ϕ + B cos ϕ.
We require that the closest impact at u0 = 1/b be reached when ϕ = π/2,
which implies B = 0 and A = 1/b, or
sin ϕ b
u0 = ⇒ r0 = . (9.59)
b sin ϕ
Note that this is a straight line in plane polar coordinates, as it should be!
Inserting this lowest-order solution as a perturbation into the right-hand
side of (9.57) gives
3m 2 3m
u + u = sin ϕ = 1 − cos 2
ϕ , (9.60)
b2 b2
for which particular solutions can be found using (9.45) and (9.46).
Combining this with the unperturbed solution (9.59) gives
sin ϕ 3m 3m 1
u= + 2 − 2 1 − cos 2ϕ . (9.61)
? b b 2b 3
Beginning with (9.60), confirm
the deflection angle (9.63). Given the orientation of our coordinate system, i.e. with the closest
approach at ϕ = π/2, we have ϕ ≈ 0 for a ray incoming from the left at
large distances. Then, sin ϕ ≈ ϕ and cos 2ϕ ≈ 1, and (9.61) yields
ϕ 2m
u≈ + . (9.62)
b b2
In the asymptotic limit r → ∞, or u → 0, this gives the angle
2m
|ϕ| ≈ . (9.63)
b
9.4 Spins in the Schwarzschild spacetime 137
Let us now finally study how a gyroscope with spin s is moving along a
geodesic γ in the Schwarzschild spacetime. Without loss of generality,
we assume that the orbit falls into the equatorial plane ϑ = π/2, and we
restrict the motion to circular orbits.
Then, the four-velocity of the gyroscope is characterised by u1 = 0 = u2
because both r = x1 and ϑ = x2 are constant.
The equations that the spin s and the tangent vector u = γ̇ of the orbit
have to satisfy are
s, u = 0 , ∇u s = 0 , ∇u u = 0 . (9.65)
The first is because s falls into a spatial hypersurface perpendicular to
the time-like four-velocity u, the second because the spin is parallel
transported, and the third because the gyroscope is moving along a
geodesic curve.
We work in the same tetrad {θ μ } introduced in (8.40) that we used to
derive the Schwarzschild solution. From (8.9), we know that
(∇u s)μ = dsμ + sν ωμν , u = u(sμ ) + ωμν (u)sν
= ṡμ + ωμν (u)sν = 0 , (9.66)
where the overdot marks the derivative with respect to the proper time τ.
With the connection forms in the Schwarzschild tetrad given in (8.50),
and taking into account that a = −b and cot ϑ = 0, we find for the
components of ṡ
ṡ0 = −ω01 (u)s1 = b e−b u0 s1 ,
e−b 3 3
ṡ1 = −ω10 (u)s0 − ω12 (u)s2 − ω13 (u)s3 = b e−b u0 s0 + u s ,
r
ṡ2 = −ω21 (u)s1 − ω23 (u)s3 = 0 ,
e−b 3 1
ṡ3 = −ω31 (u)s1 − ω32 (u)s2 = − u s , (9.67)
r
138 9 The Schwarzschild Spacetime
u̇0 = b e−b u0 u1 = 0 ,
e−b 3 2
u̇1 = −b e−b (u0 )2 − (u ) = 0 ,
r
u̇2 = 0 ,
e−b 1 3
u̇3 = − u u =0. (9.69)
r
The second of these equations implies
0 2
u 1
=− . (9.70)
? u 3 br
What is the physical meaning of
equation (9.70)?
9.4.2 Spin precession
Note that u̇μ = 0 for all μ according to (9.69). Using (9.69) and the
normalisation relation (u0 )2 − (u3 )2 = c2 , we obtain
& '
(u0 )2 u0 u3 3 0
−b u 3
˙s̄1 = b e−b 1 − s̄ = −cb e s̄ ,
(u3 )2 c u3
s̄˙2 = 0 ,
u0 1
s̄˙3 = cb e−b s̄ . (9.75)
u3
From now on, we shall drop the overbar, understanding that the si denote
the components of the spin with respect to the basis ēi .
Next, we transform the time derivative from the proper time τ to the
coordinate time t. Since
we have
u0 −a u0 b
t˙ = e = e , (9.77)
c c
or
dsi ṡi c ṡi −b
= = 0 e . (9.78)
dt t˙ u
Inserting this into (9.75) yields
dϕ ϕ̇ ce−b u3
ω≡ = = , (9.81)
dt t˙ r u0
which can be rewritten by means of (9.70),
2
u3 e−2b c2 b −2b c2 −2b
ω2 = =− e = e . (9.82)
u0 r 2 r 2r
Now, we use (9.82) to substitute the factor out front the final expression
and find the relation
# $
Ω2 = ω2 e−2b 1 + b r (9.87)
thus
r − 3m 1 − 3m/r
rb + 1 = = (9.89)
r − 2m 1 − 2m/r
and
1 − 3m/r 3m
Ω2 = ω2 e−2b = ω2 1 − . (9.90)
1 − 2m/r r
Recall that we have projected the spin into the three-dimensional space
perpendicular to the direction of motion. Thus, the result (9.91) shows
that s precesses retrograde in that space about an axis perpendicular to
the plane of the orbit, since u2 = 0.
After a complete orbit, i.e. after the orbital time τ = 2π/ω, the projection
of s onto the plane of the orbit has advanced by an angle
Ω 3m
φ = Ωτ = 2π = 2π 1 − < 2π , (9.92)
ω r
9.4 Spins in the Schwarzschild spacetime 141
according to (9.90). The spin thus falls behind the orbital motion; its
precession is retrograde. The geodetic precession frequency is
⎛ ⎞
φ − 2π ⎜⎜ ⎟⎟
= ω ⎜⎜⎜⎝ 1 − − 1⎟⎟⎟⎠
3m
ωs =
τ r
1/2
GM 3GM 3 (GM)3/2
≈− 3 2
=− (9.93)
r 2rc 2 c2 r5/2
which yields
& '−1/2
1 1
2m − dr = cdτ , (10.3)
r R
R R
r= (1 + cos η) , dr = − sin ηdη , (10.4)
2 2
143
144 10 Schwarzschild Black Holes
where we have used in the last step that 1 − cos η2 = sin2 η. This result
allows us to translate (10.3) into
√ √
R sin ηdr R R sin2 ηdη
√ =− √
2m(1 − cos η) 2 2m 1 − cos η
R3
=− (1 + cos η)dη . (10.6)
8m
R3 R3
cτ = (η + sin η) , cdτ = (1 + cos η) dη . (10.7)
8m 8m
4
radial distance r/m
0
0 2 4 6 8 10 12 14 16 18
proper time c/m
dr dr E/c
ṙ = t˙ = − . (10.9)
dt dt 1 − 2m/r
?
Next, we introduce a new, convenient radial coordinate r̄ such that Before you read on, find the func-
tion r̄(r) yourself, given (10.10).
dr
dr̄ = . (10.10)
1 − 2m/r
dr r/2m − 1 + 1 dr
= dr = dr +
1 − 2m/r r/2m − 1 r/2m − 1
) r *
= dr + 2m d ln −1 , (10.11)
2m
giving ) r *
r̄ = r + 2m ln −1 . (10.12)
2m
E/c dr E dr̄
ṙ = − =− , (10.13)
1 − 2m/r dt c dt
2
E 2 dr̄
→ E2 , (10.16)
c2 dt
and thus
dr̄
→ ±c . (10.17)
dt
Of the two signs, we have to select the negative because of r̄ → −∞, as
(10.15) shows. Therefore, an approximate solution of the equation of
motion near the singularity is r̄ ≈ c(t−t0 ) with an arbitrary constant t0 . To
be specific, we set t = 0 when r = 6m, the radius of the innermost stable
circular orbit defined in Sect. 9.2. There, r̄0 = 2m(3 + ln 2) according to
(10.12) and thus
r̄ ≈ −ct + 2m(3 + ln 2) . (10.18)
or
dr 2m
= ±c 1 − , (10.23)
dt r
10.1.3 Curvature at r = 2m
and
2a 2m 1
R22 = − 1− + 2 1 − e−2b
r r r
2m(r − 2m) 2m
=− 3 + 3 = 0 = R33 , (10.27)
r (r − 2m) r
i.e. the components of the Ricci tensor in the Schwarzschild tetrad remain
perfectly regular at the Schwarzschild radius!
We shall now try to remove the obvious problems with the Schwarzschild
coordinates by transforming (ct, r) to new coordinates (u, v), leaving ϑ
and ϕ, requiring that the metric can be written as
assuming that f will turn out to depend on r only since any dependence
on time and on the angles ϑ and ϕ is forbidden in a static, spherically-
symmetric spacetime. Then,
dr 2m
vr̄ = vr = 1 − vr (10.36)
dr̄ r
?
and the same for u. Repeat the calculation in (10.36)
with the coordinate u.
Now, we add the two equations containing F(r̄) and then add and subtract
from the result twice the third equation from (10.37). This yields
Taking the square root of this equation, we can choose the signs. The
choice
Taking partial derivatives once with respect to t and once with respect to
r̄ allows us to combine these equations to find the wave equations
which are solved by any two functions h± propagating with unit velocity,
where the signs were chosen such as to satisfy the sign choice in (10.39).
Now, since
where the primes denote derivatives with respect to the functions’ argu-
ments, we find from (10.37)
# $ # $
F(r̄) = h+ − h− 2 − h+ + h− 2 = −4h+ h− . (10.44)
150 10 Schwarzschild Black Holes
We start from outside the Schwarzschild radius, assuming r > 2m, where
also F(r̄) > 0 according to (10.35). The derivative of (10.44) with respect
to r̄ yields
# $
F (r̄) = −4 h+ h− + h+ h− (10.45)
or, with (10.44),
F h+ h−
= + . (10.46)
F h+ h−
the derivative of (10.44) with respect to time yields
# $ h+ h−
0 = −4 h+ h− − h+ h− ⇒ − =0. (10.47)
h+ h−
Now, the left-hand side depends on r̄, the right-hand side on the indepen-
dent variable r̄ + t. Thus, the two sides of this equation must equal the
same constant, which we call 2C:
or
h+ = const.eC(r̄+ct) . (10.52)
) r *2mC
v= −1 eCr sinh(Cct) , (10.56)
2m
Kruskal-Szekeres coordinates
The Kruskal-Szekeres coordinates (u, v) are related to the Schwarz-
schild coordinates (ct, r) by
) ct *
r
u= − 1 er/4m cosh ,
2m 4m
) ct *
r
v= − 1 er/4m sinh , (10.58)
2m 4m
and the scale function f is
32m3 −r/2m
f2 = e . (10.59)
r
We have (or rather, Martin Kruskal has) thus achieved our goal to replace
the Schwarzschild coordinates by others in which the Schwarzschild
metric remains prefectly regular at r = 2m. Appendix C shows how
space-times can be compactly represented in Penrose-Carter diagrams.
1.5
+
1
=t
1,
r=
0.5 II
0 III I
v
-0.5 IV r=
-1
1,
t
=
-
-1.5
-2
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
u
Figure 10.3 Illustration of the Kruskal continuation in the u-v plane. The
Schwarzschild domain r > 2m is shaded in red and marked with I, the
forbidden region r < 0 is shaded in gray.
We now want to study the collapse of an object, e.g. a star. For this
purpose, coordinates originally introduced by Arthur S. Eddington and
re-discovered by David R. Finkelstein are convenient, which are defined
by
r = r ,
ϑ = ϑ , ϕ = ϕ
) r *
ct = ct − 2m ln ± ∓1 (10.64)
2m
in analogy to the radial coordinate r̄ from (10.12), where the upper
?
and lower signs in the second line are valid for r > 2m and r < 2m,
What do the light cones look
respectively.
like in the coordinates given in
(10.64)? Since
⎧ ∓1/2
⎪
⎪
⎪
⎨ 2m − 1 (r > 2m)
r
±ct /4m
e±ct/4m = e ⎪
⎪ ∓1/2 , (10.65)
⎪
⎩ − 2m + 1
r
(r < 2m)
inserting these expressions into the Kruskal-Szekeres coordinates (10.58)
shows that they are related to the Eddington-Finkelstein coordinates by
er/4m ct /4m r − 2m −ct /4m
u= e + e
2 2m
er/4m ct /4m r − 2m −ct /4m
v= e − e , (10.66)
2 2m
such that
r − 2m r/2m r − 2m u + v
e = u2 − v 2 , ect /2m = . (10.67)
2m 2m u − v
The first of these equations shows again that r can be uniquely determined
from u and v if u2 − v2 > −1. The second equation determines t uniquely
10.3 Physical meaning of the Kruskal continuation 155
1.5
time t'/2m
0.5
0
0 0.5 1 1.5 2
radius r'/2m
respectively. Then according to (4.48) the redshift of the light from the
torch as seen by the observer is
νem k, v
1+z= = , (10.75)
νobs k, u
dr
cdtret = cdt − , (10.76)
1 − 2m/r
thus
dr2 2c dtret dr
c2 dt2 = c2 dtret
2
+ + (10.77)
(1 − 2m/r)2 1 − 2m/r
and the line element of the Schwarzschild metric transforms to
2m 2 2
ds2 = − 1 − c dtret − 2c dtret dr + r2 dΩ2 . (10.78)
r
which is possible for outgoing light rays only if dtret = 0. This shows
that such light rays must propagate along r, or k ∝ ∂r , which is of course
a consequence of our using the retarded time tret . We set the amplitude
of k such that k = κ∂r . Since ∂r , ∂r = 0 in the coordinates of the line
element (10.78), the null condition on k is satisfied for any κ.
For a distant observer at a fixed distance r 2m, the line element (10.78)
simplifies to
ds2 ≈ −c2 dtret
2
, (10.80)
which shows that the retarded time tret is also the distant observer’s
proper time.
Expanding now the astronaut’s velocity as
we find
k, v = κ∂r , v = κ gtret r t˙ret = −κt˙ret (10.82)
because ∂r , ∂r = 0 according to the metric with the line element (10.78).
The dots in these equations indicate derivatives with respect to the astro-
naut’s proper time.
158 10 Schwarzschild Black Holes
Far away from the black hole, the metric can be assumed to be Min-
kowskian. For a distant observer at rest, the four-velocity is u = ∂t . In
Minkowski coordinates, the wave vector of the light ray must be
k = κ (∂t + ∂r ) , (10.83)
ṙ/c
1 + z ≈ t˙ret = t˙ − . (10.85)
1 − 2m/r
see (9.12). To be specific, we set the constant E such that the astronaut
is at rest at infinite radius, E 2 = c2 . Requiring that the astronaut’s proper
time increases with the coordinate time, t˙ > 0 and E < 0, hence we must
set E = −c. Since ṙ < 0 for the infalling astronaut,
√
ṙ = −c 1 − δ , (10.88)
with δ ≡ 1 − 2m/r.
The redshift (10.85) can now be written
1 √ 2
1+z= 1+ 1−δ ≈ (10.89)
δ δ
to leading order close to the Schwarzschild radius, where δ → 0+. We
have seen in (10.21) that the radial coordinate of the falling astronaut
is well approximated by r ≈ 2m(1 + 2e2−ct/2m ) near the Schwarzschild
radius if the coordinate clock is set to zero at r = 6m. This enables us to
approximate δ by
2m r − 2m
δ=1− ≈ = 2e2−ct/2m (10.90)
r 2m
and the redshift by
1 + z ≈ ect/2m−2 . (10.91)
10.4 Redshift approaching the Schwarzschild radius 159
q q
F=− 2
cdt ∧ dr = − 2 e−a−b θ0 ∧ θ1 . (11.1)
r r
?
We shall verify below that a = −b also for a Schwarzschild solution with Why would the Faraday-2-form
charge, so that the exponential factor will become unity later. be given by (11.1)? Recall the
meaning of the components of
The electromagnetic energy-momentum tensor the electromagnetic field tensor.
& '
μν 1 μλ ν 1 μν αβ
T = F F λ − g F Fαβ (11.2)
4π 4
161
162 11 Charged, Rotating Black Holes
Δ r2 dr2
ds2 = − 2 dt2 + + r2 dϑ2 + sin2 ϑdϕ2 . (11.9)
r Δ
This is the Reissner-Nordström solution.
Of course, for q = 0, the Reissner-Nordström solution returns to the
Schwarzschild solution.
11.1 The Reissner-Nordström solution 163
1
∗(θ0 ∧ θ1 ) = g00 g11 ε01αβ θα ∧ θβ (11.12)
2
1
= − θ2 ∧ θ3 − θ3 ∧ θ2 = −θ2 ∧ θ3 .
2
Inserting the Schwarzschild tetrad from (8.40) yields
(4)
g=σ+g (11.15)
and have
σ = σab (xi ) dxa ⊗ dxb . (11.16)
The coefficients σab are scalar products of the two Killing vector fields k
and m,
−k, k k, m
(σab ) = , (11.17)
k, m m, m
√
ρ≡ − det σ = k, km, m + k, m2 . (11.18)
Δ := r2 − 2mr + Q2 + a2 , ρ2 := r2 + a2 cos2 ϑ ,
Σ := (r + a ) − a Δ sin ϑ .
2 2 2 2 2 2
(11.19)
Kerr-Newman solution
With these definitions, we can write the coefficients of the metric for a
charged, rotating black hole in the form
2mr − Q2 a2 sin2 ϑ − Δ
gtt = −1 + = ,
ρ2 ρ2
2mr − Q2 r 2 + a2 − Δ
gtϕ = − a sin2
ϑ = − a sin2 ϑ ,
ρ2 ρ2
ρ2 Σ2
grr = , gϑϑ = ρ2 , gϕϕ = 2 sin2 ϑ . (11.21)
Δ ρ
Figure 11.3 Roy Kerr (born 1934), New Zealand mathematician. Source:
Wikimedia Commons
Also without derivation, we quote that the vector potential of the rotating,
charged black hole is given by the 1-form
qr
? A = − 2 cdt − a sin2 ϑdϕ , (11.22)
Why is it plausible for A to have ρ
the form (11.22)? Beginning from which we obtain the Faraday 2-form
there, derive the Faraday-2-form
(11.23) yourself.
11.2 The Kerr-Newman solution 167
q 2
F = dA = r − a2 cos2 ϑ dr ∧ cdt − a sin2 ϑdϕ
ρ 4
2qra
+ 4 sin ϑ cos ϑdϑ ∧ r2 + a2 dϕ − acdt . (11.23)
ρ
For a = 0, this trivially returns to the field (11.1) for the Reissner-
Nordström solution. Sufficiently far away from the black hole, such that
a r, we can approximate to first order in a/r and write
q 2qa
F= dr ∧ cdt − a sin 2
ϑdϕ + sin ϑ cos ϑdϑ ∧ dϕ . (11.24)
r2 r
The field components far away from the black hole can now be read off
the result (11.24). Using the orthonormal basis
1 1
et = ∂ct , er = ∂r , eϑ = ∂ϑ , eϕ = ∂ϕ , (11.25)
r r sin ϑ
we find in particular for the radial component Br of the magnetic field
2qa
Br = F(eϑ , eϕ ) = cos ϑ . (11.26)
r3
In the limit of large r, the electric field thus becomes that of a point
?
charge q at the origin, and the magnetic field attains a characteristic
Find the remaining components
dipolar structure.
of the electromagnetic field of the
The Biot-Savart law of electrodynamics implies that a charge q with Kerr-Newman solution.
mass M on a circular orbit with angular momentum L has the magnetic
dipole moment
qL
μ = g , (11.27)
2Mc
where g is the gyromagnetic moment.
A magnetic dipole moment μ creates the dipole field
g freely
angular
For an interpretation of Ω, we note that freely-falling test particles on
radial orbits have zero angular momentum and thus u, m = 0. By
(11.34), this implies
νo k̃, uo
= , (11.43)
νs k̃, us
where k̃ is the wave vector of the light.
Observers at rest in a stationary spacetime have four-velocities propor-
tional to the Killing vector field k,
k
u= √ , hence k = −k, k u . (11.44)
−k, k
and k̃, ks = k̃, ko . Using this in a combination of (11.43) and (11.44),
we obtain √ √
νo k̃, ko −k, ks −k, ks
= √ = √ . (11.46)
νs k̃, ks −k, ko −k, ko
For an observer at rest far away from the black hole, k, ko ≈ −1, and
the redshift becomes
νs 1
1+z= ≈ √ = (−gtt )−1/2 , (11.47)
νo −k, ks
which tends to infinity as the source approaches the static limit.
The minimum and maximum angular velocities ω± from (11.38) both
become equal to Ω for
2
gtϕ gtt
Ω2 = = ⇒ g2tϕ − gtt gϕϕ = 0 . (11.48)
gϕϕ gϕϕ
Interestingly, writing the expression from (11.48) with the metric coeffi-
cients (11.21) leads to the simple result
0 = Δ = r2 − 2mr + Q2 + a2 , (11.51)
Angular frequency of H
The hypersurface H is rotating with the constant angular velocity
a
ΩH = , (11.54)
rH2 + a2
ergosphere
2 static limit
horizon
r/m
z/m
0
0 1 2
-1
-2
-2 -1 0 1 2
x/m
Figure 11.4 Static limit, horizon, and ergosphere for a Kerr black hole with
a = 0.75.
3
2
1
0
-1 3 6
-2 a = 0.50
a = 0.99
-3
-2 0 2 4 6 8 10
3
2
1 r/m
0
-1 3 6
-2 L= 2
L = -2
-3
-2 0 2 4 6 8 10
Figure 11.5 Trajectories of test particles in the equatorial plane of the Kerr
metric. All orbits begin at r = 10m and ϕ = 0. Top: Orbits with angular
momentum L = 0 for a = 0.5 and a = 0.9. Bottom: orbits with angular
momenta L = ±2 for a = 0.99.
uϕ uϕ c2
ωl = − ⇒ ut ut = . (11.69)
ut ut ωl − 1
Finally, using (11.67) and (11.69), we have
gtt + gtϕ ω
− u2t = −ut ut (gtt + gtϕ ω) = c2 . (11.70)
1 − ωl
If we substitute ω from (11.68) here, we obtain after a short calculation
?
Repeat the calculations leading g2tϕ − gtt gϕϕ
e2 = u2t = c2 . (11.71)
to (11.76) and (11.78) in compo- gϕϕ + 2lgtϕ + l2 gtt
nents. Why can (11.77) be called
a perpendicular projector? Can It is shown in the In-depth box “Ideal hydrodynamics in general relativity”
you confirm the non-relativistic on page 175 that the relativistic Euler equation reads
limit (11.79)?
ρc2 + p ∇u u = −c2 dp − u(p)u , (11.80)
where ρc2 and p are the density and the pressure of the ideal fluid.
Applying this equation to the present case of a stationary flow, we first
observe that
0 = Lu p = u(p) , (11.81)
thus the second term on the right-hand side of (11.80) vanishes.
Next, we introduce the dual vector u belonging to the four-velocity u.
In components, (u )μ = gμν uν = uμ . Then, from (5.32),
Lu u = uν ∂ν uμ + uν ∂μ uν = uν ∇ν uμ + uν ∇μ uν , (11.82)
μ
L f x w = f L x w + w(x)d f . (11.84)
L f u u = f Lu u − c2 d f = f ∇u u − c2 d f , (11.85)
?
making use of (11.83) in the last step.
Can you confirm (11.84)?
On the other hand, f u = u/ut = k + ωm because of (11.65), which allows
us to write
L f u u = Lk!"
u +Lωm u . (11.86)
=0
11.3 Motion near a Kerr black hole 175
π⊥ := 14 + c−2 u ⊗ u , (11.77)
∂t ρ + ∇ · #ρv $ = 0 ,
v + ∇P = 0 .
∂tv + v · ∇ (11.79)
ρ
176 11 Charged, Rotating Black Holes
f ∇u u = c2 d f + jdω . (11.89)
e(1 − ωl)
∇u u = (1 − ωl)de − eldω − eωdl + jdω
c2
= (1 − ωl)de − eωdl , (11.91)
dp ωdl
= −d ln e + , (11.93)
ρc2 +p 1 − ωl
Accretion tori
We now insert the metric coefficients (11.21) for the Kerr-Newman
solution to obtain the surfaces of constant pressure and constant l.
Assuming further a = 0, we obtain the isobaric surfaces of the accretion
flow onto a Schwarzschild black hole. With
2m
gtt = −1 + , gϕϕ = r2 sin2 ϑ , gtϕ = 0 , (11.96)
r
we find
r l2
− = const. (11.97)
r − 2m r2 sin2 ϑ
This describes toroidal surfaces around black holes, the so-called ac-
cretion tori.
10
5
z/2m 0
-5
-10
-10 10
-5 5
0 0
x/2m 5 -5 y/2m
10 -10
the black hole, its entropy would be gone, violating the second law of
thermodynamics. The same holds for gas accreted by the black hole: Its
entropy would be removed from the outside world, leaving the entropy
there lower than before.
If, however, the increased mass of the black hole led to a suitably in-
creased entropy of the black hole itself, this violation of the second law
could be remedied.
Analogy between area and entropy
Any mass and angular momentum swallowed by a black hole leads to
an increase of the area (11.98), which makes it appear plausible that
the area of a black hole might be related to its entropy.
Following Bekenstein (1973), we shall now work out this relation.
Beginning with the scaled area α = r+2 + a2 from (11.98), we have
# $
dα = 2 r+ dr+ + a · da . (11.100)
Inserting r+ from (11.99) and using that
r+ − r− =: δr = 2 m2 − Q2 − a2 , (11.101)
we find directly
& '
r+ δr + 2r+ m 2r+ Q 2r+
dα = 2 dm − dQ + 1 − a · da . (11.102)
δr δr δr
11.4 Entropy and temperature of a black hole 179
and
δr − 2r+ = − (r+ + r− ) = −2m , (11.104)
we can bring (11.102) into the form
4r+2 4r+ Q 2m
dα = dm − dQ − a · da . (11.105)
δr δr δr
Now, we need to take into account that the scaled angular momentum a
can change by changing the angular momentum L or the mass m. From
the definition (11.20), we have
⎛ ⎞
G ⎜⎜ dL L ⎟⎟ G dL adm
da = 3 ⎜⎜⎝ − 2 dm⎟⎟⎠ = 3 − . (11.106)
c m m c m m
4α 4r+ Q 4G a · dL
dα = dm − dQ − 3 , (11.107) ?
δr δr c δr
Confirm equation (11.107) by
Solving equation (11.107) for dm yields your own calculation.
· dL
dm = Θdα + ΦdQ + Ω (11.108)
Θ :=
δr
, Φ :=
r+ Q
, := G a .
Ω (11.109)
4α α c3 α
This reminds of the first law of thermodynamics if we tentatively asso-
ciate m with the internal energy, α with the entropy and the remaining
terms with external work.
Let us now see whether a linear relation between the entropy S and the
area α will lead to consistent results. Thus, assume S = γα with some
constant γ to be determined. Then, a change δα will lead to a change
δS = γδα in the entropy.
Bekenstein showed that the minimal change of the effective area is twice
the squared Planck length (1.5), thus
2G
δα = . (11.110)
c3
On the other hand, he identified the minimal entropy change of the black
hole with the minimal change of the Shannon entropy, which is derived
from information theory and is
δS = kB ln 2 , (11.111)
180 11 Charged, Rotating Black Holes
2G
kB ln 2 = δS = γδα = γ (11.112)
c3
? fixes the constant γ to
Look up Bekenstein’s arguments ln 2 kB c3
leading to the normalisation γ= . (11.113)
2 G
(11.113) of the black-hole en-
tropy (Bekenstein, J., Black
Bekenstein entropy
Holes and Entropy. Phys. Rev.
D 7 (1973) 2333). The Bekenstein entropy of a black hole is
ln 2 c3 kB
S = A, (11.114)
8π G
where A is the area of the black hole.
The quantity Θ defined in (11.109) must then correspond to the tempera-
ture of the black hole. From (11.108), we have on the one hand
∂m
Θ= . (11.115)
∂α Q,L
mc4
E = Mc2 = . (11.117)
G
Then,
∂S G ∂S ln 2 kB ∂α
= 4 = . (11.118)
∂E V c ∂m Q,L 2 c ∂m Q,L
π2 kB4
L = σAT 4 , σ= . (11.120)
603 c2
For an uncharged and non-rotating black hole, δr = 2m and A = 16πm2 ,
thus its temperature is
1 c 1 c3
T= = . (11.121)
4 ln 2 kB m 4 ln 2 kB GM
Defining the Planck temperature by
MPl c2
T Pl := = 1.42 · 1032 K (11.122)
kB
in terms of the Planck mass MPl = 2.2 · 10−5 g from (1.4), we can write
T Pl MPl
T= . (11.123)
4 ln 2 M
For a black hole of solar mass, M = M
= 2.0 · 1033 g, the temperature is
Homogeneous, Isotropic
Cosmology
183
184 12 Homogeneous, Isotropic Cosmology
?
where the overdots and primes denote derivatives with respect to ct and
Carry out the calculations lead-
r, respectively.
ing to (12.4) yourself. Can you
confirm the results?
for the connection 1-forms ωμν . From (12.5) and the results (12.4), we
can read off
R −b 2
ω01 = ω10 = a e−b θ0 + ḃe−a θ1 , ω12 = −ω21 = − e θ ,
R
Ṙ −a 2 R
ω02 = ω20 = e θ , ω13 = −ω31 = − e−b θ3 ,
R R
Ṙ cot ϑ 3
ω03 = ω30 = e−a θ3 , ω23 = −ω32 = − θ . (12.6)
R R
Ω01 = dω01 ≡ E θ0 ∧ θ1 ,
Ω02 = dω02 + ω01 ∧ ω12 ≡ Ẽ θ0 ∧ θ2 + H θ1 ∧ θ2 ,
Ω03 = dω03 + ω01 ∧ ω13 + ω02 ∧ ω23 = Ẽ θ0 ∧ θ3 + H θ1 ∧ θ3 ,
Ω12 = dω12 + ω10 ∧ ω02 ≡ −H θ0 ∧ θ2 + F̃ θ1 ∧ θ2 ,
Ω13 = dω13 + ω10 ∧ ω03 + ω12 ∧ ω23 = −H θ0 ∧ θ3 + F̃ θ1 ∧ θ3 ,
Ω23 = dω23 + ω20 ∧ ω03 + ω21 ∧ ω13 ≡ F θ2 ∧ θ3 , (12.7)
where the functions
E = e−2a b̈ − ȧḃ + ḃ2 − e−2b a − a b + a2 ,
e−2a e−2b
Ẽ = R̈ − ȧṘ − a R ,
R R
e−a−b
H= Ṙ − a Ṙ − ḃR ,
R
1
F = 2 1 − R2 e−2b + Ṙ2 e−2a ,
R
e−2a e−2b # $
F̃ = ḃṘ + b R − R (12.8) ?
R R Repeat the calculations leading to
were defined for brevity. (12.8) yourself, beginning read-
ing off the connection forms
According to (8.20), the curvature forms imply the components
(12.6).
Rαβ = Ωλα (eλ , eβ ) (12.9)
of the Ricci tensor, for which we obtain
R00 = −E − 2Ẽ , R01 = −2H , R02 = 0 = R03 ,
R11 = E + 2F̃ , R12 = 0 = R13
R22 = Ẽ + F̃ + F = R33 , R23 = 0 , (12.10)
the Ricci scalar
R = (E + 2Ẽ) + (E + 2F̃) + 2(Ẽ + F̃ + F)
= 2(E + F) + 4(Ẽ + F̃) , (12.11)
186 12 Homogeneous, Isotropic Cosmology
We can now state and prove Birkhoff’s theorem in its general form:
Birkhoff’s generalised theorem
Every C 2 solution of Einstein’s vacuum equations which is spherically
symmetric in an open subset U ⊂ M is locally isometric to a domain
of the Schwarzschild-Kruskal solution.
The proof proceeds in four steps:
e−2a
− ȧ + ḃ − a − b = 0 , (12.19)
R
implies ȧ = a , which again leads to R = ∞ through G00 = 0.
This shows that the metric reduces to the Schwarzschild metric in all
relevant cases.
Cavity in spherically-symmetric spacetime
It is a corollary to Birkhoff’s theorem that a spherical cavity in a ?
spherically-symmetric spacetime has the Minkowski metric. Indeed, Compare Birkhoff’s to Newton’s
Birkhoff’s theorem says that the cavity must have a Schwarzschild theorem.
metric with mass zero, which is the Minkowski metric.
There are good reasons to believe that the Universe at large is isotropic
around our position. The most convincing observational data are pro-
vided by the cosmic microwave background, which is a sea of blackbody
radiation at a temperature of (2.725 ± 0.001) K whose intensity is almost
exactly independent of the direction into which it is observed.
There is furthermore no good reason to believe that our position in the
Universe is in any sense prefered compared to others. We must therefore
conclude that any observer sees the cosmic microwave background as
188 12 Homogeneous, Isotropic Cosmology
Consider now the curvature tensor (3) R̄ induced on Σt (i.e. the curvature
tensor belonging to the metric h induced on Σt ). We shall write it in
components with its first two indices lowered and the following two
indices raised,
kl
(3)
R̄ = (3) R̄i j . (12.20)
12.2 Homogeneous and isotropic spacetimes 189
In this way, (3) R̄ represents a linear map from the vector space of 2-
- -
forms 2 into 2 , because of the antisymmetry of (3) R̄ with respect to
permutations of the first and the second pairs of indices. Thus, it defines
an endomorphism
.2 .2 kl
L: → , (Lω)i j = (3) R̄i j ωkl . (12.21) Caution Recall that an endo-
morphism is a linear map of a
Due to the symmetry (3.81) of (3) R̄ upon swapping the first with the vector space into itself.
second pair of indices, the endomorphism L is self-adjoint. In fact, for
-
any pair of 2-forms α, β ∈ 2 ,
kl
α, Lβ = (3) R̄i j αi j βkl = (3) R̄i jkl αi j βkl = (3) R̄kli j αi j βkl
ij
= (3) R̄kl αi j βkl = β, Lα , (12.22)
k + fk f k
fi + f j = . (12.38)
ψ
Since the two sides of these equations (one for each combination of i
and j) depend on different sets of variables, the second derivatives fi
and f j must all be equal and constant, and thus the fi must be quadratic
in xi with a coefficient of xi 2 which is independent of xi . Therefore, we
can write
k i2
3
ψ=1+ x (12.39)
4 i=1
Robertson-Walker metric
According to the preceding discussion, the homogeneous and isotropic
spatial hypersurfaces Σt must have a metric h with a line element of the
form 83 i2 3
i=1 dx
dl2 = , r 2
≡ xi 2 . (12.40)
(1 + kr2 /4)2 i=1
a(t) dxi
θ0 = cdt , θi = , (12.42)
1 + kr2 /4
dθ0 = 0 ,
ȧ dt ∧ dxi a k
dθi = − x j dx j ∧ dxi
1 + kr2 /4 (1 + kr2 /4)2 2
ȧ 0 kx j i
= θ ∧ θi + θ ∧ θj . (12.43)
ca 2a
Since the exterior derivative of the metric is dg = 0, Cartan’s first
structure equation (8.13) implies
Cartan’s second structure equation (8.13) then gives the curvature forms
ä 0
Ω0i = dω0i + ω0k ∧ ωki = θ ∧ θi , (12.48)
c2 a
k + ȧ2 /c2 i
Ωij = dωij + ωi0 ∧ ω0j + ωik ∧ ωkj = θ ∧ θj ,
a2
from which we obtain the components of the Ricci tensor
as
3ä ä k + ȧ2 /c2
R00 = − , R11 = R22 = R33 = + 2 . (12.50)
c2 a c2 a a2
12.3 Friedmann’s equations 193
ä k + ȧ2 /c2
R = Rμμ = 6 + . (12.51)
c2 a a2
) p*
T = ρ + 2 u ⊗ u + pg (12.53)
c
Adding a third of the first equation to the second, and re-writing the first
equation, we find Friedmann’s equations.
194 12 Homogeneous, Isotropic Cosmology
Friedmann’s equations
For a spatially homogeneous and isotropic spacetime with the
Robertson-Walker metric (12.41), Einstein’s field equations reduce
to Friedmann’s equations,
If we eliminate
3p
6ȧä = −8πGaȧ ρ + 2 + 2Λc2 aȧ (12.57)
c
We can use the first law of thermodynamics here because isotropy forbids
any energy currents, thus no energy can flow into or out of the volume
a3 .
Equation (12.59) yields
3p 2
a3 ρ̇ + 3ρa2 ȧ + a ȧ = 0 , (12.60)
c2
which is identical to (12.58). This demonstrates that (12.58) simply
?
expresses energy-momentum conservation. Consequently, one can show
Why are energy and momentum
that it also follows from the contracted second Bianchi identity, ∇ · T = 0.
conserved here, but not in gen-
Two limits are typically considered for (12.58). First, if matter moves eral?
non-relativistically, p ρc2 , and we can assume p ≈ 0. Then,
ρ̇ ȧ
= −3 , (12.61)
ρ a
which implies
ρ = ρ0 a−3 (12.62)
if ρ0 is the density when a = 1.
Second, relativistic matter has p = ρc2 /3, with which we obtain
ρ̇ ȧ
= −4 (12.63)
ρ a
and thus
ρ = ρ0 a−4 . (12.64)
This shows that the density of non-relativistic matter drops as expected
in proportion to the inverse volume, but the density of relativistic matter
drops faster by one order of the scale factor. An explanation will be
given below.
196 12 Homogeneous, Isotropic Cosmology
dt0 dt1
= . (12.70)
a(t0 ) a(t1 )
We can now identify the time intervals dt0,1 with the inverse frequencies
of the emitted and observed light, dti = νi−1 for i = 0, 1. This shows that
the emitted and observed frequencies are related by
ν0 a(t1 )
= . (12.71)
ν1 a(t0 )
Since the redshift z is defined in terms of the wavelengths as
λ 1 − λ0 ν 0 − ν 1
z= = , (12.72)
λ0 ν1
we find that light emitted at t0 and observed at t1 is redshifted by
λ1 a(t1 )
1+z= = . (12.73)
λ0 a(t0 )
12.4 Density evolution and redshift 197
Cosmological redshift
The expansion or contraction of spacetime according to Friedmann’s
equations causes the wavelength of light to be increased or decreased
in the same proportion as the universe itself expands or contracts.
We can now interpret the result (12.64) that the density of relativistic
matter drops by one power of a more than expected by mere dilution:
as the universe expands, relativistic particles are redshifted by another
factor a and thus loose energy in addition to their dilution.
Before we proceed, we bring the spatial line element dl from (12.40) into
a different form. We first write it in terms of spherical polar coordinates
as
dr2 + r2 (dϑ2 + sin2 ϑdϕ2 )
dl2 = (12.74)
(1 + kr2 /4)2
and introduce a new radial coordinate u defined by
r
u= . (12.75)
1 + kr2 /4
and thus
dr du
= √ . (12.79)
1 + kr /4
2
1 − ku2
In terms of the new radial coordinate u, we can thus write the spatial line
element of the metric in the frequently used form
du2
dl2 = + u2 dΩ2 , (12.80)
1 − ku2
where dΩ abbreviates the solid-angle element. The constant k can be
positive, negative or zero, but its absolute value does not matter since it
198 12 Homogeneous, Isotropic Cosmology
which is equivalent to
& '
du2
ds2 = −c2 dt2 + a2 (t) + u2 dϑ2 + sin2 ϑdϕ2 (12.84)
1 − ku2
with u related to w by (12.82), and the scale factor a(t) satisfies the
Friedmann equations (12.55).
Metrics with line elements of the form (12.83) or (12.84) are called
Robertson-Walker metrics, and Friedmann-Lemaître-Robertson-Walker
metrics if their scale factor satisfies Friedmann’s equations.
Chapter 13
We had seen in (6.16) that the separation vector n between two geodesics
out of a congruence evolves in a way determined by the equation of
geodesic deviation or Jacobi equation
ωo = k, uo , (13.2)
∇k Ei = 0 = ∇k̃ Ei (i = 1, 2) . (13.3)
199
200 13 Relativistic Astrophysics
Notice that the parallel transport along a null geodesic implies that the
Ei remain perpendicular to k̃,
In a coordinate basis {eα } and its conjugate dual basis {θi }, they can be
written as
Ei = Eiα eα with Eiα = θα (Ei ) . (13.5)
The separation vector n between rays of the bundle can now be expanded
into the basis E1,2 ,
n = nα eα = N i Ei , (13.6)
showing that its components nα in the basis {eα } are
and thus
Ei ∇2k̃ N i = R̄(k̃, E j )k̃ N j . (13.10)
Finally, we multiply equation (13.10) with E i from the left and use the
orthonormality of the vectors E1,2 ,
2 3
E i , E j = δij , (13.11)
It will turn out convenient to introduce the Weyl tensor C̄, whose compo-
nents are determined by
R
R̄αβγδ = C̄αβγδ + gα[γ Rδ]β − gβ[γ Rδ]α − gα[γ gδ]β . (13.13)
3
13.2 Gravitational lensing 201
and using E i , k̃ = 0 = k̃, k̃ together with (13.11), we find that we can
write (13.12) as
1
∇2k̃ N i = − δij R k̃, k̃ + C ij N j . (13.17)
2
Some further insight can be gained by extracting the trace-free part from
T . Since the trace is
Tr T = −R k̃, k̃ + Tr C , (13.20) ?
Why is there a factor of 1/2 in
the trace-free part of T is
front of the trace on the left-hand
1 1 side of (13.21)?
T− Tr T 12 = C − Tr C 12 =: Γ , (13.21)
2 2
202 13 Relativistic Astrophysics
where we have defined the shear matrix Γ. Notice that the symmetries
(13.14) imply that C is symmetric,
Thus, Γ is also symmetric and has only the two independent components
γ1 = C̄αβγδ k̃β k̃γ E1α E1δ − E2α E2δ , γ2 = C̄αβγδ k̃β k̃γ E1α E2δ . (13.23)
where ϕ is the polar angle on the screen spanned by the vectors E1,2 .
Before we apply the optical tidal matrix, we rotate it into its principal-
axis frame,
κ + γ1 γ2 κ+γ 0
→ (13.27)
γ2 κ − γ1 0 κ−γ
with γ2 = γ12 + γ22 , which shows that it maps the circle onto a curve
outlined by the vector
x (κ + γ) cos ϕ
≡ . (13.28)
y (κ − γ) sin ϕ
x2 y2
+ = cos2 ϕ + sin2 ϕ = 1 . (13.29)
(κ + γ)2 (κ − γ)2
Thus, for γ = 0, the originally circular cross section remains circular,
with κ being responsible for isotropically expanding or shrinking it, while
the light bundle is elliptically deformed if γ 0.
13.2 Gravitational lensing 203
∇2k̃ D = κD . (13.31)
1 4πG
κ = − G k̃, k̃ = − 4 T k̃, k̃ , (13.33)
2 c ?
Why does the cosmological con-
using Einstein’s field equations in the second step.
stant Λ not appear in the expres-
Next, we can insert the energy-momentum tensor (12.53) for a perfect sion (13.33) for the convergence?
fluid, ) p*
T = ρ + 2 u ⊗ u + pg , (13.34)
c
and use the fact that fundamental observers (i.e. observers for whom the
universe appears isotropic) have u = c∂t and u = −cdt.
The frequency of the light measured by a hypothetical fundamental
observer moving with four-velocity u and placed between the source
and the final observer is k, u. Due to our definition of k̃ = k/ωo , and
because of the cosmological redshift (12.73), we can write
2 3 k, u ω
k̃, u = = =1+z, (13.35)
ωo ωo
where ωo is the frequency measured by the final observer, and z is the
redshift relative to the final observer.
Thus, we find
If p ρc2 , the density scales like a−3 as shown in (12.62), and then
4πG −5
κ=− ρ0 a . (13.37)
c2
D = κD , (13.42)
where the prime indicates the derivative with respect to the affine param-
eter λ.
Equation (13.42) can be further simplified to reveal its very intuitive
meaning. From the metric in the form (12.83), we see that radially
propagating light rays must satisfy
dλ = cadt = a2 dw . (13.44)
D/a, i.e. the diameter with the expansion of the universe divided out.
Substituting dw for dλ, we first see that
& ) D *'
d2 ) D * 2 d 2 d d
=a a = a2 (aD − a D)
dw2 a dλ dλ a dλ
= a2 (aD − a D) . (13.45)
da ȧ da ȧ d ) ȧ * 1 d ) ȧ *2
a = = = 2 = 2 , (13.46)
dλ ca da c a da a 2c da a
which enables us to insert Friedmann’s equation (12.55) in the form
) ȧ *2 8πG ρ0 Λc2 kc2
= + − 2 (13.47)
a 3 a3 3 a
to find
da 4πG
a = = − 2 ρ0 a−4 + ka−3 = κa + ka−3 , (13.48)
dλ c
inserting κ from (13.37).
Finally, we substitute D = κD from (13.31) and a from (13.48) into
(13.45) and obtain an intuitive result.
Propagation equation for the bundle diameter
In a spatially homogeneous and isotropic spacetime, the comoving ?
diameter D of a light bundle obeys the equation The derivation of the behaviour
of light bundles from Fried-
d2 ) D * mann’s equation suggests that
2
= a3 κD − a2 D κa + ka−3
dw a it should be possible to derive
)D*
= −k , (13.49) Friedmann’s equation from the
a behaviour of light bundles. Is it?
which is a simple oscillator equation.
Equation (13.49) is now easily solved. We set the boundary conditions
such that the bundle emerges from a source point, hence D = 0 at the
source, and that it initially expands linearly with the radial distance w,
hence d(D/a)/dw = 1 there. Then, the solution of (13.49) is
⎧
⎪
⎪
⎪ k−1/2 sin k1/2 w (k > 0)
⎪
⎪
⎨
D = a fk (w) = a ⎪⎪ w (k = 0) , (13.50)
⎪
⎪
⎪
⎩|k|−1/2 sinh |k|1/2 w (k < 0)
∇·T =0 (13.54)
because the only non-vanishing of the connection forms ωα0 for a static,
axially symmetric spacetime is
With the components of the Einstein tensor given in (8.60) and the
energy-momentum tensor (13.52), the two independent field equations
read
1 −2b 1 2b 8πG
− 2 +e − =− 2 ρ
r r2 r c
1 1 2a 8πG
− 2 + e−2b 2 + = 4 p. (13.59)
r r r c
The first of these equations is equivalent to
8πG
re−2b = 1 − 2 ρr2 . (13.60)
c
Integrating, and using the mass
r
M(r) = 4π ρ(r )r2 dr , (13.61)
0
If we subtract the first from the second field equation (13.59), we find
2e−2b 8πG
(a + b ) = 4 (ρc2 + p) (13.63)
r c
208 13 Relativistic Astrophysics
or
4πG 2b 2
a = −b + e (ρc + p)r . (13.64)
c4
On the other hand, (13.62) gives
2m 2m 2m 8πG
− 2b e−2b = − = 2 − 2 ρr , (13.65)
r2 r r c
or
4πG m 2b
b = ρr − e , (13.66)
c2 r2
which allows us to write (13.64) as
m 4πG m + 4πGpr3 /c4
a = 2 + 4 pr e2b = . (13.67)
r c r(r − 2m)
Tolman-Oppenheimer-Volkoff equation
But the hydrostatic equation demands (13.58), which we combine with
(13.67) to find
Neutron stars are a possible end product of the evolution of massive stars.
When such stars explode as supernovae, they may leave behind an object
with a density so high that protons and electrons combine to neutrons
in the process of inverse β decay. Objects thus form which consist of
matter with nuclear density
ρ0 ≈ 5 · 1014 g cm−3 . (13.70)
(13.75).
0.25
0.2
0.15
x* = 5/6
0.1
0.05
0
0 0.2 0.4 0.6 0.8 1
dimension-less radius x = r/0
Instead of a postface
If these lectures lay the foundation for studying more detailed textbooks
and reading the research literature, they serve their intended purpose.
Appendix A
Electrodynamics
For α = 0, (β, γ) = (1, 2), (1, 3) and (2, 3), this gives
˙ + c∇
B × E = 0 , (A.7)
213
214 A Electrodynamics
·B
∇ =0. (A.8)
4π μ
∂ν F μν = j , (A.9)
c
where
ρc
jμ = j (A.10)
· E = 4πρ ,
∇ c∇ − E˙ = 4πj ,
×B (A.11)
respectively.
With the definition (A.1) and the Lorenz gauge condition ∂μ Aμ = 0, the
inhomogeneous equations (A.9) can be written as
4π μ
Aμ = − j , (A.12)
c
i.e. by
1
Aμ (t, x ) = d3 x dtG(t, t , x, x ) j μ (t, x ) . (A.14)
c
1 μν 1
L=− F Fμν − Aμ j μ , (A.15)
16π c
from which Maxwell’s equations follow by the Euler-Lagrange equa-
tions,
∂L ∂L
∂ν ν μ − μ = 0 . (A.16)
∂(∂ A ) ∂A
A.3 Lagrange density and energy-momentum tensor 215
F αβ Fαβ = −2(E 2 − B
2) , (A.19)
2
E 2 + B
T 00 = (A.21)
8π
of the electromagnetic field and the Poynting vector
c
cT 0i = E×B. (A.22)
4π
Appendix B
Summary of Differential
Geometry
B.1 Manifold
v : Fp → R , f → v( f ) . (B.2)
217
218 B Differential Geometry
v = vi ∂i , v( f ) = vi ∂i f . (B.4)
w : TM → R , v → w(v) . (B.5)
The space of dual vectors to a tangent vector space T M is the dual space
T ∗ M.
Specifically, the differential of a function f ∈ F is a dual vector defined
by
d f : T M → R , v → d f (v) = v( f ) . (B.6)
Accordingly, the differentials of the coordinate functions xi form a basis
{dxi } of the dual space which is orthonormal to the coordinate basis {∂i }
of the tangent space,
B.3 Tensors
t = ti j dxi ⊗ dx j . (B.10)
C : T sr → T s−1
r−1
, t → Ct (B.12)
B.4 Covariant derivative 219
such that one of the dual vector arguments and one of the vector argu-
ments are filled with pairs of basis elements and summed over all pairs.
For example, the contraction of a tensor t ∈ T11 is
Ct = t(dxk , ∂k ) = (tij ∂i ⊗ dx j )(dxk , ∂k ) = tij δki δkj = tkk . (B.13)
∇γ̇ v = 0 . (B.24)
R̄ : T M × T M × T M → T M ,
(x, y, z) → R̄(x, y)z = ∇ x ∇y − ∇y ∇ x − ∇[x,y] z . (B.29)
R̄ijkl = dxi [R̄(∂k , ∂l )∂ j ] = ∂k Γijl − ∂l Γijk + Γajl Γiak − Γajk Γial . (B.31)
B.7 Pull-back, Lie derivative, Killing fields 221
where the sums extend over all cyclic permutations of x, y, z. The first
Bianchi identity reduces the number of independent components of the
Riemann tensor to 20. In components, the second Bianchi identity can
be written
R̄ij[kl;m] = 0 , (B.34)
where the indices in brackets need to be antisymmetrised.
The Ricci tensor is the contraction of the Riemann tensor over its first
and third indices, thus its components are
R jl = R̄ijil = Rl j . (B.35)
R = Rii . (B.36)
As for vectors and dual vectors, the pull-back and the push-forward can
also be defined for tensors of arbitrary rank.
The Lie derivative of a tensor field T into direction v is given by the limit
φ∗t T − T
Lv T = lim , (B.42)
t→0 t
where φt is the flow of v. The Lie derivative quantifies how a tensor
changes as the manifold is transformed by the flow of a vector field.
The Lie derivative is linear and obeys the Leibniz rule,
L x (y + z) = L x y + L x z , L x (y ⊗ z) = L x y ⊗ z + y ⊗ L x z . (B.43)
Lv f = v( f ) = d f (v) . (B.45)
Lv d f = dLv f . (B.46)
Lv x = [v, x] . (B.47)
By its commutation with contractions and the Leibniz rule, the Lie
derivative of a dual vector w turns out to be
with v1,2 ∈ T M.
Killing vector fields K define isometries of the metric, i.e. the metric
does not change under the flow of K. This implies the Killing equation
LK g = 0 ⇒ ∇i K j + ∇ j Ki = 0 . (B.50)
B.8 Differential forms 223
Let {ei } be an arbitrary basis and {θi } its dual basis such that
dgi j = ωi j + ω ji . (B.64)
or
-2
Torsion and curvature are expressed by the torsion 2-form Θi ∈ and
-
the curvature 2-form Ωij ∈ 2 as
The torsion and curvature forms are related to the connection forms and
the dual basis vectors by Cartan’s structure equations
Penrose-Carter diagrams
With
dU dV sin(U − V)
dũ = , dṽ = , ṽ − ũ = , (C.4)
cos2 U cos2 V cos U cos V
the line elements turns into
& '
1 1 2
ds = − 2
2
dUdV + sin (V − U) dΩ .
2
(C.5)
cos U cos2 V 4
227
228 C Penrose-Carter diagrams
Using
1 1
U = (T − R) , V = (T + R) , (C.7)
2 2
we find
1 2
dUdV = dT − dR2 (C.8)
4
and thus the line element
ds2 = ω−2 (T, R) −dT 2 + dR2 + sin2 R dΩ2 (C.9)
i+
J+
T
R i0
J−
i−
Figure C.1 Penrose-Carter diagram of Minkowski spacetime. The light-
gray curves are lines of constant radius (running from i− to i+ ) and lines of
constant time (emerging from i0 ).
The following points and lines are particularly important for the causal
structure of the spacetime:
229
i+ r=0 i+
J+ J+
III
II I
i0 i0
IV
J− J−
i− r=0 i−
231
232 INDEX
tensor field, 25
Tolman-Oppenheimer-Volkoff
equation, 208
torsion, 38
forms, 112
tensor, 39
in arbitrary basis, 114
two-sphere, 16
atlas, 16
weakness of gravity, 4
Weyl tensor, 200
Einstein‘s theory of general relativity is still the valid theory of gravity
and has been confirmed by numerous tests and measurements. It
is built upon simple principles and relates the geometry of space-
time to its matter-energy content. These lecture notes begin by
introducing the physical principles and by preparing the necessary
mathematical tools taken from differential geometry. Beginning with
Einstein’s field equations, which are introduced in two different
ways in the lecture, the motion of test particles in a gravitational
field is then discussed, and it is shown how the properties of weak
gravitational fields follow from the field equations. Solutions for com-
pact objects and black holes are derived and discussed as well as
cosmological models. Two applications of general relativity to astro-
physics conclude the lecture notes.
ISBN 978-3-947732-60-9
9 783947 732609