A Brief Introduction To Semi-Riemannian Geometry and General Relativity
A Brief Introduction To Semi-Riemannian Geometry and General Relativity
general relativity
Hans Ringström
May 5, 2015
2
Contents
2 Semi-Riemannian manifolds 7
2.1 Semi-Riemannian metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Pullback, isometries and musical isomorphisms . . . . . . . . . . . . . . . . . . . . 8
2.3 Causal notions in Lorentz geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Warped product metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.5 Existence of metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.6 Riemannian distance function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.7 Relevance of the Euclidean and the Minkowski metrics . . . . . . . . . . . . . . . . 13
3 Levi-Civita connection 15
3.1 The Levi-Civita connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Parallel translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.4 Variational characterization of geodesics . . . . . . . . . . . . . . . . . . . . . . . . 22
4 Curvature 25
4.1 The curvature tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2 Calculating the curvature tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.3 The Ricci tensor and scalar curvature . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4 The divergence, the gradient and the Laplacian . . . . . . . . . . . . . . . . . . . . 29
4.5 Computing the covariant derivative of tensor fields . . . . . . . . . . . . . . . . . . 29
4.5.1 Divergence of a covariant 2-tensor field . . . . . . . . . . . . . . . . . . . . . 30
4.6 An example of a curvature calculation . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.6.1 Computing the connection coefficients . . . . . . . . . . . . . . . . . . . . . 32
4.6.2 Calculating the components of the Ricci tensor . . . . . . . . . . . . . . . . 33
4.7 The 2-sphere and hyperbolic space . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
i
ii CONTENTS
The two basic examples are the Euclidean scalar product and the Minkowski scalar product.
Example 3. The Euclidean scalar product on Rn , 1 ≤ n ∈ Z, here denoted gEucl , is defined as
follows. If v = (v 1 , . . . , v n ) and w = (w1 , . . . , wn ) are two elements of Rn , then
n
X
gEucl (v, w) = v i wi .
i=1
The vector space Rn equipped with the Euclidean scalar product is called the (n-dimensional)
Euclidean scalar product space. The Minkowski scalar product on Rn+1 , 1 ≤ n ∈ Z, here denoted
gMin , is defined as follows. If v = (v 0 , v 1 , . . . , v n ) and w = (w0 , w1 , . . . , wn ) are two elements of
Rn+1 , then
Xn
gMin (v, w) = −v 0 w0 + v i wi .
i=1
The vector space Rn+1 equipped with the Minkowski scalar product is called the (n+1-dimensional)
Minkowski scalar product space.
In order to distinguish between different scalar products, it is convenient to introduce the notion
of an index.
1
2 CHAPTER 1. SCALAR PRODUCT SPACES
Definition 4. Let (V, g) be a scalar product space. Then the index, say ι, of g is the largest
integer that is the dimension of a subspace W ⊆ V on which g is negative definite.
As in the case of Euclidean geometry, it is in many contexts convenient to use particular bases,
such as an orthonormal basis; in other words, a basis {ei } such that g(ei , ej ) = 0 for i 6= j and
g(ei , ei ) = ±1 (no summation on i).
Lemma 5. Let (V, g) be a scalar product space. Then there is an integer d ≤ n := dim V and a
basis {ei }, i = 1, . . . , n, of V such that
• g(ei , ej ) = 0 if i 6= j.
• g(ei , ei ) = −1 if i ≤ d.
• g(ei , ei ) = 1 if i > d.
Proof. Let {vi } be a basis for V and let gij = g(vi , vj ). If G is the matrix with components gij ,
then G is a symmetric matrix. There is thus an orthogonal matrix T so that T GT t is diagonal. If
Tij are the components of T , then the ij’th component of T GT t is given by
!
X X X X
Tik Gkl Tjl = Tik g(vk , vl )Tjl = g Tik vk , Tjl vl .
k,l k,l k l
it thus follows that g(wi , wj ) = 0 if i 6= j. Due to the non-degeneracy of the scalar product,
g(wi , wi ) 6= 0. We can thus define a basis {Ei } according to
1
Ei = wi .
|g(wi , wi )|1/2
Then g(Ei , Ei ) = ±1. By renumbering the Ei , one obtains a basis with the properties stated in
the lemma.
If g is definite, the last statement of the lemma is trivial. Let us therefore assume that 0 < d < n.
Clearly, the index ι of g satisfies ι ≥ d. In order to prove the opposite inequality, let W be a
subspace of V such that g is negative definite on W and such that dim W = ι. Let N be the
subspace of V spanned by {ei }, i = 1, . . . , d, and ϕ : W → N be the map defined by
d
X
ϕ(w) = − g(w, ei )ei .
i=1
this equality is a consequence of the fact that if we take the scalar product of ei with the left hand
side minus the right hand side, then the result is zero for all i (so that non-degeneracy implies
that (1.1) holds). If ϕ(w) = 0, we thus have
n
X
w= g(w, ei )ei .
i=d+1
1.2. ORTHONORMAL BASES ADAPTED TO SUBSPACES 3
Compute
n
X n
X
g(w, w) = g(w, ei )g(w, ej )g(ei , ej ) = g(w, ei )2 ≥ 0.
i,j=d+1 i=d+1
Since g is negative definite on W , this implies that w = 0. Thus ϕ is injective, and the lemma
follows.
If V is an n-dimensional Riemannian scalar product space, then there is a linear isometry from
V to the n-dimensional Euclidean scalar product space. If V is an n + 1-dimensional Lorentz
scalar product space, then there is a linear isometry from V to the n + 1-dimensional Minkowski
scalar product space. Due to this fact, and the fact that the reader is assumed to be familiar with
Euclidean geometry, we here focus on the Lorentz setting.
In order to understand Lorentz scalar product spaces better, it is convenient to make a few more
observations of a linear algebra nature. To begin with, if (V, g) is a scalar product space and W
is a subspace of V , then
W ⊥ = {v ∈ V : g(v, w) = 0 ∀w ∈ W }.
In contrast with the Riemannian setting, W + W ⊥ does not equal V in general.
Exercise 9. Give an example of a Lorentz scalar product space (V, g) and a subspace W of V
such that W + W ⊥ 6= V .
Even though W + W ⊥ does not in general equal V , it is of interest to find conditions on W such
that the relation holds. One such condition is the following.
Definition 13. Let W be a subspace of a scalar product space V . Then W is said to be non-
degenerate if g|W is non-degenerate.
Lemma 14. Let W be a subspace of a scalar product space V . Then W is non-degenerate if and
only if V = W + W ⊥ .
Proof. Due to Lemma 10 and (1.2), it is clear that W + W ⊥ = V if and only if W ∩ W ⊥ = {0}.
However, W ∩ W ⊥ = {0} is equivalent to W being non-degenerate.
Corollary 15. Let W1 be a subspace of a scalar product space (V, g). If W1 is non-degenerate,
then W2 = W1⊥ is also non-degenerate. Thus Wi , i = 1, 2, are scalar product spaces with indices
ιi ; the scalar product on Wi is given by gi = g|Wi . If ι is the index of V , then ι = ι1 +ι2 . Moreover,
there is an orthonormal basis {ei }, i = 1, . . . , n, of V which is adapted to W1 and W2 in the sense
that {ei }, i = 1, . . . , d, is a basis for W1 and {ei }, i = d + 1, . . . , n, is a basis for W2 .
Proof. Since W1 is non-degenerate and W2⊥ = W1 (according to Lemma 10), Lemma 14 implies
that
V = W1 + W2 = W2⊥ + W2 .
Applying Lemma 14 again implies that W2 is non-degenerate. Defining gi as in the statement of
the corollary, it is clear that (Wi , gi ), i = 1, 2, are scalar product spaces. Due to Lemma 5, we
know that each of these scalar product spaces have an orthonormal basis. Let {ei }, i = 1, . . . , d,
be an orthonormal basis for W1 and {ei }, i = d + 1, . . . , n, be an orthonormal basis for W2 .
Then {ei }, i = 1, . . . , n, is an orthonormal basis of V . Since ι1 equals the number of elements
of {ei }, i = 1, . . . , d, with squared norm equal to −1, and similarly for ι2 and ι, it is clear that
ι = ι1 + ι2 .
Definition 16. Let (V, g) be a Lorentz scalar product space. Then a vector v ∈ V is said to be
The classification of a vector v ∈ V according to the above is called the causal character of the
vector v.
The importance of this terminology stems from its connection to the notion of causality in physics.
According to special relativity, no information can travel faster than light. Assuming γ to be a
curve in the Minkowski scalar product space (γ should be thought of as the trajectory of a physical
object; a particle, a spacecraft, light etc.), the speed of the corresponding object relative to that
of light is characterized by the causal character of γ̇ with respect to the Minkowski scalar product.
1.3. CAUSALITY FOR LORENTZ SCALAR PRODUCT SPACES 5
If γ̇ is timelike, the speed is strictly less than that of light, if γ̇ is lightlike, the speed equals that
of light.
In Minkowski space, if v = (v 0 , v̄) ∈ Rn+1 , where v̄ ∈ Rn , then
where |v̄| denotes the usual norm of an element v̄ ∈ Rn . Thus v is timelike if |v 0 | > |v̄|, lightlike
if |v 0 | = |v̄| =
6 0 and spacelike if |v 0 | < |v̄| or v = 0. The set of timelike vectors consists of two
components; the vectors with v 0 > |v̄| and the vectors with −v 0 > |v̄|. Choosing one of these
components corresponds to a choice of so-called time orientation (a choice of what is the future
and what is the past). Below we justify these statements and make the notion of a time orientation
more precise. However, to begin with, it is convenient to introduce some additional terminology.
Definition 17. Let (V, g) be a scalar product space and W ⊆ V be a subspace. Then W is said
to be spacelike if g|W is positive definite; i.e., if g|W is nondegenerate of index 0. Moreover, W is
said to be lightlike if g|W is degenerate. Finally, W is said to be timelike if g|W is nondegenerate
of index 1.
Let (V, g) be a Lorentz scalar product space. If u ∈ V is a timelike vector, the timecone of V
containing u, denoted C(u), is defined by
The opposite timecone is defined to be C(−u). Note that C(−u) = −C(u). If v ∈ V is timelike,
then v has to belong to C(u) or C(−u). The reason for this is that (Ru)⊥ is spacelike; cf.
Lemma 18. The following observation will be of importance in the discussion of the existence of
Lorentz metrics.
Lemma 20. Let (V, g) be a Lorentz scalar product space and v, w ∈ V be timelike vectors. Then
v and w are in the same timecone if and only if g(v, w) < 0.
Proof. Consider a timecone C(u) (where we, without loss of generality, can assume that u is a
unit timelike vector). Due to Corollary 15, there is an orthonormal basis {eα }, α = 0, . . . , n, of
V such that e0 = u. Then v ∈ C(u) if and only if v 0 > 0, where v = v α eα . Note also that if
x = xα eα , then x is timelike if and only if |x0 | > |x̄|, where x̄ = (x1 , . . . , xn ) and |x̄| denotes the
ordinary Euclidean norm of x̄ ∈ Rn .
Let v and w be timelike and define v α , wα , v̄ and w̄ in analogy with the above. Compute
where · denotes the ordinary dot product on Rn . Since v and w are timelike, |v 0 | > |v̄| and
|w0 | > |w̄|, so that
|v̄ · w̄| ≤ |v̄||w̄| < |v 0 w0 |.
Thus the first term on the right hand side of (1.3) is bigger in absolute value than the second
term. In particular, g(v, w) < 0 if and only if v 0 and w0 have the same sign.
Assume that v and w are in the same timecone; say C(u). Then v 0 , w0 > 0, so that g(v, w) < 0
by the above. Assume that g(v, w) < 0 and fix a timelike unit vector u. Then v 0 and w0 have the
same sign by the above. If both are positive, v, w ∈ C(u). If both are negative, v, w ∈ C(−u). In
particular, v, w are in the same timecone. The lemma follows.
6 CHAPTER 1. SCALAR PRODUCT SPACES
As a consequence of Lemma 20, timecones are convex; in fact, if 0 ≤ a, b ∈ R are not both zero and
v, w ∈ V are in the same timecone, then av + bw is timelike and in the same timecone as v and w.
In particular, it is clear that the timelike vectors can be divided into two components. A choice
of time orientation of a Lorentz scalar product space is a choice of timecone, say C(u). A Lorentz
scalar product space with a time orientation is called a time oriented Lorentz scalar product space.
Given a choice of time orientation, the timelike vectors belonging to the corresponding timecone
are said to be future oriented. Let v be a null vector and C(u) be a timecone. Then g(v, u) 6= 0.
If g(v, u) < 0, then v is said to be future oriented, and if g(v, u) > 0, then v is said to be past
oriented.
Chapter 2
Semi-Riemannian manifolds
The main purpose of the present chapter is to define the notion of a semi-Riemannian manifold
and to describe some of the basic properties of such manifolds.
Let (M, g) be a semi-Riemannian manifold. If (xi ) are local coordinates, the corresponding com-
ponents of g are given by
gij = g(∂xi , ∂xj ),
and g can be written
g = gij dxi ⊗ dxj .
Since gij are the components of a non-degenerate matrix, there is a matrix with components g ij
such that
g ij gjk = δki .
Note that the functions g ij are smooth, whenever they are defined. Moreover, g ij = g ji . In fact,
g ij are the components of a smooth, symmetric contravariant 2-tensor field. As will become clear,
this construction is of central importance in many contexts.
Again, the basic examples of metrics are the Euclidean metric and the Minkowski metric.
7
8 CHAPTER 2. SEMI-RIEMANNIAN MANIFOLDS
Definition 24. Let (xi ), i = 1, . . . , n, be the standard coordinates on Rn . Then the Euclidean
metric on Rn , denoted gE , is defined as follows. Let (v 1 , . . . , v n ), (w1 , . . . , wn ) ∈ Rn and
i ∂ n i ∂
∈ Tp Rn .
v=v ∈ Tp R , w = w
∂xi p ∂xi p
Then
n
X
gE (v, w) = v i wi .
i=1
Then
n
X
gM (v, w) = −v 0 w0 + v i wi .
i=1
Exercise 25. Give an example of a smooth manifold M , a Lorentz manifold (N, h) and a smooth
immersion F : M → N such that F ∗ h is not a semi-Riemannian metric on M .
Example 27. Let Sn ⊂ Rn+1 denote the n-sphere and ιSn : Sn → Rn+1 the corresponding
inclusion. Then the round metric on Sn , gSn , is defined by gSn = ι∗Sn gE ; cf. Definition 24. Let
H n denote the set of x ∈ Rn+1 such that gMin (x, x) = −1 and let ιH n : H n → Rn+1 denote the
corresponding inclusion. Then the hyperbolic metric on H n , gH n , is defined by gH n = ι∗H n gM ; cf.
Definition 24.
Remark 28. Both (Sn , gSn ) and (H n , gH n ) are Riemannian manifolds (we shall not demonstrate
this fact in these notes; the interested reader is referred to, e.g., [1, Chapter 4] for a more detailed
discussion). Note that there is a certain symmetry in the definitions: Sn is the set of x ∈ Rn+1
such that gEucl (x, x) = 1 and H n is the set of x ∈ Rn+1 such that gMin (x, x) = −1.
2.2. PULLBACK, ISOMETRIES AND MUSICAL ISOMORPHISMS 9
assuming that this integral makes sense. If S is an oriented submanifold of Rn , the volume of S
is then defined to be the volume of the oriented Riemannian manifold (S, ι∗ gE ).
A fundamental notion in semi-Riemannian geometry is that of an isometry.
It will be useful to keep in mind that a semi-Riemannian metric induces an isomorphism between
the sections of the tangent bundle and the sections of the cotangent bundle.
Remark 32. Here X∗ (M ) denotes the smooth sections of the cotangent bundle; i.e., the one-forms.
Proof. Note that, given X ∈ X(M ), it is clear that X [ is linear over C ∞ (M ). Due to the tensor
characterization lemma, [2, Lemma 12.24, p. 318], is is thus clear that X [ ∈ X∗ (M ). In addition,
it is clear that the map taking X to X [ is linear over C ∞ (M ).
In order to prove injectivity of the map, assume that X [ = 0. Then g(X, Y ) = 0 for all Y ∈ X(M ).
In particular, given p ∈ M , g(Xp , v) = 0 for all v ∈ Tp M . Due to the non-degeneracy of the metric,
this implies that Xp = 0 for all p ∈ M . Thus X = 0 and the map is injective.
In order to prove surjectivity, let η ∈ X∗ (M ). To begin with, let us try to find a vectorfield X on
a coordinate neighbourhood U such that X [ = η on U . If ηi are the components of η with respect
to local coordinates, then we can define a vectorfield on U by
∂
X = g ij ηj .
∂xi
In this expression, g ij are the components of the inverse of the matrix with components gij . Then
X is a smooth vectorfield on U . Moreover,
Thus X [ = η. Due to the uniqueness, the local vectorfields can be combined to give an X ∈ X(M )
such that X [ = η. This proves surjectivity.
10 CHAPTER 2. SEMI-RIEMANNIAN MANIFOLDS
The maps
[ : X(M ) → X∗ (M ), ] : X∗ (M ) → X(M )
are sometimes referred to as musical isomorphisms. In the physics literature, where authors prefer
to write everything in coordinates, the maps ] and [ are referred to as raising and lowering indices
using the metric; if ηi are the components of a one-form, then g ij ηj are the components of the
corresponding vectorfield; if X i are the components of a vectorfield, then gij X j are the components
of the corresponding one-form. However, the musical isomorphisms are just a special case of a
general construction. If A is a tensor field of mixed (k, l)-type, then we can, for example, lower
one of the indices of A according to
gii1 Aji11···i
···jl .
k
The result defines a tensor field of mixed (k − 1, l + 1)-type. Again, this is just a special case of
a construction called contraction. The idea is the following. If Aij11···i
···jl are the components of a
k
tensor field of mixed (k, l)-type with respect to local coordinates, then setting ir = js = i and
summing over i yields a tensor field of mixed (k − 1, l − 1)-type. For example g i1 i2 ηj1 are the
components of a tensor field of mixed (2, 1)-type. Applying the contraction construction to i2 and
j1 yields the components of η ] .
Exercise 33. Let A be a tensor field of mixed (2, 2)-type. The components of A with respect to
local coordinates are Aji11ij22 . Prove that Aij11ii (where Einstein’s summation convention is enforced)
are the components of a tensor field of mixed (1, 1)-type.
Definition 34. Let (M, g) be a Lorentz manifold. A time orientation of (M, g) is a choice of time
orientation of each scalar product space (Tp M, gp ), p ∈ M , such that the following holds. For each
p ∈ M , there is an open neighbourhood U of p and a smooth vectorfield X on U such that Xq is
future oriented for all q ∈ U . A Lorentz metric g on a manifold M is said to be time orientable if
(M, g) has a time orientation. A Lorentz manifold (M, g) is said to be time orientable if (M, g) has
a time orientation. A Lorentz manifold with a time orientation is called a time oriented Lorentz
manifold.
Remark 35. Here gp denotes the scalar product induced on Tp M by g. The requirement that
there be a local vectorfield with the properties stated in the definition is there to ensure the
“continuity” of the choice of time orientation.
A choice of time orientation for a Lorentz manifold corresponds to a choice of which time direction
corresponds to the future and which time direction corresponds to the past. In physics, time
oriented Lorentz manifolds are of greater interest than non-time oriented ones. For this reason,
the following terminology is sometimes introduced.
Let us now introduce some of the notions of causality that we shall use.
In case (M, g) is a spacetime, it is also possible to speak of future directed timelike vectors etc.
Note, however, that a causal curve is said to be future directed if and only if γ̇(t) is future oriented
for all t in the domain of definition of γ. Our requirements concerning vector fields is similar.
g = π1∗ g1 + (f ◦ π1 )2 π2∗ g2 .
One special case of this construction is obtained by demanding that f = 1. In that case, the
resulting warped product is referred to as a semi-Riemannian product manifold. One basic example
of a warped product is the following.
where t is the coordinate on the first factor in I × R3 and xi , i = 1, 2, 3, are the coordinates on
the last three factors. The geometry of most models of the universe used by physicists today are
of the type (M, g). What varies from model to model is the function f .
Proposition 41. Every smooth manifold with or without boundary admits a Riemannian metric.
Proof. Assume that there is a nowhere vanishing smooth vectorfield X on M . Let h be a Rieman-
nian metric on M (such a metric exists due to Proposition 41). By normalizing X if necessary, we
can assume h(X, X) = 1. Define g according to
g = −2X [ ⊗ X [ + h,
where X [ is defined in Lemma 31. Then
g(X, X) = −2[X [ (X)]2 + 1 = −1.
Given p ∈ M , let e2 |p , . . . , en |p ∈ Tp M be such that e1 |p , . . . , en |p is an orthonormal basis for
(Tp M, hp ), where e1 |p = Xp . Then {ei |p } is an orthonormal basis for (Tp M, gp ). Moreover, it is
clear that the index of gp is 1. Thus (M, g) is a Lorentz manifold. Since we can define a time
orientation by requiring Xp to be future oriented for all p ∈ M , it is clear that M admits a time
orientable Lorentz metric.
Assume now that M admits a time orientable Lorentz metric g. Fix a time orientation. Let {Uα }
be an open covering of M such that on each Uα there is a timelike vector field Xα which is future
pointing (that such a covering exists is a consequence of the definition of a time orientation; cf.
Definition 34). Let {φα } be a partition of unity subordinate to the covering {Uα }. Define X by
X
X= φα Xα .
α
Fix a p ∈ M . At this point, the sum consists of finitely many terms, so that
k
X
Xp = ai Xi,p ,
i=1
where 0 < ai ∈ R and Xi,p ∈ Tp M are future oriented timelike vectors. Due to Lemma 20 it is
then clear that Xp is a future oriented timelike vector. In particular, X is thus a future oriented
timelike vectorfield. Since such a vectorfield is nowhere vanishing, it is clear that M admits a
non-zero vector field.
It is of course natural to ask what happens if we drop the condition that (M, g) be time orientable.
However, in that case there is a double cover which is time orientable (for those unfamiliar with
covering spaces, we shall not make any use of this fact). It is important to note that the existence
of a Lorentz metric is a topological restriction; not all manifolds admit Lorentz metrics. As an
orientation in the subject of Lorentz geometry, it is also of interest to make the following remark
(we shall not make any use of the statements made in the remark in what follows).
Remark 43. If (M, g) is a spacetime such that M is a closed manifold (in other words, M is
compact and without boundary), then there is a closed timelike curve in M . In other words,
there is a future oriented timelike curve γ in M such that γ(t1 ) = γ(t2 ) for some t1 < t2 in
the domain of definition of γ. This means that it is possible to travel into the past. Since this
is not very natural in physics, spacetimes (M, g) such that M is closed are not very natural
(in contrast with the Riemannian setting). For a proof of this statement, see [1, Lemma 10,
p. 407]. In general relativity, one often requires spacetimes to satisfy an additional requirement
called global hyperbolicity (which we shall not define here) which involves additional conditions
concerning causality. Moreover, globally hyperbolic spacetimes (M, g), where n + 1 = dim M , are
topologically products M = R × Σ where Σ is an n-dimensional manifold.
with (M, g). Since the basic properties of the Riemannian distance function are described in [2,
pp. 337–341], we shall not do so here.
t0 =t + t0 , (2.1)
x0 =Ax + vt + x0 . (2.2)
The special relativistic perspective. The classical laws of physics transform well under changes
of coordinates of the form (2.1)–(2.2). However, it turns out that Maxwell’s equations do not.
This led Einstein to use a different starting point, namely that the speed of light is the same in all
intertial frames. One consequence of this assumption is that time is no longer absolute. Moreover,
if one wishes to compute the associated changes of coordinates when going from one intertial
frame to another, they are different from (2.1)–(2.2). The group of transformations (taking the
coordinates of one intertial frame to the coordinates of another frame) that arise when taking
this perspective is called the group of Lorentz transformations. The main point of introducing
Minkowski space is that the group of isometries of Minkowski space are exactly the group of
Lorentz transformations.
14 CHAPTER 2. SEMI-RIEMANNIAN MANIFOLDS
Chapter 3
Einstein’s equations of general relativity relate the curvature of a spacetime with the matter
content of the spacetime. In order to understand this equation, it is therefore important to
understand the notion of curvature. This subject has a rich history, and here we only give a quite
formal and brief introduction to it. One way to define curvature is to examine how a vector is
changed when parallel translating it along a closed curve in the manifold. In order for this to
make sense, it is of course necessary to assign a meaning to the notion of “parallel translation”. In
the case of Euclidean space, the notion is perhaps intuitively clear; we simply fix the components
of the vector with respect to the standard coordinate frame and then change the base point.
Transporting a vector in Euclidean space (along a closed curve) in this way yields the identity
map; one returns to the vector one started with. This is one way to express that the curvature of
Euclidean space vanishes. Using a rather intuitive notion of parallel translation on the 2-sphere,
one can convince oneself that the same is not true of the 2-sphere.
In order to proceed to a formal development of the subject, it is necessary to clarify what is meant
by parallel translation. One natural way to proceed is to define an “infinitesimal” version of this
notion. This leads to the definition of a so-called connection.
• ∇X Y is linear over C ∞ (M ) in X,
• ∇X Y is linear over R in X,
• ∇X (f Y ) = X(f )Y + f ∇X Y for all X, Y ∈ X(M ) and all f ∈ C ∞ (M ).
The expression ∇X Y is referred to as the covariant derivative of Y with respect to X for the
connection ∇.
15
16 CHAPTER 3. LEVI-CIVITA CONNECTION
∇X (a1 Y1 + a2 Y2 ) = a1 ∇X Y1 + a2 ∇X Y2
Let (M, g) be a semi-Riemannian manifold. Our next goal is to argue that there is preferred
connection, given the metric g. However, in order to single out a preferred connection, we have to
impose additional conditions. One such condition would be to require that
for all X, Y, Z ∈ X(M ). In what follows, it is going to be a bit cumbersome to use the notation
g(X, Y ). We therefore define h·, ·i by
hX, Y i = g(X, Y );
we shall use this notation both for vectorfields and for individual vectors. With this notation,
(3.1) can be written
XhY, Zi = h∇X Y, Zi + hY, ∇X Zi. (3.2)
A connection satisfying this requirement is said to be metric. However, it turns out that the
condition (3.2) does not determine a unique connection. In fact, we are free to add further
conditions. One such condition would be to impose that ∇X Y − ∇Y X can be expressed in terms
of only X and Y , without any reference to the connection. Since ∇X Y − ∇Y X is antisymmetric,
one such condition would be to require that
∇X Y − ∇Y X = [X, Y ] (3.3)
for all X, Y ∈ X(M ). A connection satisfying this criterion is said to be torsion free. Remarkably,
it turns out that conditions (3.2) and (3.3) uniquely determine a connection, referred to as the
Levi-Civita connection.
Theorem 47. Let (M, g) be a semi-Riemannian manifold. Then there is a unique connection ∇
satisfying (3.2) and (3.3) for all X, Y, Z ∈ X(M ). It is called the Levi-Civita connection of (M, g).
Moreover, it is characterized by the Koszul formula:
2h∇X Y, Zi =XhY, Zi + Y hZ, Xi − ZhX, Y i − hX, [Y, Z]i + hY, [Z, X]i + hZ, [X, Y ]i. (3.4)
Proof. Assume that ∇ is a connection satisfying (3.2) and (3.3) for all X, Y, Z ∈ X(M ). Compute
where we have applied (3.2) and (3.3). In the fourth and fifth steps, we also rearranged the terms
and used the antisymmetry of the Lie bracket. Note that this equation implies that (3.4) holds.
This leads to the uniqueness of the Levi-Civita connection. The reason for this is the following.
Assume that ∇ and ∇ ˆ both satisfy (3.2) and (3.3). Then, due to the Koszul formula,
ˆ X Y, Zi = 0
h∇X Y − ∇
ˆ XY
∇X Y = ∇
for all X, Y ∈ X(M ). In other words, there is at most one connection satisfying the conditions
(3.2) and (3.3). Given X, Y ∈ X(M ), let θX,Y be defined by the condition that 2θX,Y (Z) is given
by the right hand side of (3.4). It can then be demonstrated that θX,Y is linear over C ∞ (M ); in
other words,
θX,Y (f1 Z1 + f2 Z2 ) = f1 θX,Y (Z1 ) + f2 θX,Y (Z2 )
for all X, Y, Zi ∈ X(M ), fi ∈ C ∞ (M ), i = 1, 2 (we leave it as an exercise to verify that this is
true). Due to the tensor characterization lemma, [2, Lemma 12.24, p. 318], it thus follows that
θX,Y is a one-form. By appealing to Lemma 31, we conclude that there is a smooth vectorfield
]
θX,Y such that
]
hθX,Y , Zi = θX,Y (Z),
where the right hand side is given by the right hand side of (3.4). We define ∇X Y by
]
∇X Y = θX,Y .
Then ∇ is a function from X(M ) × X(M ) to X(M ). However, it is not obvious that it satisfies the
conditions of Definition 44. Moreover, it is not obvious that it satisfies (3.2) and (3.3). In other
words, there are five conditions we need to verify. Let us verify the first condition in the definition
of a connection. Note, to this end, that 2h∇f X Y, Zi is given by the right hand side of (3.4), with
X replaced by f X. However, a straightforward calculation shows that if you replace X by f X in
(3.4), then you obtain f times the right hand side of (3.4). In other words,
h∇f X Y, Zi = hf ∇X Y, Zi.
2h∇X (f Y ), Zi =Xhf Y, Zi + f Y hZ, Xi − ZhX, f Y i − hX, [f Y, Z]i + hf Y, [Z, X]i + hZ, [X, f Y ]i
=X(f )hY, Zi + f XhY, Zi + f Y hZ, Xi − Z(f )hX, Y i − f ZhX, Y i
+ Z(f )hX, Y i − f hX, [Y, Z]i + f hY, [Z, X]i + X(f )hZ, Y i + f hZ, [X, Y ]i
=2f h∇X Y, Zi + 2hX(f )Y, Zi = 2hf ∇X Y + X(f )Y, Zi.
∇X (f Y ) = f ∇X Y + X(f )Y.
Thus the third condition of Definition 44 is satisfied. We leave it to the reader to verify that (3.2)
and (3.3) are satisfied.
Exercise 48. Prove that the connection ∇ constructed in the proof of Theorem 47 satisfies the
conditions (3.2) and (3.3) for all X, Y, Z ∈ X(M ).
18 CHAPTER 3. LEVI-CIVITA CONNECTION
Thus
∇X Y = X(Y k ) + Γkij X i Y j ∂k .
When defining parallel transport, we shall use the following consequence of this formula.
Lemma 49. Let (M, g) be a semi-Riemannian manifold and ∇ be the associated Levi-Civita
connection. Let v ∈ Tp M for some p ∈ M , and Y ∈ X(M ). Let Xi ∈ X(M ), i = 1, 2, be such that
Xi,p = v. Then
(∇X1 Y )p = (∇X2 Y )p .
for all X, Xi ∈ X(γ), ai ∈ R, f ∈ C ∞ (I) and Y ∈ X(M ), i = 1, 2. Moreover, this map has the
property that
d
hX1 , X2 i = hX10 , X2 i + hX1 , X20 i. (3.9)
dt
Remark 52. How to interpret the expression ∇γ(t) Y appearing in (3.8) is explained in Lemma 49
and Definition 50.
Proof. Let us begin by proving uniqueness. Let X ∈ X(γ). Then we can write X as
where the X i are smooth functions on X −1 (U ) (where U is the set on which the coordinates
(xi ) are defined). Assume now that we have derivative operator satisfying (3.6)–(3.8). Applying
(3.6)–(3.8) to (3.10) yields
dX i dX i
X 0 (t) = (t)∂i |γ(t) + X i (t)(∂i |γ )0 (t) = (t)∂i |γ(t) + X i (t)∇γ 0 (t) ∂i . (3.11)
dt dt
Since the right hand side only depends on the Levi-Civita connection, we conclude that uniqueness
holds.
In order to prove existence, we can define X 0 (t) by (3.11) for t ∈ X −1 (U ). It can then be verified
that the corresponding derivative operator satisfies the conditions (3.6)–(3.9); we leave this as an
exercise. Due to uniqueness, these coordinate representations can be patched together to produce
an element X 0 ∈ X(γ).
Exercise 53. Prove that the derivative operator defined by the formula (3.11) has the properties
(3.6)–(3.9).
It is of interest to write down a formula for X 0 in local coordinates. Let (xi ) be local coordinates,
γ i = xi ◦ γ and X i be defined by (3.10). Then, since
dγ i
γ 0 (t) = (t)∂i |γ(t) ,
dt
(3.11) implies
dX i dγ j
X 0 (t) = (t)∂i |γ(t) + X i (t) (t)Γkji [γ(t)]∂k |γ(t)
dt dt
(3.12)
dX k dγ j
= (t) + X i (t) (t)Γkji [γ(t)] ∂k |γ(t) .
dt dt
Given the derivative operator of Proposition 51, we are now in a position to assign a meaning to
the expression parallel translation used in the introduction to the present chapter.
Definition 54. Let (M, g) be a semi-Riemannian manifold, I ⊆ R be an open interval and
γ : I → M be a smooth curve. Then X ∈ X(γ) is said to be parallel along γ if and only if X 0 = 0.
Note that, in local coordinates, the equation X 0 = 0 is a linear equation for the components of
X; cf. (3.12). For this reason, we have the following proposition (cf. also the arguments used to
prove the existence of integral curves of vectorfields).
20 CHAPTER 3. LEVI-CIVITA CONNECTION
Due to Proposition 55 we are in a position to define parallel translation along a curve. Given
assumptions as in the statement of Proposition 55, let t0 , t1 ∈ I. Then there is a map
P : Tγ(t0 ) → Tγ(t1 )
defined as follows. Given ξ ∈ Tγ(t0 ) M , let X ∈ X(γ) be such that X 0 = 0 and X(t0 ) = ξ. Define
P (ξ) = X(t1 ). Here P depends (only) on γ, t0 and t1 . In some situations, it may be useful to
indicate this dependence, but if these objects are clear from the context, it is convenient to simply
write P . The map P is called parallel translation along γ from γ(t0 ) to γ(t1 ). Parallel translation
has the following property.
Proof. We leave it to the reader to prove that parallel translation is a vector space isomorphism. In
order to prove that it is an isometry, let v, w ∈ Tp0 M and V, W ∈ X(γ) be such that V 0 = W 0 = 0,
V (t0 ) = v and W (t0 ) = w. Then P (v) = V (t1 ) and P (w) = W (t1 ). Compute
Z t1
d
hP (v), P (w)i =hV (t1 ), W (t1 )i = hV (t0 ), W (t0 )i + hV, W idt
t0 dt
Z t1
=hv, wi + (hV 0 , W i + hV, W 0 i) dt = hv, wi,
t0
where we have used property (3.9) of the derivative operator 0 , as well as the fact that V 0 = W 0 = 0.
The proposition follows.
Let us analyze what parallel translation means in the case of Euclidean space and Minkowski
space. Let I and γ be as in the statement of Proposition 55, where (M, g) is either Euclidean
space or Minkowski space. Note that the Christoffel symbols of gE and gM vanish with respect to
standard coordinates on Rn and Rn+1 respectively. An element X ∈ X(γ) is therefore parallel if
and only if the components of X with respect to the standard coordinate vectorfields are constant
(just as we stated in the introduction). In particular, the result of the parallel translation does
not depend on the curve. It is of importance to note that, even though this is true in the case of
Euclidean space and Minkowski space, it is not true in general.
3.3 Geodesics
A notion which is extremely important both in Riemannian geometry and in Lorentz geometry is
that of a geodesic. In Riemannian geometry, geodesics are locally length minimizing curves. In
the case of general relativity (Lorentz geometry), geodesics are related to the trajectories of freely
falling test particles, as well as the trajectories of light.
Keeping (3.12) in mind, geodesics are curves which with respect to local coordinates (xi ) satisfy
the equation
γ̈ k + Γkij ◦ γ γ̇ i γ̇ j = 0,
(3.13)
where we use the notation
dγ i d2 γ i
γ i = xi ◦ γ, γ̇ i = , γ̈ i = .
dt dt2
It is important to note that, even though (3.13) is an ODE, it is (in contrast to the equation
X 0 = 0 for a fixed curve γ) a non-linear ODE. Due to the fact that (3.13) is an autonomous ODE
for γ and the fact that the Christoffel symbols are smooth functions, it is clear that geodesics are
smooth curves. Due to local existence and uniqueness results for ODE’s, we have the following
proposition.
Proposition 60. Let (M, g) be a semi-Riemannian manifold, p ∈ M and v ∈ Tp M . Then there
is a unique geodesic γ : I → M with the properties that
Proof. Since the uniqueness is clear from the definition, let us focus on existence.
Local existence and uniqueness. To begin with, note that there is an open interval I0 contain-
ing 0 and a unique geodesic β : I0 → M such that β 0 (0) = v; this is an immediate consequence of
applying standard results concerning ODE’s to the equation (3.13). In other words, local existence
and uniqueness holds.
Global uniqueness. In order to proceed, we need to prove global uniqueness. In other words,
we need to prove that if Ii , i = 0, 1, are open intervals containing 0 and βi : Ii → M are geodesics
such that βi0 (0) = v, then β0 = β1 on I0 ∩ I1 . In order to prove this statement, let A be the set of
t ∈ I0 ∩ I1 such that β00 (t) = β10 (t). Note that if t ∈ A, then β0 (t) = π ◦ β00 (t) = π ◦ β10 (t) = β1 (t),
where π : T M → M is the projection taking a tangent vector to its base point. In other words,
if we can prove that A = I0 ∩ I1 , it then follows that β0 = β1 on I0 ∩ I1 . Since 0 ∈ A, it is clear
that A is non-empty. Due to local uniqueness, A is open. In order to prove that A is closed, let
t1 ∈ I0 ∩ I1 belong to the closure of A. Then there is a sequence sj ∈ A such that sj → t1 . Since
βi0 : Ii → T M are smooth maps, it is clear that
β10 (t1 ) = lim β10 (sj ) = lim β00 (sj ) = β00 (t1 ).
j→∞ j→∞
Thus t1 ∈ A. Summing up, A is an open, closed and non-empty subset of I0 ∩I1 . Thus A = I0 ∩I1 .
In other words, global uniqueness holds.
Existence. Let Ia , a ∈ A, be the collection of open intervals Ia ⊆ R such that
• 0 ∈ Ia ,
• there is a geodesic γa : Ia → M such that γa0 (0) = v.
Due to local existence, we know that this collection of intervals is non-empty. Define
[
I= Ia .
a∈A
Then I ⊆ R is an open interval containing 0. Moreover, due to global uniqueness, we can define
a geodesic γ : I → M such that γ 0 (0) = v; simply let γ(t) = γa (t) for t ∈ Ia . Finally, it is clear,
by definition, that I is maximal.
22 CHAPTER 3. LEVI-CIVITA CONNECTION
The geodesic constructed in Proposition 60 is called the maximal geodesic with initial data given
by v ∈ Tp M .
Exercise 61. Prove that the maximal geodesics in Euclidean space and in Minkowski space are
the straight lines.
In general relativity, timelike geodesics are interpreted as the trajectories of freely falling test
particles and null geodesics are interpreted as the trajectories of light. In particular, in Lorentz
geometry, we can think of the timelike geodesics as freely falling observers. Moreover, if γ : I → M
is a future oriented timelike geodesic in a spacetime and t0 < t1 are elements of I, then
Z t1
1/2
(−hγ 0 (t), γ 0 (t)i) dt
t0
is the proper time between t0 and t1 as measured by the observer γ. Since the integrand is constant,
it is clear that if I = (t− , t+ ) and t+ < ∞, then the amount of proper time the observer can measure
to the future is finite (there is an analogous statement concerning the past if t− > −∞). This can
be thought of as saying that the observer leaves the spacetime in finite proper time. One way to
interpret this is that there is a singularity in the spacetime. It is therefore of interest to analyze
under what circumstances I 6= R. To begin with, let us introduce the following terminology.
Definition 64. Let (M, g) be a semi-Riemannian manifold and γ : I → M be a maximal geodesic
in (M, g). Then γ is said to be a complete geodesic if I = R. A semi-Riemannian manifold, all of
whose maximal geodesics are complete is said to be complete.
Euclidean space and Minkowski space are both examples of complete semi-Riemannian manifolds.
On the other hand, removing one single point from Euclidean space or Minkowski space yields
an incomplete semi-Riemannian manifold. In other words, the notion of completeness is very
sensitive. Moreover, it is clear that in order to interpret the presence of an incomplete causal
geodesic as the existence of a singularity (as is sometimes done), it is necessary to ensure that
the spacetime under consideration is maximal in some natural sense. Nevertheless, trying to sort
out conditions ensuring that a spacetime (which is maximal in some natural sense) is causally
geodesically incomplete is a fundamental problem. Due to the work of Hawking and Penrose,
spacetimes are causally geodesically incomplete under quite general circumstances. The relevant
results, which are known under the name of “the singularity theorems”, are discussed, e.g., in [1,
Chapter 14].
The function ν should be thought of as a variation of the curve γ(t) = ν(t, 0). Let
Z t1
L(s) = |h∂t ν(t, s), ∂t ν(t, s)i|1/2 dt.
t0
A natural question to ask is: what are the curves γ such that for every variation ν (as above,
with an appropriate degree of regularity and fixing the endpoints t0 and t1 ), L0 (0) = 0? Roughly
speaking, it turns out to be possible to characterize geodesics as the curves for which L0 (0) for all
such variations. In particular, geodesics in Riemannian geometry are the locally length minimizing
curves.
Here we shall not pursue this perspective further, but rather refer the interested reader to, e.g.,
[1, Chapter 10].
24 CHAPTER 3. LEVI-CIVITA CONNECTION
Chapter 4
Curvature
The notion of curvature arose over a long period of time. Some of the history can be found in [3].
Here we proceed in a more formal way. As indicated at the beginning of the previous chapter, one
way to define curvature is through parallel translation along a closed curve. Here we define the
curvature tensor via an “infinitesimal” version of this idea.
Proof. That R is linear over the real numbers is clear. The only thing we need to prove is thus
that
R(f X)Y Z =f RXY Z, (4.2)
RX(f Y ) Z =f RXY Z, (4.3)
RXY (f Z) =f RXY Z. (4.4)
We prove one of these equalities and leave the other two as exercises. Compute
RXY (f Z) =∇X [f ∇Y Z + Y (f )Z] − ∇Y [X(f )Z + f ∇X Z] − [X, Y ](f )Z − f ∇[X,Y ] Z
=X(f )∇Y Z + XY (f )Z + Y (f )∇X Z + f ∇X ∇Y Z − Y X(f )Z − X(f )∇Y Z
− Y (f )∇X Z − f ∇Y ∇X Z − [X, Y ](f )Z − f ∇[X,Y ] Z
=f RXY Z.
25
26 CHAPTER 4. CURVATURE
By an argument similar to the proof of the tensor characterization lemma, cf. [2, pp. 318–319],
Proposition 65 implies that it is possible to make sense of Rxy z for x, y, z ∈ Tp M . In fact, choosing
any vector fields X, Y, Z ∈ X(M ) such that Xp = x, Yp = y and Zp = z, we can define Rxy z by
hRxy z, zi = 0 (4.9)
for all x, y, z ∈ Tp M . In order to prove (4.9), compute (using (3.2) and the fact that [X, Y ] = 0)
hRXY Z, Zi =h∇X ∇Y Z − ∇Y ∇X Z, Zi
=Xh∇Y Z, Zi − h∇Y Z, ∇X Zi − Y h∇X Z, Zi + h∇X Z, ∇Y Zi
1 1 1
= XY hZ, Zi − Y XhZ, Zi = [X, Y ]hZ, Zi = 0.
2 2 2
due to (4.7). Adding up the four cyclic permutations of this equation and using (4.5) and (4.6)
yields (4.8). We leave the details to the reader.
∇ej ek = Γijk ei .
In case ei = ∂i , the connection coefficients are the Christoffel symbols given by (3.5). However,
for a general frame, the relation Γkij = Γkji does typically not hold. This is due to the fact that
the Lie bracket [ei , ej ] typically does not vanish. Note that the information concerning the Lie
k
brackets is contained in the functions γij defined by
k
[ei , ej ] = γij ek .
Let us compute
Rijkm = ∂j Γm m l m l m
ik − ∂i Γjk + Γik Γjl − Γjk Γil .
Moreover, in this case, the Γkij ’s are given by (3.5). With respect to the standard coordinates,
the Christoffel symbols of the Euclidean metric and the Minkowski metric vanish. In particular,
the associated curvature tensors thus vanish. Moreover, this property (essentially) characterizes
Euclidean space and Minkowski space. To prove this statement is, however, non-trivial.
The general strategy for computing the components of the curvature tensor is the following. First,
choose a suitable local frame. Which frame is most appropriate depends on the context. Sometimes
it is convenient to use a coordinate frame, but sometimes it is easier to carry out the computations
with respect to an orthonormal frame. Once a choice of frame has been made, one first calculates
k
the functions γij determined by the Lie bracket. Then, one calculates the coefficients Γkij using
the Koszul formula, (3.4). After this has been done, the components of the curvature can be
calculated using (4.12). Needless to say, this is a cumbersome process in most cases.
In terms of local coordinates, the components of the Ricci tensor are given by
Again, the Ricci tensor of Euclidean space and Minkowski space vanish. In what follows, we
denote the tensor field whose components are given by (4.13) by Ric. In other words, if Rijkm are
the components of the curvature tensor relative to a frame {ei }, then
Ric(ei , ek ) = Rijkj .
The Ricci tensor is an extremely important object in semi-Riemannian geometry in general, and
in general relativity in particular. It is of interest to derive alternate formulae for the Ricci tensor.
Lemma 70. Let (M, g) be a semi-Riemannian manifold and {ei } be an orthonormal frame such
that hei , ei i = i (no summation on i). Then
X
Ric(X, Y ) = j hRej X Y, ej i. (4.14)
j
where all the components are calculated with respect to the frame {ei }. Assume now that the
frame is orthonormal so that hei , ej i = 0 if i 6= j and hei , ei i = i (no summation on i), where
i = ±1. Letting l = j in (4.15) then yields
(no summation on j), where we have appealed to (4.5). Summing over j now yields
X
Ric(ei , ek ) = j hRej ei ek , ej i.
j
S = g ij Rij .
4.4. THE DIVERGENCE, THE GRADIENT AND THE LAPLACIAN 29
note that this map is bilinear over the smooth functions and thus defines a (1, 1)-tensor field due to
the tensor characterization lemma. The components of this tensor field with respect to coordinates
would in physics notation be written ∇i X j . They are given by
Contracting the components of this tensor field yields a smooth function. We define the divergence
of X ∈ X(M ), written divX, to be the function which in local coordinates is given by
divX = ∇i X i = ∂i X i + Γiik X k .
gradf = (df )] .
In local coordinates,
gradf = g ij (∂i f )∂j .
The Laplacian of a function. Finally, taking the divergence of the gradient yields the Laplacian
∆f = div(gradf ).
In the case of Euclidean space, this definition yields the ordinary Laplacian. However, in the case
of Minkowski space, it yields the wave operator.
Exercise 71. Prove that (∇X η)(Y ) defined by (4.16) is linear over the smooth functions in
the argument Y (so that ∇X η is a one-form due to the tensor chracterization lemma). Prove,
moreover, that
∇X η = (∇X η ] )[ .
The components of this tensor field with respect to local coordinates is given by
where Γkij are the Christoffel symbols associated with the coordinates (xi ). Physicists would write
this equation as
∇i ηj = ∂i ηj − ηk Γkij .
In order to generalize this to tensorfields, let T be a tensorfield of type (k, l). We can then think
of T as a map from X(M )∗ × · · · × X∗ (M ) × X(M ) × · · · × X(M ) (k copies of X(M )∗ and l copies
of X(M )) to C ∞ (M ) which is multilinear over the smooth functions. If η1 , . . . , ηk ∈ X∗ (M ) and
Y, X1 , . . . , Xl ∈ X(M ), then ∇X T is defined by the relation
(∇X T )(η1 , . . . , ηk , X1 , . . . , Xl )
=X[T (η1 , . . . , ηk , X1 , . . . , Xl )]
(4.17)
− T (∇X η1 , η2 , . . . , ηk , X1 , . . . , Xl ) − · · · − T (η1 , . . . , ηk−1 , ∇X ηk , X1 , . . . , Xl )
− T (η1 , . . . , ηk , ∇X X1 , X2 , . . . , Xl ) − · · · − T (η1 , . . . , ηk , X1 , . . . , Xl−1 , ∇X Xl ).
Exercise 72. Prove that (∇X T )(η1 , . . . , ηk , X1 , . . . , Xl ) defined by the formula (4.17) is linear
over the smooth functions in η1 , . . . , ηk , X1 , . . . , Xl . Due to the tensor characterization lemma,
this implies that ∇X T is a tensorfield of type (k, l).
Exercise 73. Let (M, g) be a semi-Riemannian manifold and let ∇ be the associated Levi-Civita
connection. Prove that ∇g = 0.
We here follow this convention by denoting the components of the tensorfield defined by (4.18) by
∇i Tjk . We use this notation also in the case that the components are calculated with respect to
a frame as opposed to only coordinate frames. However, which frame we use should be clear from
the context. Note that ∇i Tjk = ∇i Tkj (this is a consequence of the fact that T is symmetric). We
define divT to be the one-form whose components are given by
(divT )k = g ij ∇i Tjk .
(divT )k = ∇i Tik ;
first you raise the first index and then you contract with the second index.
4.6. AN EXAMPLE OF A CURVATURE CALCULATION 31
where it is taken for granted that all the indices are calculated with respect to the frame {ei }.
Say that X = X i ei is a smooth vector field. Then
X X
(divT )(X) = X k (divT )(ek ) = X k (divT )k = X k i (∇ei T )(ei , ek ) = i (∇ei T )(ei , X).
i i
To conclude X
(divT )(X) = i (∇ei T )(ei , X).
i
Then the idea is to use the Koszul formula (3.4) to calculate the connection coefficients, defined
by
∇eα eβ = Γλαβ eλ . (4.22)
λ λ λ
Since γαβ = −γβα , it is sufficient to compute γαβ for α < β. Compute
f˙
[e0 , ei ] = − ∂i = Hei ,
f2
where H is the function defined by
f˙
H=− .
f
32 CHAPTER 4. CURVATURE
α
Thus γ0i = 0 unless α = i and
i
γ0i = H.
λ
for all α = 0, . . . , n. To conclude, the only γαβ ’s that do not vanish are
i i
γ0i = H, γi0 = −H, (4.23)
Lemma 74. Let (M, g) be a semi-Riemannian manifold and let {eα }, α = 0, . . . , n, be an or-
thonormal frame on an open subset U of M . Define the connection coefficients Γλαβ by the formula
λ
(4.22) and γαβ by (4.21). Then
1 β
Γλαβ = α
−λ α γβλ + λ β γλα λ
+ γαβ (4.24)
2
(no summation on any index), where α = g(eα , eα ).
where we used the fact that the frame is orthonormal in the second step and gαβ = heα , eβ i. On
the other hand,
2h∇eα eβ , eµ i = 2hΓναβ eν , eµ i = 2Γναβ gµν .
1
Γναβ gµν = ν ν ν
−γβµ gαν + γµα gβν + γαβ gµν .
2
1
Γλαβ = ν
g λµ gαν + γµα
ν
g λµ gβν + γαβ
ν
g λµ gµν
−γβµ
2
1 α β λ
= −λ α γβλ + λ β γλα + γαβ
2
(no summation on any index), where we have used the fact that gαβ = α δαβ . The lemma
follows.
4.6. AN EXAMPLE OF A CURVATURE CALCULATION 33
Calculating the connection coefficients in the case of the metric (4.19). Let us now
return to the metric (4.19). Considering the formula (4.24), it is of interest to note the following.
µ
In all the terms on the right hand side, the indices in the γαβ ’s are simply permutations of the
λ µ
indices in Γαβ . In our particular setting, the only combination of indices in γαβ that (may) give a
non-zero result is if one of the indices is 0 and the other two are equal and belong to {1, . . . , n}.
Let us compute, using (4.24),
1
Γi0i = 0 i i
−γii + γi0 + γ0i = 0,
2
1
Γii0 = −γ0i
i 0 i i
+ γii + γi0 = −γ0i = −H,
2
1
Γ0ii = −γi0i i 0 i
+ γ0i + γii = γ0i = H.
2
To conclude, the only connection coefficients which are non-zero are
Ric(eµ , eν ) = eα (Γα λ α α λ α λ α
µν ) + Γµν Γαλ − eµ (Γαν ) − Γαν Γµλ − γαµ Γλν , (4.26)
hReα eµ eν , eα i =heα (Γβµν )eβ + Γλµν Γβαλ eβ − eµ (Γβαν )eβ − Γλαν Γβµλ eβ − γαµ
λ
Γβλν eβ , eα i
=α eα (Γα λ α α λ α λ α
µν ) + Γµν Γαλ − eµ (Γαν ) − Γαν Γµλ − γαµ Γλν
(no summation on α), where we have used the fact that heα , eβ i = α δαβ . Combining this obser-
vation with (4.27) yields (4.26).
The components of the Ricci tensor of the metric (4.19). Let us now compute the
components of the Ricci tensor of the metric (4.19). Since the Ricci tensor is symmetric, it is
sufficient to compute Ric(eµ , eν ) for µ ≤ ν. Before computing the individual components, let us
make the following observations. Since the Γα α
µν ’s only depend on t, eλ (Γµν ) = 0 unless λ = 0.
Keeping in mind that the only non-zero connection coefficients are given by (4.25), we conclude
that
eα (Γα 0
µν ) = e0 (Γµν ) = 0
unless µ = ν = i. Moreover,
eα (Γα 0
ii ) = e0 (Γii ) = Ḣ.
34 CHAPTER 4. CURVATURE
The 00-component of the Ricci tensor. Compute, using the above observations as well as the
fact that the only non-zero connection coefficients are given by (4.25),
Ric(e0 , ei ) = eα (Γα λ α α λ α λ α
0i ) + Γ0i Γαλ − e0 (Γαi ) − Γαi Γ0λ − γα0 Γλi = 0;
since Γλ0β = 0 regardless of what λ and β are, the first, second and fourth terms on the right hand
side vanish; since Γ00i = 0 and Γjji = 0 regardles of the values of i and j, it is clear that Γα
αi = 0 (so
that the third term on the right hand side vanishes); in order for the first factor in the fifth term
to be non-vanishing, we have to have λ = α = j for some j = 1, . . . , n, cf. (4.23), but if λ = α = j,
then the second factor in the fifth term vanishes.
The ij-components of the Ricci tensor, i 6= j. If i 6= j, then
Proof. Let
ψ : (0, π) × (0, 2π) → S2
be defined by
ψ(θ, φ) = (sin θ cos φ, sin θ sin φ, cos θ). (4.30)
2
Then ψ is a diffeomorphism onto its image, the image of ψ is dense in S and
ψ ∗ gS2 = dθ ⊗ dθ + sin2 θ dφ ⊗ dφ. (4.31)
We leave the verification of these statements as an exercise. Due to these facts (and the smoothness
of the Ricci tensor and the metric), it is sufficient to verify that
g = dθ ⊗ dθ + sin2 θ dφ ⊗ dφ
satisfies Ric = g for 0 < θ < π and 0 < φ < 2π.
The metric g is such that we are in the situation considered in Section 4.6; replace t with θ; x1
with φ; n with 1; f (t) with sin θ; and with 1. As in Section 4.6, we also introduce the frame
1
e0 = ∂ θ , e1 = ∂φ .
sin θ
Note that
1
H=− ∂θ sin θ = − cot θ.
sin θ
Moreover,
cos2 θ 1
Ḣ = 1 + 2 = .
sin θ sin2 θ
Appealing to Lemma 76 then yields
1 cos2 θ
Ric(∂θ , ∂θ ) = Ric(e0 , e0 ) = 2 − = 1.
sin θ sin2 θ
Similarly, Ric(e1 , e1 ) = 1 and Ric(e0 , e1 ) = 0. The proposition follows.
Exercise 80. Let ψ be defined by (4.30). Prove that ψ is a diffeomorphism onto its image, that
the image of ψ is dense in S2 and that (4.31) holds.
36 CHAPTER 4. CURVATURE
4.7.2 The curvature of the upper half space model of hyperbolic space
Let us define Un by
Un = {x = (x1 , . . . , xn ) ∈ Rn : xn > 0}.
Moreover, define
n
1 X i
gUn = dx ⊗ dxi .
(xn )2 i=1
Then (Un , gUn ) is called the upper half space model of n-dimensional hyperbolic space. Here, we
do not sort out the relation between this model and the metric defined in Example 27, but we
calculate the Ricci tensor of gUn .
Lemma 81. Let 1 ≤ n ∈ Z and let Un+1 and gUn+1 be defined as above. Then
Ric[gUn+1 ] = −ngUn+1 ,
Denoting x0 by t, introducing f by f (t) = e−t , and letting = 1, we are exactly in the situation
considered in Section 4.6. Compute
f˙
H = − = 1.
f
Lemma 76 then yields Ric = −ng. The lemma follows.
Bibliography
[1] O’Neill, B.: Semi Riemannian Geometry. Academic Press, Orlando (1983)
[2] Lee, J. M.: Introduction to Smooth Manifolds (Second Edition). Springer, New York (2013)
[3] Spivak, M.: A comprehensive introduction to differential geometry, Volume two. Publish or
Perish, Inc., Houston, Texas, USA (1979)
37
Index
38
INDEX 39
Riemannian, 7 geodesic, 22
semi-Riemannian, 7 submanifold, 11
time orientable, 10 subspace, 5
Minkowski tangent vector, 10
metric, 8 vector, 4
scalar product, 1 vector field, 10
scalar product space, 1 Spacetime, 10
Musical isomorphisms, 10 Submanifold
semi-Riemannian, 8
Non-degenerate spacelike, 11
subspace of a scalar product space, 4 Subspace
Null lightlike, 5
geodesic, 22 spacelike, 5
vector, 4 timelike, 5
Opposite timecone, 5 Tangent vector
Orthonormal causal, 11
basis, 2 lightlike, 10
Parallel spacelike, 10
translation along a curve, 20 timelike, 10
vector field along a curve, 19 Time orientable
Preserving scalar products, 3 Lorentz manifold, 10
metric, 10
Ricci tensor, 27 Time orientation
Riemann curvature tensor, 25 manifold, 10
Riemannian scalar product space, 6
manifold, 7 Time oriented
metric, 7 Lorentz manifold, 10
scalar product, 3 Timecone, 5
scalar product space, 3 opposite, 5
Round metric on Sn , 8 Timelike
curve, 10
Scalar curvature, 28 geodesic, 22
Scalar product, 1 subspace, 5
Euclidean, 1 tangent vector, 10
index, 2 vector, 4
Lorentz, 3 vector field, 10
Minkowski, 1 Timelike vectors
non-degenerate, 1 future oriented, 6
Riemannian, 3 Torsion free
symmetric, 1 connection, 16
Scalar product space, 1
Euclidean, 1 Upper half space model, 36
Minkowski, 1
Riemannian, 3 Vector
time orientation, 6 causal character, 4
Semi-Riemannian lightlike, 4
manifold, 7 null, 4
metric, 7 spacelike, 4
submanifold, 8 timelike, 4
Semi-Riemannian manifold Vector field
complete, 22 lightlike, 10
Spacelike spacelike, 10
curve, 10 timelike, 10
40 INDEX
Vectorfield
causal, 11
Warped product, 11