You are on page 1of 10

Survey on the Geometry of Spacetime: Lorentzian

Manifolds and Singularity Theorems


Alexander Wickes

Friday 8th June, 2012

Introduction
Einstein’s relativistic theory of gravitation—general relativity—will shortly be a century
old. At its core is one of the most beautiful and revolutionary conceptions of modern
science—the idea that gravity is the geometry of four-dimensional curved spacetime.
— J. Hartle [1]

The modern study of gravitation is primarily grounded in the theory of general relativity,
where space and time are modeled together as points on 4-dimensional manifolds similar
to Riemannian manifolds but without the requirement that the metric be positive-definite.
The manifolds have a similar structure to their Riemannian counterparts but pick up an
additional causal structure and associated physical notions. Since Einstein’s development of
this geometric theory of spacetime in the early 20th century, the field of Lorentzian Geometry
has flourished and there has been much interplay between the physical study of gravitation
and relativity and the mathematical study of differential geometry. General relativity has
been one of the most successful physical theories to date.
In the mid-20th century however, Hawking and Penrose developed their famous singularity
theorems, which intuitively state that under certain physically reasonable assumptions
(primarily, what are known as energy conditions), a Lorentzian manifold representing spacetime
will not be geodesically complete. That is to say, geodesics are not guaranteed to extend
forever and may terminate in a finite amount of time, as at the center of a black hole. This
may imply that our understanding of space and time in the relativistic context is not entirely
accurate or complete.
Interestingly, the validity of the energy conditions at the heart of the singularity theorems
has recently come into question. In particular, the strong energy condition used in this
paper is hardly ever taken seriously anymore because there are simple examples of matter
configurations which violate it [2] (nonetheless it is helpful to start with when learning
singularity theorems). More generally, it was recently found that all of the commonly
accepted energy conditions are violated by subtle quantum effects and in fact in certain
classical systems as well [3]. Fortunately, mathematicians are undaunted by the unreality of
the situation and will undoubtedly appreciate the relevant geometry for its own sake.

1
This paper contains a broad overview of Lorentzian geometry, concepts of general relativity,
causal structure, and singularity theorems, largely following in the example of [4]. Also in this
paper we use the notions of the gradient of a scalar field on a manifold and the divergence of
vector fields, which generalizes directly to Lorentzian geometry from Riemannian geometry.
For reference, see for example do Carmo [5].

1 Lorentzian Geometry and General Relativity


In any physical theory, we are interested in physical events and their relationships with one
another. In relativity, these events collectively make up what is referred to as spacetime,
and are typically given the temporal coordinate t and spatial coordinates x, y, and z in
the Cartesian case, or else some other spatial coordinate system. More specifically, general
relativity models spacetime as a 4-dimensional Lorentzian manifold, a special kind of non-
Riemannian manifold having a metric which need not be positive-definite. Lorentzian
geometry is interesting in its own right because many of the results of Riemannian geometry
carry over, and there is additional causal structure as well, which will be discussed later.
Definition 1.1. A pseudo-Riemannian manifold (M, g) is a differentiable manifold M
equipped with a smooth, symmetric, non-degenerate 2-form g, called the metric, defining a
generalized inner product h · , · i. Note that this differs from a Riemannian manifold in that
the metric is no longer required to be positive-definite and so does not necessarily define a
proper inner product.
Definition 1.2. Let (M, g) be an n-dimensional pseudo-Riemannian manifold. Since the
metric is non-degenerate, the metric tensor has n nonzero eigenvalues. The signature (p, q)
of the metric denotes that there are p negative eigenvalues and q positive eigenvalues. This is
also denoted by (−, . . . , −, +, . . . , +) where there are p minus signs and q plus signs. If M
has signature (0, n) then M is a Riemannian manifold. If M has signature (1, n − 1) then
M is called a Lorentzian manifold.
Example 1.3. The simplest Riemannian manifold is Euclidean space, that is, Rn with
the Euclidean metric
g = (dx1 )2 + · · · + (dxn )2 . (1)
Similarly, the simplest Lorentzian manifold is Minkowski space, or Rn+1 with the Lorentz
metric
g = −(dx0 )2 + (dx1 )2 + · · · + (dxn )2 . (2)
As previously noted, Lorentzian geometry shares many features of Riemannian geometry
and so we do not need to work too much to get important results. For example, the Theorem
of Levi-Civita still holds: given a Lorentzian manifold, there exists a unique symmetric
affine connection which is compatible with the metric. The proof is just the same as in
Riemannian geometry (cf. do Carmo, [5]). Furthermore, this implies that the notion of
geodesics from Riemannian geometry carries over precisely to Lorentzian geometry.
Given all this, it is tempting to assume that all results carry over from Riemannian
geometry. However, this is simply not the case. First and foremost, note that it is not true
that every manifold admits a Lorentz metric [6].

2
On the other hand, loosening the signature of the metric results in interesting features
of Lorentzian geometry that are not found in Riemannian geometry. In particular, tangent
vectors now split into three classes.
Definition 1.4. Let M be a Lorentzian manifold, let p ∈ M and let v ∈ Tp M . Then v is
said to be
i. Timelike if hv, vi < 0,
ii. Spacelike if hv, vi > 0,
iii. Null or lightlike if hv, vi = 0.
This terminology may seem arbitrary at first, but it does make sense in the context of
physical notions. In relativity, no information can travel faster than the speed of light c.
Consider 4-dimensional Minkowski space with coordinates (t, x, y, z), in units where c = 1.
Given any spacetime point p, there is a hypercone of null geodesics extending from that point,
representing the paths of light rays traveling towards, passing through, and moving away from
the point. This hypercone is referred to as the lightcone. Furthermore, under the action of
a proper orthochronous Lorentz transformation (an element of the time-preserving special
orthogonal group SO+ (1, 3), which gives the change of basis between inertial reference frames),
a timelike vector can be made to point entirely along the time axis, and will furthermore
always be timelike no matter the inertial frame. Similarly, a spacelike vector can be made to
point entirely off the time axis and will always be spacelike; and lightlike vector will always
point along the lightcone, and will always be lightlike.
Note that the same terminology as above applies to curves in spacetime, γ : I ⊂ R → M ,
referring to the tangent vector γ̇(t). In particular, if γ is a geodesic, its classification is
constant, since  
d D γ̇
hγ̇, γ̇i = 2 , γ̇ = 0 (3)
dt Dt
for geodesics, just as in Riemannian geometry. Furthermore, if a curve is timelike then we
have  2  2  2  2
dx dy dz dt
+ + < =1 (4)
dt dt dt dt
so a timelike curve represents the path of a particle, with speed smaller than the speed of
light, unity. We call the quantity Z
τ (γ) = |γ̇| dt (5)
I
1/2
the proper time of the curve γ, where |v| = hv, vi is the length of the vector v.
In Minkowski space, the geodesics are the usual straight lines, representing the unchanging
path of a particle in a proper vacuum. Einstein’s theory developed the idea that gravitation
could be represented as the curvature of a Lorentzian manifold. This way, all particles,
including massless ones such as the photon, would follow paths according to the curvature of
spacetime due to gravity, an experimentally observed fact unaccounted for by Newtonian
gravity. Einstein worked out the equations for the metric of spacetime due to a distribution of
energy and momentum, telling us which Lorentzian manifolds represent physical gravitational
fields.

3
Definition 1.5. A Lorentzian manifold (M, g) satisfies the Einstein field equation if the
metric satisfies
1
Ric − gR = T (6)
2
where Ric is the Ricci curvature tensor, R is the scalar curvature, and T is the energy-
momentum tensor.
Example 1.6. Some common examples of the energy-momentum tensor are as follows.
i. Vacuum — There is no matter or energy,

T = 0. (7)

ii. Cosmological constant — There is dark matter and energy driving the acceleration
of distances induced by the metric

T = −Λg, Λ ∈ R. (8)

iii. Pressureless perfect fluid — Described by a rest mass density function ρ ∈ C ∞ (M )


describing the mass distribution, and a timelike unit velocity field U ∈ Vect(M )
describing the flow of the mass distribution, with

T = ρU [ ⊗ U [ (9)

where [ : Tp M → Tp∗ M is the musical (flat) isomorphism.

2 Causal Structure of Spacetime


We now turn to the causal structure of spacetime. Lorentzian manifolds have many features
which do not appear in Riemannian geometry, with mostly physically interpretable results.
Definition 2.1. Let M be a Lorentzian manifold and let p ∈ M . Two vectors v, w ∈ Tp M
have the same (opposite) time orientation if hv, wi < 0 (> 0). The manifold is time-
orientable if there exists a non-vanishing timelike vector field T ∈ Vect M .
If M is time-orientable then the tangent spaces can be consistently time-oriented; that
is, any two vector fields will have either the same or opposite time orientations everywhere.
Given a non-vanishing vector field T , a vector v ∈ Tp M with the same (opposite) orientation
as T is said to be future-pointing (past-pointing) with respect to T . A manifold is
time-oriented if we have chosen a preferred orientation according to some nonvanishing
vector field.
Definition 2.2. Let M be a time-oriented Lorentzian manifold.
i. A timelike curve γ : I ⊂ R → M is called future-directed if γ̇ is future-pointing.

ii. The chronological future of p ∈ M is the set I + (p) of all points to which p can be
joined by a future-directed timelike curve.

4
iii. A future-directed causal curve is a curve γ : I → M such that γ̇ is timelike or null
and future-pointing if non-zero.

iv. The causal future of p ∈ M is the set J + (p) of all points to which p can be connected
by a future-directed causal curve.

There are of course analogous definitions made by replacing future and plus signs above
with past and minus signs. Locally, these can be understood just as in the case of Minkowski
space. This is stated more precisely in the following proposition, from which we also see that
the implications in fact hold globally for Minkowski space.

Proposition 2.3. Let M be a time-oriented Lorentzian manifold. Then each point p0 ∈ M


has an open neighborhood U ⊂ M such that if p and q are points in the submanifold U then

i. there exists a unique geodesic joining p to q (U is said to be geodesically convex);

ii. q ∈ I + (p) iff there exists a future-directed timelike geodesic joining p and q;

iii. J + (p) = I + (p); and

iv. q ∈ J + (p) iff there exists a future-directed timelike or null geodesic joing p and q.

Proof. The existence of geodesically convex neighborhoods given an affine connection is a


basic result from Riemannian geometry [7] and holds equally well in Lorentzian geometry.
The rest of the proof is rather long and beside the point of this paper, so we omit it (cf.
[4]).
From this a simple corollary follows.

Corollary 2.4. Let M be a time-oriented Lorentzian manifold and p ∈ M . Then


If q ∈ I + (p) and r ∈ I + (q) then r ∈ I + (p).
If q ∈ J + (p) and r ∈ J + (q) then r ∈ J + (p).
I + (p) is an open set.

In the context of Lorentzian geometry, there are no curves of minimal length between
two points; any two points can be joined by piecewise null curves, with total length zero.
However, there exist curves of maximal length, which turn out to be timelike geodesics.
This is codified by the following proposition, which intuitively tells us that if two observers
meet at some point in spacetime, are separated, and meet again at a later point, then the
free-falling one (i.e. the one following a geodesic) will always measure a longer proper having
passed.

Proposition 2.5 (Twin Paradox). Let M be a time-oriented Lorentzian manifold and let
p0 ∈ M . Then there exists a geodesically convex open neighborhood U ⊂ M such that on the
submanifold U , the following holds: If p ∈ U , qinI + (p), γ is the timelike normal geodesic
joining p and q, and c is any arbitrary timelike normal curve joining p and q then τ (c) ≤ τ (γ),
with equality if and only if c = γ.

5
Proof. Choose U to be the set furnished by proposition 2.3. Let c : [0, 1] → U be a timelike
curve with c(0) = p and c(1) = q. We can write c as

c(t) = expp (r(t)n(t)) (10)




where r(t) ≥ 0 and n(t), n(t) = −1. Then we have
 
ċ(t) = d(expp )c(t) ṙ(t)n(t) + r(t)ṅ(t) , (11)

which we can rewrite as


ċ(t) = ṙ(t)X(t) + Y (t) (12)
where X and Y are orthonormal vector fields along c(t) with X timelike and Y spacelike.
Then we find that
Z 1
hṙX + Y, ṙX + Y i 1/2 dt

τ (c) = (13)
0
Z 1
ṙ hX, Xi + hY, Y i 1/2 dt
2
= (14)
0
Z 1
2
2 1/2
= ṙ − hY, Y i
dt (15)
0
Z 1
≤ ṙ(t) dt = r(1) = τ (γ), (16)
0

where we have used the fact that ṙ(t) > 0 since γ̇ is future-pointing. We see that equality
holds if and only if Y = 0, since Y is spacelike. In this case then we have c = γ.

Definition 2.6. A Lorentzian manifold M is said to be stably causal if it admits a global


time function, that is, a smooth function t : M → R such that grad t is timelike.

Remark 2.7. Notice that a stably causal Lorentzian manifold is time-orientable. It is


conventional to choose the time orientation defined by − grad t, so that t increases along
future-directed timelike curves. This implies that there are no closed timelike curves, which
provides us with a notion of causality—a particle cannot go back in time and affect its own
history.

Definition 2.8. Let M be a time-oriented Lorentzian manifold.

i. A smooth future-directed causal curve c : (a, b) → M is said to be future-inextendible


if limt→b c(t) does not exist.

ii. Given a subset S ⊂ M , the past domain of dependence of S is the set D− (S) of all
points p ∈ M such that any future-inextendible causal curve starting at p intersects S.

Once again we have the analogous definitions made by swapping future and past and
changing the sign. We also have the entire domain of dependence of S, the set D(S) =
D+ (S) ∪ D− (S).

6
Definition 2.9. A Lorentzian manifold M is said to be globally hyperbolic if it is stably
causal and there exists a global time function t : M → R such that the time slices (or
Cauchy surfaces) Σa = t−1 (a) satisfy D(Σa ) = M , that is, their domain of dependence is
the whole spacetime.
We conclude this section by noting that most physically reasonable spacetimes are globally
hyperbolic. We can only expect to make reasonable predictions using data on a time slice, and
furthermore one can formulate a well-behaved quantum field theory on a globally hyperbolic
spacetime [8].

3 The Singularity Theorem


In certain Lorentzian manifolds, such as the Schwarzschild spacetime produced by a spherically
symmetry matter configuration or the various cosmological models, there arise singularities
beyond which a timelike geodesic cannot be extended. We say that a spacetime is singular
if it is not geodesically complete. Unfortunately, the Hopf-Rinow Theorem does not carry
over to Lorentzian geometry due to the lack of proper distance functions. We will find in fact
that most realistic spacetimes are singular; this is the crux of the singularity theorems.
First, we define an important condition on a spacetime which is based on physically
reasonable ideas which we will not go into.
Definition 3.1. A Lorentzian manifold M n satisfies the strong energy condition if
Ric(V, V ) ≥ 0 for any timelike vector field V ∈ Vect M . If n ≥ 2 and the spacetime
also satisfies the Einstein field equation, this is equivalent to the requirement that the
energy-momentum tensor T satisfies
tr T
T (V, V ) ≥ hV, V i . (17)
n−1
Example 3.2. It is helpful to see the meaning of the strong energy condition in some basic
matter models, as in those of example 1.6.
Vacuum — Trivially satisfied.
Cosmological constant — Equivalent to Λ ≤ 0.
Pressureless perfect fluid — Equivalent to ρ ≥ 0.
We now make some definitions and define some variables which will be used through the
rest of the discussion.
Definition 3.3. Let M be a globally hyperbolic Lorentzian manifold and let Σ be a time
slice. Let γ be a timelike geodesic orthogonal to Σ starting from p ∈ Σ. If γ has no conjugate
points between γ(0 and γ(t0 ) then there exists an open neighborhood V of γ([0, t0 ]) which can
be foliated with a timelike congruence, a family of timelike normal geodesics orthogonal
to S. Let X ∈ Vect V be the unit timelike vector field formed by the tangent vectors of
the geodesics making up the congruence, and let K = ∇(X [ ) be the symmetric tensor
corresponding to the covariant derivative of X [ . Then the function θ = div X = tr K is called
the expansion of the congruence on V . Additionally, a point is said to be conjugate to Σ
if it is a conjugate point on one of the geodesics in V .

7
Remark 3.4. To give some intuition, note that it can be seen by the divergence theorem that
θ gives the logarithmic rate of change along X of the volume element of time slices along the
geodesics.
Proposition 3.5. Let M be a globally hyperbolic Lorentzian manifold satisfying the strong
energy condition. Let Σ be a time slice and p ∈ Σ a point where θ = θ0 < 0. If the geodesic γ
through p can be extended to a distance t0 = −n/θ0 > 0 to the future of Σ, then it contains at
least one point conjugate to Σ.
Proof. Suppose that there are no points on γ([0, t0 ]) conjugate to Σ. Then there exists an
open neighborhood V of the geodesic image as described in definition 3.3. It can be seen
that the Raychaudhuri equation holds,
X(θ) + tr K 2 + Ric(X, X) = 0. (18)
Since the spacetime satisfies the strong energy condition, we have Ric(X, X) ≥ 0. Furthermore,
by the Cauchy-Schwarz inequality we have
(tr A)2 ≤ n tr A> A (19)
for any n × n matrix A. Then since K is symmetric we have
θ2
tr K 2 = tr K > K ≥ . (20)
n
Now since X(θ) = Dθ Dt
, the covariant derivative along γ, the strong energy condition along
with the above inequality give
Dθ θ2
+ ≤ 0. (21)
Dt n
Integrating this equation up to t gives us the relation
1 1 t
≥ + . (22)
θ θ0 n
Since θ(0)−1 = θ0−1 < 0 and θ(t)−1 > 0 for t > t0 it follows by the mean value theorem that θ
must blow up at some point in [0, t0 ], which contradicts that θ is a smooth function on V .
Proposition 3.6. Let M be a globally hyperbolic Lorentzian manifold, Σ a time slice, p ∈ M
and γ a timelike geodesic through p and orthogonal to Σ. If there exists a conjugate point
between Σ and p then γ does not maximize the length among smooth timelike curves joining
p with Σ.
Sketch. Let q be a conjugate point along γ between p and Σ. Then there exists a timelike
geodesic γ orthogonal to Σ and approximately through q whose length is approximately
the same as that of γ. Now let U be a geodesically convex neighborhood of q furnished by
proposition 2.3. Let r be a point in U along γ between q and Σ, and let s be a point in U
along γ between q and p. Since U is geodesically convex, by proposition 2.3 there exist unique
timelike geodesics γrs joining r and s and γsp joining s and p. Now if γ 0 is the restriction of
γ joining r and Σ, then the piecewise smooth curve γ 0 = γ 0 + γrs + γsp has strictly greater
length than γ by the Twin Paradox. Now we can easily smooth the edges of γ 0 to construct
a smooth timelike curve joining p and Σ whose length is greater than that of γ.

8
The following proposition and theorem are very important in establishing the singularity
theorem. However, their proofs are long and tedious and detract from the point of this paper.
Therefore, we will point out the necessary ingredients, but defer to [4] for details.
Proposition 3.7. Let M be a globally hyperbolic spacetime, Σ a time slice and p ∈ D+ (Σ).
Then D+ (Σ) ∩ J − (p) is compact.
Notes. The proof is of a primarily topological nature. It hinges on the fact that we can form
a basis for the topology of M from simple neighborhoods. A simple neighborhood U ⊂ M
is a geodesically convex open set diffeomorphic to an open ball and whose boundary is a
compact submanifold of a geodesically convex open set. The proof also relies largely on the
causal structure of Lorentzian manifolds.
Corollary 3.8. Let M be a globally hyperbolic spacetime and let p, q ∈ M . Then
i. J + (p) is closed, and
ii. J + (p) ∩ J − (q) is compact.
Theorem 3.9. Let M be a globally hyperbolic spacetime, Σ a time slice and p ∈ D+ (Σ).
Then among the timelike curves connecting p to Σ there exists a timelike curve with maximal
length. This curve is a timelike geodesic orthogonal to Σ.
Sketch. The proof is again topological in nature. Denote by T (Σ, p) the space of all timelike
curves connecting p to Σ with the topology induced by the Hausdorff metric. The proof
involves first showing that the proper time function τ : T (Σ, p) → R, giving the proper time
τ (c) of the curve c ∈ T (Σ, p), is upper semicontinuous. Once we know the function is upper
semicontinuous, we can extend it continuously to the closure T (Σ, p). Then we know that
τ obtains its maximum at some point of T (Σ, p), from which we can construct a timelike
geodesic orthogonal to Σ with the same length, using geodesically convex neighborhoods and
the Twin Paradox.
Finally, we come to the celebrated singularity theorem.
Theorem 3.10. Let M be a globally hyperbolic spacetime satisfying the strong energy con-
dition, and suppose that the expansion satisfies θ ≤ θ0 < 0 on a time slice Σ. Then M is
singular.
Proof. We will show by contradiction that no future-directed timelike geodesic orthogonal
to Σ can be extended to proper time greater than τ0 = −n/θ0 > 0 to the future of Σ. So,
suppose that there exists a future-directed timelike geodesic γ orthogonal to Σ and defined
on the interval [0, τ0 + ] for some  > 0. Putting p = γ(τ0 + ), we have by theorem 3.9 that
there is a maximal timelike geodesic γ orthogonal to Σ, so that τ (γ) ≥ τ (γ) = τ0 + . But by
proposition 3.5 we know that γ must have a conjugate point in the interval [0, τ0 ], and then
by proposition 3.6 γ must not be maximal, a contradiction!
This concludes our discussion on the geometry of spacetime and singularity theorems. We
have laid out the basics of relativity and Lorentzian geometry, explored the causal structure
of Lorentzian manifolds, and proved a classical singularity theorem, a major result in both
geometry and physics.

9
4 References
[1] J. Hartle, Gravity: An Introduction to Einstein’s General Relativity, Pearson Education
(2003)
[2] S. Hawking and G. F. R. Ellis, The Large Scale Structure of Space-Time, Cambridge
University Press (1973)
[3] M. Visser and C. Barcelo, Energy Conditions and Their Cosmological Implications,
arXiv:gr-qc/0001099v1 (2000)
[4] J. Natario, Relativity and Singularities – A Short Introduction for Mathematicians,
Resenhas 6:309-335 (2005); arXiv:math/0603190v4
[5] M. do Carmo, Riemannian Geometry, Birkhauser (1993)
[6] J. M. Lee, Introduction to Smooth Manifolds, Springer (2003)
[7] S. Kobayashi and K. Nomizu, Foundations of differential geometry, vol. I and II, Wiley,
(1996)
[8] S. Hawking and R. Penrose, The Nature of Space and Time, Princeton University Press
(1996)

10

You might also like