Lorentzian Cobordisms, Compact Horizons and the Generic Condition

Eric Larsson

Master of Science Thesis Stockholm, Sweden 2014

Lorentzian Cobordisms, Compact Horizons and the Generic Condition

Eric Larsson

Master’s Thesis in Mathematics (30 ECTS credits) Degree programme in Engineering Physics (300 credits) Royal Institute of Technology year 2014 Supervisor at KTH was Mattias Dahl Examiner was Mattias Dahl

TRITA-MAT-E 2014:29 ISRN-KTH/MAT/E--14/29--SE

Royal Institute of Technology School of Engineering Sciences

KTH SCI SE-100 44 Stockholm, Sweden

URL: www.kth.se/sci

iii

Abstract

We consider the problem of determining which conditions are necessary for cobordisms to admit Lorentzian metrics with certain properties. In particu- lar, we prove a result originally due to Tipler without a smoothness hypothe- sis necessary in the original proof. In doing this, we prove that compact hori- zons in a smooth spacetime satisfying the null energy condition are smooth.

We also prove that the ”generic condition” is indeed generic in the set of Lorentzian metrics on a given manifold.

Acknowledgements

I would like to thank my advisor Mattias Dahl for invaluable advice and en- couragement. Thanks also to Hans Ringström.

Special thanks to Marc Nardmann for feedback on Chapter 2.

Contents

Contents iv

1 Lorentzian cobordisms 1 1.1 Existence of Lorentzian cobordisms ...... 2 1.2 Lorentzian cobordisms and causality ...... 4 1.3 Lorentzian cobordisms and energy conditions ...... 8 1.3.1 C 2 null hypersurfaces ...... 9 1.3.1.1 The null Weingarten map ...... 9 1.3.1.2 Generator flow on C 2 null hypersurfaces ...... 10 1.3.2 Structure of horizons ...... 16 1.3.2.1 Abstract horizons ...... 16 1.3.2.2 Cauchy horizons ...... 17 1.3.3 Tipler’s original proof ...... 18 1.3.4 Properties of nonsmooth horizons ...... 20 1.3.5 A smoothness theorem ...... 23 1.3.5.1 Outline of the proof ...... 23 1.3.5.2 Flow sets and generator flow ...... 23 1.3.5.3 Generator flow is area-preserving ...... 26 1.3.5.4 Smoothness from area-preserving generator flow . 38 1.3.6 Cobordisms satisfying energy conditions are trivial ...... 41 1.3.6.1 Cylinder extensions ...... 41 1.3.6.2 Tipler’s theorem without smoothness hypothesis . 43

2 Generic metrics satisfy the generic condition 46 2.1 Outline of the argument ...... 46 2.2 Construction of the bundles Ri and Bi ...... 48 2.3 Viewing a curvature tensor R as an endomorphism Rb ...... 49 2.4 Construction of NGi,t and NGi,` ...... 50 2.4.1 Definitions ...... 50 2.4.2 Manifoldness and codimension ...... 51 2.5 The map α ...... 55 CONTENTS v

2.6 Transversality ...... 57

A Background 61 A.1 Preliminaries ...... 61 A.1.1 Manifolds and fiber bundles ...... 61 A.1.2 Cobordisms ...... 62 A.2 Lorentzian geometry ...... 62 A.2.1 Cauchy developments and Cauchy horizons ...... 64 A.2.2 Causality conditions ...... 64 A.3 Energy conditions ...... 65 A.4 Geometric measure theory ...... 66 A.4.1 Regularity of functions ...... 66 A.4.2 Measure zero ...... 66 A.4.3 Density functions ...... 67 A.5 Jet bundles ...... 70 A.5.1 Jets of maps between manifolds ...... 71 A.5.2 Jets of sections of a smooth fiber bundle ...... 72 A.6 Whitney topologies ...... 72

B Assorted proofs 74 B.1 Complete generators ...... 74 B.2 Geodesically spanned null hypersurfaces ...... 84 B.3 Generalized Jet Transversality ...... 85

Bibliography 93 Introduction

An intriguing question which can be posed in the theory of general relativity is that of topology change: Is it possible for a spacelike slice of spacetime at one time to have a different topology than that of a spacelike slice at some other time? One way of making this question precise is through the concept of a Lorentzian cobordism (see Definition 1.0.2); that is, a spacetime whose boundary consists of disjoint spacelike submanifolds. The question about whether topology change is possible can then be interpreted as the question of whether physically interest- ing nontrivial Lorentzian cobordisms exist. For this question to be interesting the cobordism needs to have some compactness property. We will consider both the case when the cobordism is compact, and the case when the cobordism has the weaker property of causal compactness (see Definition 1.0.3). Of course, the most interesting cobordisms from the perspective of topology change are the connected cobordisms, so we will mainly be concerned with those.

Earlier results

The existence of Lorentzian cobordisms when no geometrical conditions are im- posed is essentially a problem of differential topology. It is equivalent to the exis- tence of a cobordism with a vector field with a prescribed direction at the bound- ary, and the problem of characterizing pairs of manifolds which are cobordant in this sense was considered by Reinhart [29]. We discuss this in Section 1.1. When geometrical conditions are imposed, it is significantly more difficult to construct Lorentzian cobordisms which are not diffeomorphic to S [0,1] for some × manifold S. There are two classical results about the non-existence of nontrivial Lorentzian cobordisms under certain hypotheses: In 1967 it was shown by Geroch [16] that nontrivial Lorentzian cobordisms can exist only if they contain closed timelike curves, and in 1977 it was shown by Tipler [32] that nontrivial Lorentzian cobordisms satisfying certain energy conditions cannot exist. It was noted by Borde [4] in 1994 that both of these theorems apply to noncompact but causally compact Lorentzian cobordisms as well, thereby establishing nonexistence of cer- tain Lorentzian cobordisms between noncompact manifolds when geometrical vii conditions are imposed. In contrast to the classical non-existence results, Nardmann [26] has more re- cently obtained a positive result: Any cobordism, subject to some topological con- ditions in low dimensions, which has a Lorentzian metric at all can be endowed with a metric satisfying the strong geometrical condition known as the ”dominant energy condition”. However, they are only ”weak Lorentzian cobordisms” in the sense that the boundary is not spacelike. The paper [33] is concerned with explicit constructions of Lorentzian cobor- disms using Morse theory. The related paper [34] constructs nontrivial Lorentzian cobordism-like objects satisfying certain causality and energy conditions. How- ever, the latter paper does not require the cobordisms to satisfy the compactness assumptions used elsewhere, and so the the problem considered there is com- pletely different than the problem considered by Geroch and Tipler. A survey of the problem of topology change can be found in [4]. In proving Tipler’s theorem, one works with a compact Cauchy horizon, and the question arises of which regularity a Cauchy horizon has. It was shown in [10] that Cauchy horizons are not necessarily smooth. In [3, Section IV] it was shown that a compact ”almost everywhere C 2” horizon satisfying the null energy con- dition is everywhere C 1. In [5, Section 4] it was asked whether compact Cauchy horizons are always smooth. This question was answered in the negative in [6], where it was also suggested that an energy condition might be sufficient for mak- ing compact horizons smooth.

Overview of the present work

The main results of this thesis are Theorem 1.3.41 concerning smoothness of com- pact Cauchy horizons subject to energy conditions, and Theorem 2.6.3 about the genericity of the ”generic condition”. It has been known for some time (see for instance [10, Section 1]) that the proof of Tipler’s result in [32] makes an implicit smoothness assumption, and hence is not sufficient to prove the theorem as it is stated. This has not been corrected in [4] where Borde proves the result with weaker hypotheses. Tipler’s original proof uses arguments from a part of the proof of the Hawking Singularity Theorem [19, p.295-298], and the same implicit assumption can be found there as well. A similar mistake was made in the original proof of the Hawking Area Theorem, and has since been corrected by Chru´sciel,Delay, Galloway and Howard [8]. Significant work was necessary to fill in the gaps in the proof of the Hawking Area Theorem, and the proof in [8] is technical. Fortunately, their methods may be adapted to the setting of Tipler’s theorem and we do so in Section 1.3 in order to present a more careful proof of the nonexistence of Lorentzian cobordisms which satisfy certain viii INTRODUCTION energy conditions. In the process of giving a complete proof of Tipler’s theorem, we prove in The- orem 1.3.41 that compact horizons in a spacetime satisfying the null energy condi- tion are smooth, thereby significantly generalizing the theorem in [3, Section IV], and providing an answer to the the question raised in [6, Section 4] of whether compact horizons which satisfy energy conditions are smooth. In Geroch’s original proof [16] of the nonexistence of Lorentzian cobordisms without closed timelike curves, and also in Borde’s proof [4] of the result with weaker assumptions, there is a small detail missing. In Section 1.2 we present a proof of the result with this detail added. In Chapter 2 we prove a result which to our knowledge is completely new. It concerns a certain curvature condition known as the ”generic condition”. The name implies that this condition should be generic in some sense. We prove that this is indeed the case; the generic condition is generic in the Whitney C 4 topology on the set of all Lorentzian metrics on a given manifold. Previous work in this di- rection has been performed by Beem and Harris (see [1] and [2]). They considered the generic condition at a point, and derived geometrical consequences of having many non-generic vectors at the same point. Appendix A contains notation and conventions, while Appendix B contains proofs which are not central to the development of the main results. Lorentzian cobordisms1

In this chapter, we will discuss the conditions under which topologically nontriv- ial Lorentzian cobordisms can exist. In Section 1.1 we briefly discuss this question without any geometrical constraints. We continue by addressing in Section 1.2 how causality conditions affect the existence of nontrivial Lorentzian cobordisms, and conclude in Section 1.3 with a discussion of the relation between energy con- ditions and nontrivial Lorentzian cobordisms. Notation and conventions can be found in Appendix A. We will need several different notions of Lorentzian cobordisms. Since the word ”cobordism” generally refers to a compact space, we will define the notion of a ”Lorentzian pseudocobordism”:

Definition 1.0.1. Let S1 and S2 be manifolds of dimension n without boundary. A Lorentzian pseudocobordism between S and S is a Lorentzian (n 1)-manifold 1 2 + M, the boundary of which is the disjoint union S S , such that S and S are 1 t 2 1 2 spacelike, and M admits a nonvanishing timelike vector field which is inward- directed on S1 and outward-directed on S2.

The classical notion of a Lorentzian cobordism is the following.

Definition 1.0.2. A Lorentzian pseudocobordism M between S1 and S2 is a com- pact Lorentzian cobordism (or simply Lorentzian cobordism) if M is compact.

It turns out, as was noted by Borde (see [4]), that many of the theorems about Lorentzian cobordisms continue to hold when the property of compactness is re- placed by the property of ”causal compactness”. We will call the resulting object a ”causally compact Lorentzian cobordism”:

Definition 1.0.3. A spacetime M is causally compact if I(p) is compact for each p M. ∈ Causal compactness captures the concept of ”compact in time”,while allowing the spacetime to be non-compact in the spatial directions.

Definition 1.0.4. A Lorentzian pseudocobordism M between S1 and S2 is called a causally compact Lorentzian pseudocobordism if M is causally compact. 2 CHAPTER 1. LORENTZIAN COBORDISMS

Remark 1.0.5. Of course, we immediately see that every (compact) Lorentzian cobordism is also a causally compact Lorentzian pseudocobordism.

Finally, we define the notion of a ”weak Lorentzian cobordism”, following [26]. The difference between a Lorentzian cobordism and a weak Lorentzian cobordism is that the boundary of a weak Lorentzian cobordism does not need to be space- like. However, we still require that there is a timelike vector field transverse to the boundary.

Definition 1.0.6. Let S1 and S2 be manifolds of dimension n without boundary. A weak Lorentzian pseudocobordism between S and S is a Lorentzian (n 1)- 1 2 + manifold M, the boundary of which is the disjoint union S S , such that M 1 t 2 admits a nonvanishing timelike vector field which is inward-directed on S1 and outward-directed on S2.

Definition 1.0.7. A weak Lorentzian pseudocobordism is a weak Lorentzian cobor- dism if it is compact.

Each weak Lorentzian cobordism may be transformed into a Lorentzian cobor- dism by a homotopy of the metric making the light cones narrower. The difference between the notions arises when one considers Lorentzian cobordisms satisfying energy conditions. We will discuss this further at the end of Section 1.3.

Definition 1.0.8. A Lorentzian pseudocobordism M between S1 and S2 is topolog- ically trivial if it is diffeomorphic to S [0,1]. 1 ×

1.1 Existence of Lorentzian cobordisms

Suppose that we are given two compact manifolds S1 and S2 without boundary. A complete characterization of when there is a Lorentzian cobordism between them follows from a result due to Reinhart [29]. The existence result we present in this section is attributed to both Reinhart and Misner (in dimension 3) in [16]. We prove a lemma to explicitly translate Reinhart’s result to our setting.

Lemma 1.1.1. Suppose that M is a compact n-manifold with boundary S S . 1 t 2 Then there is a Lorentzian metric g which makes M a Lorentzian cobordism be-

tween S1 and S2 if and only if there is a nonvanishing vector field V on M which is inward-directed on S1 and outward-directed on S2.

Proof. The existence of such a vector field V is included in the definition of a Lorentzian cobordism, so one implication is trivial. For the converse, suppose that such a vector field V is given. Choose an arbi- trary Riemannian metric σ on M. (This can be done by lifting metrics from Rn to 1.1. EXISTENCE OF LORENTZIAN COBORDISMS 3 coordinate neighborhoods of M and patching them together using a partition of unity. The result is a Riemannian metric since the space of Riemannian metrics is convex.) We may normalize V so that it has unit length in σ. Define a tensor g on M by g(X ,Y ) σ(X ,Y ) Cσ(X ,V )σ(Y ,V ) = − where C is a constant to be determined. Note that if C 1 then g is a Lorentzian > metric on M. This is seen by considering a σ-orthonormal basis containing Vp for the tangent space T M at some point p. Let i {0,1}. For each p S , there is p ∈ ∈ i some value of C 1 making T S spacelike, since V is by hypothesis transverse > p i to Tp Si . All smaller values of C also make Tp Si spacelike. Moreover, If some value of C makes T S spacelike for some point p S , then T S is also space- p i ∈ i q i like for all q in a neighborhood of p. Since S S is compact a finite number of 1 ∪ 2 such neighborhoods with corresponding values of Ci suffice to cover them, and hence C min C 1 makes S and S spacelike hypersurfaces in the Lorentzian = i > 1 2 manifold-with-boundary M. This completes the proof.

Theorem 1.1.2. Let S1 and S2 be compact manifolds. Then there is a Lorentzian cobordism between them if and only if they have the same Stiefel-Whitney numbers and Euler numbers.

Proof. By [29, Theorem 1] the hypotheses about Stiefel-Whitney numbers and Eu- ler numbers are equivalent to there being a cobordism between S1 and S2 with a nonvanishing vector field which is inward-directed on S1 and outward-directed on S2. By the previous lemma, this is equivalent to there being a Lorentzian cobor- dism between S1 and S2.

In particular we have the following result in dimension 4, due to the fact that any two 3-manifolds are cobordant (see Theorem A.1.9 and Theorem A.1.10) and have the same Euler numbers (as can be seen by Poincaré duality, see [18, Corol- lary 3.37]).

Corollary 1.1.3. Let S1 and S2 be compact manifolds of dimension 3. Then there is a Lorentzian cobordism between them.

These results do not answer the question about existence of causally compact Lorentzian pseudocobordisms. We leave the following question unanswered.

Problem 1.1.4. Given two manifolds S1 and S2, and a (not necessarily compact) manifold M with boundary S S , does there exists a metric on M which makes 1 t 2 M a causally compact Lorentzian pseudocobordism between S1 and S2? 4 CHAPTER 1. LORENTZIAN COBORDISMS

1.2 Lorentzian cobordisms and causality

The classical result about the relation between Lorentzian cobordisms and causal- ity conditions is a theorem due to Geroch [16, Theorem 2]. It states that a topolog- ically nontrivial Lorentzian cobordism cannot satisfy the chronology condition. It has been noted previously (see [26, Footnote 2, p. 28]) that the theorem is false as it is stated by Geroch. These issues may be resolved by requiring, as we have done in the definition of a Lorentzian cobordism, that the decomposition ∂M S S = 1 t 2 is such that an inward-directed timelike vector field on S1 is future-directed, while an inward-directed timelike vector field on S1 is past-directed. However, there is a detail missing, both from Geroch’s original proof and from the proof given by Borde [4, Theorem 1]. We will first provide a complete proof of the theorem, using the weaker compactness condition of causal compactness suggested by Borde [4], and then discuss which detail is missing from the earlier proofs.

Theorem 1.2.1. Let n 1, let S ,S be n-manifolds without boundary (not nec- ≥ 1 2 essarily compact, nor necessarily a priori connected). Let (M,g) be a connected

causally compact weak Lorentzian pseudocobordism between S1 and S2 which sat- isfies the chronology condition. Then there is a diffeomorphism ϕ: S [0,1] M 1 × → such that the submanifold ϕ({x} [0,1]) is timelike for every x S ; in particular M × ∈ 1 is topologically trivial, and S1 and S2 are diffeomorphic.

Proof. Let V denote the timelike vector field which is assumed to exist on M. Our

first goal is to prove that each flow line of V intersects S1 precisely once and S2 precisely once. Consider an inextendible future-directed flow line γ. Either γ has a future endpoint, or it does not. If it does not, then γ restricts to a timelike curve [0, ) M. Since it is timelike, its image is contained in I (γ(0)), which is com- ∞ → + pact by causal compactness of M. Hence the curve has some limit point x, in the sense that it intersects each sufficiently small neighborhood of x an infinite number of times. By modifying the curve slightly one may obtain a timelike curve which passes through x twice with the same tangent vector, and this contradicts the hypothesis that M is chronological. Hence γ has a future endpoint.

The future endpoint of γ belongs either to S1, to the interior of M, or to S2. The first case is impossible since V is inward-directed on S1. In the second case V must be zero at the endpoint, which contradicts the hypothesis that V is nonzero

at each point. Hence the future endpoint of γ belongs to S2. By similar arguments, γ has a past endpoint which belongs to S1. We have noted that each flow line of V intersects S1 and S2 at precisely one point each. Let Lp denote the flow line through a point p. Let U2 be an open neighborhood of S2 such that no integral curve passes the boundary of U2 twice. Let η2 be a function on M which is 1 outside of U2, and 0 precisely on S2. Consider 1.2. LORENTZIAN COBORDISMS AND CAUSALITY 5

the vector field η2V . The flow lines of this vector field on M \ S2 are the same as those of V . Since each point in M \ S2 belongs to some such flow line and hence can be reached by flowing from S1, and each point of S1 can be flowed an arbitrary time along η2V without leaving M \ S2, the vector field η2V induces a diffeomor- phism ψ : S [0, ) M \S . By choosing an orientation-preserving diffeomor- 1 1 × ∞ → 2 phism ρ : [0, ) [0,1) we get a diffeomorphism φ : S [0,1) M \ S defined ∞ → 1 1 × → 2 by φ (p,t) ψ (p,ρ(t)). 1 = 1 Similarly, choosing a neighborhood U1 of S1 and a smooth function η1 which is 1 outside of U and 0 on S we get a diffeomorphism φ : S (0,1] M \ S . 1 1 2 2 × → 1 Note that the flow lines of φ1 and φ2 are the same as for the original vector field V , so in particular it holds for each p S that φ ({p} (0,1)) φ ({q} (0,1)) for a ∈ 1 1 × = 2 × unique q S . We introduce a function mapping each point p to the correspond- ∈ 2 ing point q: Define the function q : S S by 1 → 2 φ ({p} (0,1)) φ ({q(p)} (0,1)). 1 × = 2 × A priori, the only thing we know about the function q is that it is bijective. We will 1 now show that it is in fact a diffeomorphism. By composing φ1 and φ2− we get a diffeomorphism β: S (0,1) S (0,1). This diffeomorphism is of the form 1 × → 2 × (p,t) (q(p),α(p,t)) 7→ for some function α. Precomposing β with the inclusion i : S , S (0,1) given 1 → 1 × by p (p,1/2) and postcomposing it with the projection π: S (0,1) S we see 7→ 2 × → 2 that π β i q. ◦ ◦ = Since all three of these maps are smooth, so is q. Similarly, using an inclusion i : S , S (0,1) and the projection π: S (0,1) S we see that 2 → 2 × 1 × → 1 1 1 π β− i q− ◦ ◦ = is also smooth. This shows that q is a diffeomorphism. In particular, S1 and S2 are diffeomorphic. 1 Let βˆ denote the map (q− ,id) β, i.e. the map βˆ(p,t) (p,α(p,t)). This ◦ = map is a composition of diffeomorphisms, and hence itself a diffeomorphism. Consider the quotient space of S [0,1) S (0,1] defined by identifying each 1 × t 1 × (p,t) S (0,1) with βˆ(p,t). This space is a smooth manifold diffeomorphic to ∈ 1 × S [0,1]. Define the map φ: S [0,1) S (0,1] M by 1 × 1 × t 1 × →  φ1(p,t) if (p,t) S1 [0,1), φ(p,t) ∈ × = φ (q(p),t) if (p,t) S (0,1]. 2 ∈ 1 × 6 CHAPTER 1. LORENTZIAN COBORDISMS

This map descends to a map S [0,1] M since if (p,t) S (0,1) then 1 × → ∈ 1 × 1 (φ (q,id) βˆ)(p,t) (φ (q,id) (q− ,id) β)(p,t) (φ β)(p,t) φ (p,t). 2 ◦ ◦ = 2 ◦ ◦ ◦ = 2 ◦ = 1 This map is bijective and agrees locally with the the diffeomorphisms onto their images φ and φ (q,id), and so is a diffeomorphism. We have then constructed a 1 2 ◦ diffeomorphism S [0,1] M. Moreover, by construction it holds for each p S 1 × → ∈ 1 that ϕ({p} [0,1]) is an integral curve of the timelike vector field V , so it is a timelike × curve. This completes the proof.

Remark 1.2.2. Geroch’s original proof [16, Theorem 2] of this statement, after it has been shown that each integral curve intersects both S1 and S2, proceeds by constructing a Morse function ϕ using parametrizations of the integral curves. However, it is not shown that this function is continuous, and this fact is nontriv- ial. Borde’s proof [4, Theorem 1] does not construct a diffeomorphism explicitly, but rather implies that such a diffeomorphism can be constructed from a suit- able parametrization of the integral curves. However, one would need to show that such a parametrization can be chosen smoothly. In our proof, this is done by using the cut-off functions η1 and η2 to ”normalize” the parametrizations of the integral curves to have infinite length.

In a sense, Geroch’s theorem completely answers the question of which causal- ity conditions can hold for a topologically nontrivial causally compact Lorentzian pseudocobordism. Such a space cannot be chronological, but is on the other hand always non-totally vicious since points on the boundaries cannot be part of any closed timelike curve. This means that we know where on the causal ladder de- scribed in Section A.2.2 a topologically non-trivial pseudocobordism belongs. The remaining question then concerns causality of topologically trivial pseudocobor- disms.

Example 1.2.3 (Chronology of trivial cobordisms). Figure 1.1 shows a topologi- cally trivial cobordism which is not chronological. Figure 1.3 provides a further illustration of some of the timelike curves in that cobordism. By a similar con- struction, shown in Figure 1.4, one may obtain a metric on S1 [0,1] for which the × inward-directed normal vector field is future-directed at both boundary compo- nents.

Example 1.2.4 (Causality of trivial cobordisms). Figure 1.2 shows a chronological but non-causal topologically trivial Lorentzian cobordism.

By using the theory of Cauchy horizons, which will be described in greater detail in Section 1.3, one may show the following. We suggest that the reader refers to Section 1.3 for the terminology used in the proof. 1.2. LORENTZIAN COBORDISMS AND CAUSALITY 7

Figure 1.1: A topologically trivial Figure 1.2: A topologically trivial non-chronological cobordism S1 chronological cobordism S1 [0,1] × × [0,1]. It contains a closed timelike which is not causal. It contains a curve shown in blue. closed lightlike curve shown in red.

Proposition 1.2.5. Let M be a causally compact Lorentzian pseudocobordism be- tween S1 and S2. Suppose that M is strongly causal. Then M is globally hyperbolic.

Proof. Choose the time orientation so that an inward-directed vector field at S1 is future-directed. Consider the Cauchy horizon H +(S1). If it is empty, then M is globally hyperbolic and we are done. If not, choose some point p H +(S ) and ∈ 1 consider the intersection H +(S ) I (p). Choose a generator γ through p. Since 1 ∩ − J −(p) I −(p) by [7, Corollary 2.4.19] we know that imγ is contained in the com- ⊆ © ª pact set I −(p). This means that the set γ(k) k N must have a point of accumu- ∈ lation q in the compact set I −(p). Each sufficiently small neighborhood of q is then intersected an infinite number of times by the causal curve γ, contradicting the hypothesis that M is strongly causal. Hence H +(S1) must be empty and we conclude that M is globally hyperbolic.

We leave the remaining possibility as an open problem.

Problem 1.2.6. Does there exist a causally compact Lorentzian pseudocobordism which is causal but not strongly causal? 8 CHAPTER 1. LORENTZIAN COBORDISMS

Figure 1.3: Some timelike curves in the cobordism S1 [0,1] shown in × Figure 1.1. The black arrows indi- cate that the vertical edges should be identified.

Figure 1.4: Timelike curves for a met- ric on S1 [0,1] which reverses the × time orientation.

1.3 Lorentzian cobordisms and energy conditions

In Section 1.2 we saw a theorem of Geroch [16, Theorem 2] stating that a topolog- ically nontrivial Lorentzian cobordism cannot satisfy the chronology condition. A result from 1977 by Tipler [32, Theorems 3 and 4] further implies that a nontriv- ial Lorentzian cobordism cannot satisfy certain energy conditions. Unfortunately Tipler’s original proof, the methods of which are also used in [19, p.295-298] for proving Hawking’s singularity theorem, is flawed in that it is implicitly assumed that a certain Cauchy horizon is C 2. In this section, we will prove Tipler’s result (Theorem 1.3.46) by a method which does not require such an assumption. The machinery we use is from [8], where similar arguments were used to prove Hawk- ing’s Area Theorem. Our proof of Tipler’s theorem also yields a smoothness result for compact Cauchy horizons subject to energy conditions. See Theorem 1.3.41. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 9

1.3.1 C 2 null hypersurfaces

1.3.1.1 The null Weingarten map

In this section we summarize properties of C 2 null hypersurfaces which we will need later. For details, see [15], [22], [14, Section II.1] or [8, Appendix A]. A null hypersurface H in a spacetime M is characterized by the fact that each

tangent space Tp H contains a unique (up to scaling) null vector Kp . It holds that the tangent space Tp H consists of those vectors of Tp M which are orthogonal to Kp . This means that any normal vector field K of H consists entirely of null vec- tors. We will call the integral curves of these vector fields generators of H. By [15, Proposition 3.1] these generators (when given a suitable parametrization) are . By straightforward computations it holds that

X ,Y X 0,Y 0 and K ,Y K ,Y 0 〈 〉 = 〈 〉 〈∇X 〉 = 〈∇X 0 〉

whenever X ,Y Tp H and X X 0 λ1K and Y Y 0 λ2K for some real numbers ∈ − = − = ± λ1,λ2. Inspired by this, we work instead with the quotient space Tp H RK . To ± simplify notation we will denote this space by Tp H K , and the corresponding vector bundle by T H±K . This quotient is independent of the particular choice of null vector field K , since all such vector fields differ only by scaling. We now define the null Weingarten map of H with respect to K by

b : T H±K T H±K , K → b (X ) K . K = ∇X This map is not independent of the particular choice of null vector field K . How-

ever, if f is a smooth nowhere vanishing function then b f K f bK since K is null. 2 1 = Note that if H is C , then K can be chosen C so that bK is continuous. Since all our spacetimes are time-oriented we may restrict attention to future-directed null vector fields K . This means that we can associate to each null hypersurface a family of null Weingarten maps which differ only by positive scaling. Since K is null the spacetime metric induces an inner product on T H±K . Using this inner product, we may define the null second fundamental form of H with respect to K by B (X ,Y ) b (X ),Y . K = 〈 K 〉 A straightforward computation shows that BK is symmetric. We will need the fol- lowing theorem, a proof of which can be found in [22, Theorem 30].

Theorem 1.3.1. Let H be a smooth null hypersurface in a spacetime M. Then the null second fundamental form of H is identically zero if and only if H is a totally submanifold of M. 10 CHAPTER 1. LORENTZIAN COBORDISMS

Remark 1.3.2. The theorem as stated in [22, Theorem 30] applies to null subman- ifolds in general, regardless of codimension, and so requires the submanifold to be ”irrotational”. This condition is automatically satisfied for null hypersurfaces.

Finally, we define the null mean curvature θK of a null hypersurface with re- spect to a null vector field K as the trace of the null Weingarten map:

θ trb . K = K

Recall that if K 0 λK is another null vector field then b λb . Hence θ λθ . = K 0 = K K 0 = K This means that the sign of θK is independent of the particular future-directed null vector field K used to compute θK . We will sometimes omit the vector field K from the notation. Recall that the integral curves of a null vector field K are reparametrizations of geodesics. If K is chosen to agree with γ˙ of a geodesic segment γ with affine parameter s, and b(s) is the family of null Weingarten maps with respect to K along γ, then 2 b0 b Re 0. (1.1) + + = Here 0 denotes derivative along γ, and Re denotes the fiberwise endomorphism ± ± Re: T H K T H K defined from the Riemann curvature tensor R by letting → Re(X ) R(X ,γ˙)γ˙. Note that it is not obvious that the derivative b0 exists, since b = is a priori only continuous. A proof of the fact that the derivative does exist and satisfies equation (1.1) can be found in [8, Proposition A.1]. From equation (1.1) one can derive the ”Raychaudhuri equation”. In particu- lar, one may derive a certain differential inequality for the null mean curvature. Let b be the null Weingarten map of a C 2 null hypersurface with respect to a future- directed null vector field K (scaled to give an affine parametrization of an integral θ 2 curve), and let θ denote the trace of b. Let S b n 2 id. Then the trace of b is = − − trb2 θ2/(n 2) tr(S2) so taking the trace of equation (1.1) yields = − + 2 θ 2 θ0 tr(S ) Ric(K ,K ) 0. (1.2) + n 2 + + = − Since b and id are self-adjoint, so is S. Hence tr(S2) 0 so ≥ θ2 θ0 Ric(K ,K ) 0. (1.3) + n 2 + ≤ − This is the differential inequality we will use later.

1.3.1.2 Generator flow on C 2 null hypersurfaces

A null vector field on a C 2 null hypersurface gives rise to a family of diffeomor- phisms which flow points along the vector field. The integral curves of such a vec- tor field are called generators, and we will refer to such a flow as a generator flow. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 11

The generator flow for time t will typically be denoted βt . Given a Riemannian metric σ with volume form ω, the Jacobian determinant J(βt ) with respect to σ is the real valued function defined by (β )∗ω J(β )ω. In this section, we will show t = t that Jacobian determinant of a generator flow with respect to a certain family of Riemannian metrics is related to the null mean curvature of the hypersurface. We choose to work with a past-directed vector field T since this is the case in which we will apply the lemma.

Lemma 1.3.3. Let H be a C 2 null hypersurface in a spacetime (M,g) of dimension n 1. Let V be an arbitrary unit timelike vector field on M, and define a Riemannian + metric σ on M by σ(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ). = + Let T denote the unique past-directed lightlike σ-unit vector field on H and let ω denote the σ-volume form induced on H. Let θ denote the null mean curvature of H with respect to the future-directed null vector field T . Then the Lie derivative of − ω with respect to T is L ω θω. T = −

Proof. Choose some point p H at which to evaluate L ω. Let e ,e ,...,e be a ∈ T 1 2 n g-orthogonal basis for Tp H such that

• e T , 1 = p • g(e ,V ) 0 for i 2,3,...,n, i = = • g(e ,e ) 1 for i 2,3,...,n. i i = = Recall that integral curves of null vector fields on null hypersurfaces are geodesic segments. Let γ be a segment of the integral curve of T through p with an affine parametrization. Extend the basis e1,...,en along γ by letting

• e T , 1 = • e 0 for i 2,3,...,n. ∇e1 i = = Here denotes covariant derivative with respect to g. Note that we do not yet ∇ n know that (ei )i 1 is a frame for H, since we first need to show that the ei are tan- gent to H. This= will follow from Claim II below.

Claim I: g(e ,V ) 1/p2 on all of γ 1 = Since e1 is g-lightlike and σ-unit

2 2 1 σ(e ,e ) g(e ,e ) 2¡g(e ,V )¢ 2¡g(e ,V )¢ = 1 1 = 1 1 + 1 = 1 12 CHAPTER 1. LORENTZIAN COBORDISMS proving the claim.

Claim II: g(e ,e ) 0 on all of γ 1 i = By choice of the basis at p, we know that g(e ,e ) 0 at p. It holds on γ that 1 i = e ¡g(e ,e )¢ g( e ,e ) g(e , e ) g( e ,e ) 1 1 i = ∇e1 1 i + 1 ∇e1 i = ∇e1 1 i since e 0. Since the integral curve of T is a reparametrization of the geodesic ∇e1 i = γ it holds that e λe for some function λ along γ. Hence ∇e1 1 = 1 e ¡g(e ,e )¢ λg(e ,e ). 1 1 i = 1 i

This can be seen as an ordinary differential equation for g(e1,ei ) along γ, and to- gether with the initial value g(e ,e ) 0 at p, the uniqueness theorem of ordinary 1 i = differential equations tells us that g(e ,e ) 0 on all of γ. 1 i = Claim III: g(e ,e ) δ for i, j 2,3,...,n on all of γ i j = i j = By choice of the basis at p, the identity holds at p. By choice of the extensions ei and e j

e g(e ,e ) g( e ,e ) g(e , e ) g(0,e ) g(e ,0) 0. 1 i j = ∇e1 i j + i ∇e1 j = j + i = Viewing this as an initial value problem tells us that g(e ,e ) δ on all of γ. i j = i j Claim IV: σ(e ,e ) p2g(e ,V ) for i 2,3,...,n on all of γ 1 i = i = Using the previous claims

2 σ(e1,ei ) g(e1,ei ) 2g(e1,V )g(ei ,V ) 0 g(ei ,V ) p2g(ei ,V ). = + = + p2 =

Claim V: σ(e ,e ) δ 2g(e ,V )g(e ,V ) for i, j 2,3,...,n on all of γ i j = i j + i j = Using the previous claims

σ(e ,e ) g(e ,e ) 2g(e ,V )g(e ,V ) δ 2g(e ,V )g(e ,V ). i j = i j + i j = i j + i j

Claim VI: det((σ(e ,e )) ) 1 i j i,j = Let a σ(e ,V ) for i 1,2,...,n. The matrix A with entries σ(e ,e ) can then by i = i = i j the previous claims be written as

 1 p2a p2a  2 3 ··· p 2   2a2 1 2a2 2a2a3  A  + 2 ···. = p2a3 2a2a3 1 2a  + 3 ···  . . . .  . . . .. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 13

Let   1 0 0 0  ···  p2a2 1 0 0  − ··· B  p2a3 0 1 0 . = − ···  p2a 0 0 1  − 4 ···  . . . . .  ...... Then detB 1. Moreover = 1 p2a p2a  2 3 ···   0 1 0  BA  ···. = 0 0 1  ···  . . . .  . . . ..

Hence det(BA) 1. This means that = det(BA) det A 1, = detB = proving the claim.

Proof part VII(Computation of LT ω): The volume form ω induced on H by the Riemannian metric σ can be expressed in the frame e1,e2,...,en as q ω det((σ(e ,e )) )e1 e2 en = i j i,j ∧ ∧ ··· ∧ where the ei are the covectors defined by ei (e ) 1 and ei (e ) 0 for i j. By the i = j = 6= previous claim the determinant is equal to 1, so

ω e1 e2 en = ∧ ∧ ··· ∧ on all of γ. We will use Cartan’s formula to compute the Lie derivative LT ω, so we need to extend the frame ei to a neighborhood of γ. Extend e1,...,en to a frame such that e T . Extend the dual frame e1,...,en in the natural way by letting 1 = ei (e ) 1 and ei (e ) 0 for i j. Rescale e if necessary so that ω e1 e2 en i = j = 6= n = ∧ ∧···∧ holds everywhere. We will now use this frame to compute LT ω at the point p. By Cartan’s formula L ω i dω d(i ω). T = T + T Since ω is an n-form on an n-manifold, dω 0. Hence = L ω d(i ω). T = T Since ω e1 e2 en = ∧ ∧ ··· ∧ d(i ω) d(e1(T )e2 en) d(e1(e )e2 en) d(e2 en). T = ∧ ··· ∧ = 1 ∧ ··· ∧ = ∧ ··· ∧ 14 CHAPTER 1. LORENTZIAN COBORDISMS

Hence n X k 2 k 1 k k 1 n LT ω ( 1) e e − de e + e . = k 2 − ∧ ··· ∧ ∧ ∧ ∧ ··· ∧ = We now compute dek , or rather the part of dek which does not contain any e j for j {1,k}; all such terms would be annihilated when we insert this expression into ∉ the large wedge product above. Since dek is a two-form this means that only one of its terms, (dek )(e ,e )e1 ek , is interesting. Now 1 k ∧ (dek )(e ,e ) e (ek (e )) e (ek (e )) ek ([e ,e ]) ek ([e ,e ]) 1 k = 1 k − k 1 − 1 k = − 1 k since ek (e ) 1 and ek (e ) 0 everywhere close to γ. We express the Lie bracket, k = 1 = evaluated at the point p, using the spacetime metric g:

(dek )(e ,e ) ek ([e ,e ]) ek ( e e ) ek ( e ). 1 k = − 1 k = − ∇e1 k − ∇ek 1 = ∇ek 1 Recall that denotes covariant derivative with respect to g. We have used that the ∇ choice of e ,...,e is such that e 0 for all k. We now know that 2 n ∇e1 k = dek (ek ( e ))e1 ek ... = ∇ek 1 ∧ + where the dots signify terms containing some e j for j {1,k}. Hence it holds at the ∉ point p that

k 2 k 1 k k 1 n ( 1) e e − de e + e − ∧··· ∧ ∧ ∧ ∧ ··· ∧ k 2 k 1 k 1 k k 1 n ( 1) e e − (e ( e ))e e e + e = − ∧ ··· ∧ ∧ ∇ek 1 ∧ ∧ ∧ ··· ∧ k k 2 k 1 2 k 1 k k 1 n ( 1) ( 1) − (e ( e ))e e e − e e + e = − − ∇ek 1 ∧ ∧ ··· ∧ ∧ ∧ ∧ ··· ∧ (ek ( e ))e1 e2 en = ∇ek 1 ∧ ∧ ··· ∧ k 2 1 2 k 1 where the additional factor of ( 1) − is due to commuting e with e ,...,e − . − Hence n n X k 1 2 n X k LT ω (e ( ek e1))e e e (e ( ek e1))ω. = k 2 ∇ ∧ ∧ ··· ∧ = k 2 ∇ = = Since e T and e ,...,e are g-orthonormal and g-orthogonal to e , 1 = 2 n 1 n n n X k X X e ( ek e1) g(ek , ek e1) g(ek , ek ( T )). k 2 ∇ = k 2 ∇ = − k 2 ∇ − = = = ± Recall from Section 1.3.1.1 that the quotient Tp H RT has an inner product in- n ± duced by g such that the image of (ei )i 2 under the projection Tp H Tp H RT = Pn → forms an orthonormal basis. This means that k 2 g(ek , ek ( T )) is the trace of = ∇ − the null Weingarten map b T defined in Section 1.3.1.1. This trace is the null mean − curvature θ of H with respect to T . We can then conclude that − L ω θω T = − at p. Since p was arbitrary, this completes the proof. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 15

The following lemma essentially consists of integrating the Lie derivative LT ω to relate the null mean curvature θ to the Jacobian determinant of the generator flow.

Lemma 1.3.4. Let H be a C 2 null hypersurface in a spacetime (M,g). Let σ be a Riemannian metric on M of the form

σ(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ) = + for some g-unit timelike vector field V on M. Let T be the unique past-directed σ- unit null vector field on H, and let θ be the null mean curvature of H with respect to T . Fix t 0 and let β : H H denote the flow along T for time t (whenever − > t → defined). Suppose that p is such that β (p) is defined for all s [0,t]. Let J(β ) s ∈ t denote the Jacobian determinant of βt with respect to σ. Then µ Z t ¶ J(βt )(p) exp θ(βs(p))ds . = − 0

Proof. Let ω denote the volume form of σ. The Jacobian determinant J(βt )(p) is characterized by

β∗(ω ) J(β )(p)ω . t βt (p) = t p Let α: [0,t] R denote the function such that →

β∗(ω ) α(s)ω . s βs (p) = p Note that α(0) 1 since β is the identity, and that α(t) J(β )(p). For simplicity = 0 = t of notation, we will abbreviate ωβs (p) by ωs. By definition of the Lie derivative ¯ d ¯ (LT ω)βs (p) ¯ βτ∗ωs τ. = dτ¯τ 0 + = Using the function α

βτ∗ωs τ β∗sβ∗s τωs τ β∗sα(s τ)ω0 α(s τ)β∗sω0. + = − + + = − + = + − Note that the pullbacks used in the above computation make sense for sufficiently small τ since β∗s is defined at p when s [0,t] and β∗s τ is defined at βs τ(p) when − ∈ + + s τ belongs to some sufficiently small open neighborhood of [0,t]. By definition + of α it holds that β∗s ωs α(s)ω0 so that ωs α(s)β∗sω0. Hence = = − α(s τ) βτ∗ωβs τ(p) + ωs. + = α(s) Returning to the Lie derivative, we now know that ¯ d ¯ α(s τ) α0(s) (LT ω)βs (p) ¯ + ωs ωs. = dτ¯τ 0 α(s) = α(s) = 16 CHAPTER 1. LORENTZIAN COBORDISMS

On the other hand, Lemma 1.3.3 tells us that

(L ω) θ(β (p))ω . T βs (p) = − s s Hence α0(s) θ(βs(p)). α(s) = − Solving this differential equation subject to the initial condition that α(0) 1 we = see that µ Z t ¶ α(t) exp θ(βs(p))ds . = − 0 Since α(t) J(β )(p) this completes the proof. = t

1.3.2 Structure of horizons

We begin by defining the abstract concept of a ”horizon” (following [8]), and state some previously known results about the regularity of horizons. We then prove that the Cauchy horizons we will work with are horizons in this sense.

1.3.2.1 Abstract horizons

Definition 1.3.5. Consider a spacetime. We say that an embedded topological hypersurface is past null geodesically ruled if every point on the hypersurface be- longs to a past inextendible null geodesic contained in the hypersurface. These geodesics are called generators.

Remark 1.3.6. Note that if a past null geodesically ruled hypersurface is a C 2 null hypersurface, then these generators are the same as those defined in Sec- tion 1.3.1.1.

Definition 1.3.7. Consider a spacetime. A horizon is an embedded, achronal, past null geodesically ruled, closed (as a set) topological hypersurface.

Remark 1.3.8. One may just as well define a horizon to be future null geodesically ruled. Indeed, in [8] the distinction is made between a ”past horizon” and a ”fu- ture horizon”. However, since we will work only with future Cauchy horizons it is convenient to restrict our attention to past null geodesically ruled horizons.

Remark 1.3.9. If an open subset of a horizon is past null geodesically ruled, then its generators are the restrictions of the generators of the horizon.

Note that we have assumed no smoothness in the definition. Note also that the generators through a point of a horizon are by no means necessarily unique. In fact, we have the following theorem (see Theorem 3.5 in [3] and Proposition 3.4 in [10]). 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 17

Theorem 1.3.10. A horizon is differentiable precisely at those points which belong to a single generator.

We also note that horizons are null hypersurfaces whenever they are differen- tiable, so that the generators of a C 2 horizon are precisely the integral curves of the null vector fields:

Proposition 1.3.11. If a horizon H is differentiable at a point p, then Tp H is a null hyperplane.

Proof. Since p belongs to a lightlike geodesic segment contained in H, we know that Tp H contains null vectors. If Tp H were to contain a timelike vector, then there would be a timelike curve in H with this tangent vector. This would contra- dict achronality of H, and hence Tp H must be a null hyperplane.

Finally, we note that generators can only intersect in common endpoints.

Proposition 1.3.12. Let H be a horizon, and suppose that p is an interior point of a generator Γ. Then there is no other generator containing p.

Proof. Suppose that some other generator Γ0 contained p. Let q be a point to the past of p along Γ0, and let r be a point to the future of p along Γ. By following Γ0 from q to p and then Γ from p to r we have connected q and r by a causal curve which not a null geodesic. By [7, Proposition 2.6.9] this curve cannot be achronal. Since the image of the curve belongs to H, this contradicts achronality of H.

1.3.2.2 Cauchy horizons

We now connect the statements in Section 1.3.2.1 about abstract horizons to the particular case of a Cauchy horizon in a spacetime. We begin by quoting [15, Proposition 2.7]. A similar statement can be found in [7, Proposition 2.10.6].

Proposition 1.3.13. Let S be an achronal subset of a spacetime M. Then the set 0 H +(S) \ edge(S), if nonempty, is an achronal C hypersurface of M ruled by null geodesics, each of which either is past inextendible in M or has a past endpoint on edge(S).

Corollary 1.3.14. Let M be a spacetime. Suppose that S M is an achronal set with ⊆ edge(S) . Then H +(S) is a horizon in the sense of Definition 1.3.7. = ;

Proof. The proposition tells us that H +(S) is a topological hypersurface which is achronal and past null geodesically ruled. To see that H +(S) is closed, note that it by definition is the difference of a closed set and an open set. This completes the proof. 18 CHAPTER 1. LORENTZIAN COBORDISMS

We conclude with a lemma allowing us to apply Corollary 1.3.14 to closed spacelike hypersurfaces.

Lemma 1.3.15. Let M be a spacetime and let S be a spacelike hypersurface which is closed as a set. Then edge(S) . = ; Proof. Let p S. After choosing a sufficiently small neighborhood U of p, the set ∈ S U is a Cauchy surface for U. In particular there is a time function on U the zero ∩ set of which is S U. By continuity, the time function evaluated along any timelike ∩ curve from I −(p,U) to I +(p,U) must take the value zero, and so every such curve must intersect S. This means that edge(S) S . Since S is closed and edge(S) S ∩ = ; ⊆ by definition, this completes the proof.

1.3.3 Tipler’s original proof

This section is given only for motivation. No other section depends logically on it. The following is the theorem about cobordisms which is proved (but not ex- plicitly stated) by Tipler in [32, Theorems 3 and 4]. Tipler did not mention the 2 need for the condition that H +(S1) is C . In the following sections, ending with Theorem 1.3.46, we will prove as a corollary to the smoothness theorem of Sec- tion 1.3.5 that this condition is not necessary. However, we begin by correctly stat- ing what Tipler’s original proof shows, and summarizing the idea of that proof. We have added some details not present in the original proof, and left others out. For instance, Tipler discusses other energy conditions than the strict lightlike conver- gence condition, as will we in later sections.

Theorem 1.3.16. Let n 2, let S ,S be compact n-dimensional manifolds and let ≥ 1 2 (M,g) be a compact connected Lorentzian cobordism between S1 and S2 which sat- isfies the strict lightlike convergence condition. Suppose moreover that H +(S1) is C 2. Then there exists a diffeomorphism ϕ: S [0,1] M such that the submani- 1 × → fold ϕ({x} [0,1]) is g-timelike for every x S ; in particular, S is diffeomorphic to × ∈ 1 1 S2.

Proof. For a rigorous proof, it is necessary to embed M in a manifold without boundary (as we will do in Section 1.3.6.1) to apply propositions about manifolds without boundary. For the purposes of this proof sketch, we ignore this complica- tion. If M satisfies the chronology condition then Theorem 1.2.1 gives the desired conclusion. Suppose for contradiction that M does not satisfy the chronology condition. Then it contains a closed timelike curve γ. We first wish to show that

H +(S ) is nonempty. Let q imγ. Note that γ is a timelike past inextendible curve 1 ∈ through q which does not intersect S1, so (according to Proposition A.2.13) it is 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 19 the case that q D (S ). Let Γ be an arbitrary curve contained in M from some ∉ + 1 point of S D+(S ) to q. Suppose that Γ does not contain a point of H +(S ). 1 ⊆ 1 1 By definition H +(S ) D (S )\ I −(D+(S )), so for each r Γ either r D (S ) of 1 = + 1 1 ∈ ∉ + 1 r I −(D+(S )). Let ∈ 1 t sup{s Γ(t) D (S )} = | ∈ + 1 and r Γ(t). = Then r D (S ) since this set is closed. Suppose to get a contradiction that r ∈ + 1 belongs to I −(D+(S )). This set is open so there is a point r Γ(t ²) for some 1 ² = + ² 0 such that r I −(D+(S )). Hence there is a point p D+(S ) such that there > ² ∈ 1 ∈ 1 is a timelike curve β from r to p. By maximality of t we know that r D (S ) so ∉ + 1 there is some past inextendible timelike curve with future endpoint r which does not meet S1. Concatenating α and β we get a past inextendible timelike curve with future endpoint p which does not meet S , contradicting that p D+(S ). 1 ∈ 1 Hence r I −(D+(S )), so by definition of H +(S ) D (S )\ I −(D+(S )) it holds ∉ 1 1 = + 1 1 that r H +(S1). Hence H +(S1) M is nonempty. ∈ ∩ 2 1 Since H +(S1) is assumed to be C , it has a C null vector field tangent to the null generators. The null generators are totally past imprisoned in the compact set M, so by Lemma B.1.1 they are past complete. Let θ denote the null mean curvature of H +(S1). From inequality (1.3) and the lightlike convergence condition we see that if θ 0 at some point then θ will diverge within a finite affine distance > to the past. This would contradict smoothness of θ, so we conclude that θ 0 ≤ everywhere. Let V be a timelike vector field and introduce the Riemannian metric σ defined by σ(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ). = + Let ω be the volume form induced on H +(S1) by σ and let T be the σ-unit past- directed null vector field on H +(S ). Let β : H +(S ) H +(S ) denote the flow 1 t 1 → 1 along T . Then Lemma 1.3.3 tells us that d Z Z ω θω. dt βt (H +(S1)) = − H +(S1) Since θ 0 we then know that ≤ d Z ω 0. dt βt (H +(S1)) ≥

However, βt is defined on all of H +(S1) since the generators of H +(S1) have no past endpoints, and hence β (H +(S )) H +(S ) so that t 1 ⊆ 1 d Z ω 0. dt βt (H +(S1)) ≤ 20 CHAPTER 1. LORENTZIAN COBORDISMS

Figure 1.5: Part of the future Cauchy horizon of a spacelike rectangle in (2 1)- + dimensional Minkowski space. Generators of the Cauchy horizon are shown in blue. A more complicated example in which no open subset of the horizon is dif- ferentiable is given in [10].

Hence it must hold that θ 0 on all of H +(S ). This, together with the strict light- = 1 like convergence condition, contradicts inequality (1.3). Hence M must satisfy the chronology condition, and we are done.

1.3.4 Properties of nonsmooth horizons

In general, Cauchy horizons are not C 2 hypersurfaces: Figure 1.5 shows an exam- ple of a non-C 2 Cauchy horizon. This particular example is ”almost C 2” in the sense that it has a dense open subset which is C 2, so we would expect that many results about C 2 hypersurfaces are applicable to this example. However, it was shown in [10] that Cauchy horizons are not necessarily almost C 2. This means that the proof of Tipler’s theorem needs to be modified to deal with the possibil- ity of horizons of lower regularity. For this reason, and to prove Theorem 1.3.41 about conditions which make Cauchy horizons smooth, we need some results about general horizons. The definitions and results in this section can be found in [8].

Definition 1.3.17. Let M be a spacetime, and let (a,b) Σ O M be an open × =∼ ⊆ subset such that each slice {t} Σ is spacelike and each curve (a,b) {p} is timelike. × × Let N M be a hypersurface. A function f : Σ (a,b) is said to be a graphing ⊆ → function of N if N O {(f (x),x) x Σ}. ∩ = | ∈ Theorem 2.2 of [8] says that any locally achronal hypersurface, in particular 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 21 any horizon, is semi-convex. (See Definition A.4.2.) This implies (see [8, Propo- sition 2.1]) every point on the horizon has a globally hyperbolic spacetime neigh- borhood ( a,a) Σ in which the horizon has a graphing function f for which there − × is a subset ΣA Σ such that l ⊆

1. Σ \ ΣAl has measure zero,

2. f is differentiable at all points of ΣAl ,

3. f is twice-Alexandrov-differentiable at all points x ΣA . In other words, ∈ l there is a quadratic form D2 f (x) such that for all y Σ ∈ 1 f (y) f (x) d f (x)(y x) D2 f (x)(x y,x y) o( x y 2). − − − = 2 − − + | − |

Moreover, it is shown in [8] that this notion is coordinate invariant: If p (f (x),x) = with x ΣA for one globally hyperbolic neighborhood of p, then p (f˜(x˜),x˜) ∈ l = with x˜ Σ˜ A for any other globally hyperbolic neighborhood ( a˜,a˜) Σ˜ of p, with ∈ l − × corresponding graphing function f˜. Hence the following definition makes sense.

Definition 1.3.18. Let H be a horizon in a spacetime. Denote by HAl the set of all points p H which are images under a graphing function of one of the corre- ∈ sponding sets ΣAl . We will call HAl the set of Alexandrov points of the horizon.

Remark 1.3.19. By the definition of semi-convexity a semi-convex function is the sum of a C 2 function and a convex function, and hence locally Lipschitz. This means that horizons are Lipschitz hypersurfaces.

We will now define the null mean curvature θAl and the null second funda- mental form BAl of HAl , following [8]. More precisely, we will define θAl and BAl on the intersection of HAl with a globally hyperbolic coordinate neighborhood O. This definition is not coordinate invariant. However, θAl and BAl are defined up to pointwise scaling by a positive function, so the sign of θAl is globally well defined.

Definition 1.3.20. Let H be a horizon in a spacetime (M,g). Choose a globally hyperbolic coordinate neighborhood O ( a,a) Σ of some point in H, and let = − × f : Σ ( a,a) be the graphing function of H in this neighborhood. Let the func- → − tion t :( a,a) Σ ( a,a) be the projection. For each point p O HA , with − × → − ∈ ∩ l x such that p (f (x),x), define k(p) dt d f (x). This makes sense since f is = = − + differentiable at all such points. Let K be the vector field dual to k with respect to g. Let e T O be the vector which is g-dual to dt. Choose a basis e ,...,e for 0 ∈ p 1 n Tp H such that 22 CHAPTER 1. LORENTZIAN COBORDISMS

• e K , n = p • g(e ,e ) 1 if 1 i n 1, i i = ≤ ≤ − • g(e ,e ) 0 if i j, i j = 6= • g(e ,e ) 0 if 1 i n 1. i 0 = ≤ ≤ −

We now define θAl and BAl using the coordinate formulae

n 1 X− i j ³ 2 µ ´ θAl el el Di j f Γi j kµ , = l 1 − =

a b a b i j ³ 2 µ ´ BA (X ea,Y e ) X Y e e D f Γ kµ . l b = a b i j − i j In the definitions above, we have constructed an ”artificial” covariant derivative µ k D2 f Γ k using the Alexandrov second derivative D2 f of f . We empha- ∇i j ≈ i j − i j µ size again that this definition of θAl is not independent of the coordinate system. However, the definitions using different coordinate systems differ only by a pos- itive multiplicative constant. In particular, the sign of θAl is invariantly defined. (See [8, Proposition 2.5].)

2 Remark 1.3.21. If H is C , then θAl and BAl agree with the null mean curvature θK and null second fundamental form BK defined in Section 1.3.1.1.

Remark 1.3.22. By [8, Theorem 5.1] the (1,1) tensor bAl associated to BAl satisfies equation (1.1) from Section 1.3.1.1.

We will later derive formulae involving θAl for the area of a horizon, and know- ing the sign of θAl will yield inequalities between different areas. In our case, the generators of the horizon will be past complete, so we will have use of the fol- lowing result. It is a generalization of [8, Proposition 4.17], and the proof of that proposition is sufficient for proving the generalization as well, since the proof is local.

Proposition 1.3.23. Let M be a spacetime, and let H be a horizon in M. Let Hb be an open subset of H which is also past null geodesically ruled. Suppose that the generators of Hb are complete in the past direction and that the null energy condition holds. Then

θA 0 on HA Hb . l ≤ l ∩ Remark 1.3.24. The result in [8] is expressed with the opposite time orientation compared to our setting. Consequently we obtain the inequality θA 0 instead l ≤ of θA 0. l ≥ 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 23

1.3.5 A smoothness theorem

In [5, Section 4] the question was posed whether a compact Cauchy horizon is necessarily smooth. A negative answer was given by the same authors in [6, Sec- tion 4], where it was also mentioned that compactness together with some energy condition might be sufficient to guarantee smoothness. In this section, we show using methods from [8] that this is indeed the case.

1.3.5.1 Outline of the proof

The theorem which will be proved in this section is Theorem 1.3.41, stating that compact horizons in a spacetime which satisfies the null energy condition are smooth. We first give an outline of the proof. Horizons are Lipschitz null hyper- surfaces, and so differentiable almost everywhere. At the points of differentiability there is a unique (up to scaling) null tangent vector, giving rise to an almost ev- erywhere defined vector field on the horizon. By restricting to a suitably chosen subset of the horizon, we may define a flow along this vector field. One may then construct a C 1,1 manifold containing, locally, this chosen subset, and extend the flow to a Lipschitz flow on the C 1,1 manifold. This is sufficient regularity to express how the area of a set changes under the flow, and this change of area is the central idea of the proof. To measure area, we introduce a Riemannian metric σ on the spacetime. With a suitably chosen such metric, the change in area is related to the

Alexandrov null mean curvature θAl of the Cauchy horizon. The argument for this 2 relation between area change and θAl proceeds via a C approximation of a part of the local C 1,1 approximation of the original horizon. Once the relation between

θAl and area change has been established, knowledge of the sign of θAl gives an inequality for area change under the flow. A sufficient condition under which the

sign of θAl may be determined is that all null geodesics in the horizon are com- plete in the past direction, together with an energy condition. That the generators are complete follows from Lemma B.1.1. By these arguments, we determine that the flow increases area. However, the flow maps a subset of the horizon into itself, thereby decreasing area. Hence the only possibility is that the flow conserves area. We show in Lemma 1.3.39 that this condition implies that the horizon is smooth.

1.3.5.2 Flow sets and generator flow

We wish to generalize the notion of the generator flow on C 2 null hypersurfaces discussed in Section 1.3.1.2 to possibly nonsmooth horizons. In other words, we want a flow along generators of a horizon H. However, since some points belong to several generators it is in general not possible to do this on all of H. Instead, we construct a smaller subset on which to define the flow. 24 CHAPTER 1. LORENTZIAN COBORDISMS

Definition 1.3.25. Let H be a horizon in a spacetime M. Define the total flow set of H to be the set A (H) of points p H such that the following conditions are 0 ∈ satisfied:

• There is a unique generator Γ of H passing through p.

• The point p belongs to the interior of Γ.

• Each interior point of Γ is an Alexandrov point.

Let σ be a Riemannian metric on M. For δ 0 define the δ-flow set of H with > respect to σ to be the set

A (H,σ) {p A (H) The generator through p exists δ = ∈ 0 | for a σ-distance greater than δ to the past and to the future}.

Remark 1.3.26. Note that the total flow set is the union of all δ-flow sets:

[ A0(H) Aδ(H,σ). = δ 0 > Remark 1.3.27. When the context allows it, we will sometimes drop H and σ from the notation and write simply A0 or Aδ.

For the next proposition, we will need the following result, a proof of which can be found in [8, Theorem 5.6].

Lemma 1.3.28. Let H be a horizon in a spacetime M of dimension n 1. Suppose + that S is a C 2 hypersurface intersecting H properly transversally (in the sense that if q S H and the tangent space T H exists then T S is transverse to T H). Define ∈ ∩ q q q S {q S H q is an interior point of a generator of H}, 0 = ∈ ∩ | S {q S all interior points of the generator 1 = ∈ 0 | through q are Alexandrov points of H}. Then S has full (n 1)-dimensional Hausdorff measure in S . 1 − 0 Proposition 1.3.29. Let H be a horizon in a spacetime (M,g) of dimension n 1, + and let σ be a Riemannian metric on M. Let hn be the n-dimensional Hausdorff measure induced by the distance function induced by σ. Then the total flow set A0 of H has full hn-measure in the sense that

hn(H \ A ) 0. 0 = 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 25

Proof. To show that H\A0 has measure zero, it is sufficient to show that each point p H has an open neighborhood U M such that hn (U (H \ A )) 0, for H can ∈ ⊆ ∩ 0 = be covered by countably many such neighborhoods since it is second-countable. The idea of the proof is to construct a spacetime of one dimension greater than M and apply Lemma 1.3.28 in this higher-dimensional setting. To this end, choose a globally hyperbolic neighborhood U M of p H dif- ⊆ ∈ feomorphic to ( a,a) Σ, where Σ Rn. With this notation, we mean that each − × ⊆ slice {t} Σ is spacelike. We may choose the zero slice to be such that p {0} Σ. × ∈ × Let Mc M I denote the product manifold with the metric g g ds2, where s = × b = + refers to the coordinate in the open interval I. Let Hb H I. Let π: Mc M de- = × → note the projection. We now verify that Hb is a horizon in Mc, which follows easily from H being a horizon in M. Since H is an embedded topological hypersurface, 1 so is Hb . By definition of the product topology, Hb π− (H) is closed. If there were = some timelike curve γ between two points of Hb , then π γ would be a timelike ◦ curve between two points of H contradicting achronality of H, so Hb must also be achronal. (To see that π γ is indeed timelike, note that by definition of g it holds ◦ b that g(π V ) g(V ) for all vectors V .) Finally Hb is past null geodesically ruled ∗ ≤ b since if Γ is a past inextendible null M-geodesic contained in H, then Γ {s} is a × past inextendible null Mc-geodesic contained in Hb for each s I. ∈ We now wish to construct, after possibly decreasing a or shrinking I, a diffeo- morphism ρ : I ( a,a) such that the hypersurface → −

S : ©(t,q,s) ( a,a) Σ I t ρ(s)ª = ∈ − × × | = is spacelike. A possible choice of basis for the tangent space T S T R Σ R (t,q,s) ⊆ q × × of S at some point (t,q,s) consists of a basis for the tangent space of Σ together with the vector (ρ0(s),0,1). The basis of Tq Σ consists of spacelike vectors since Σ is spacelike, and if ρ0(s) is sufficiently close to zero then (ρ0(s),0,1) is also spacelike (since (0,0,1) is spacelike by definition of gb, and the set of spacelike vectors at a point is open). This means that for each (t,q) ( a,a) Σ, there is some c(t,q) 0 ∈ − × > such that if ζ c(t,q) then for any s I the vector (0,ζ,1) T is spacelike. < ∈ ∈ (t,q,s) Since g is smooth, c can be chosen smooth. Hence c takes some minimum on every compact subset of ( a,a) Σ, This minimum is positive since c is positive on − × ( a,a) Σ, after possibly shrinking Σ and a. We may then find some real number − × ζ such that 0 ζ c(t,q) for all (t,q) ( a,a) Σ. Letting ρ(s) ζ(s s ) where < < ∈ − × = − 0 s0 is the midpoint of I, and subsequently shrinking I or a to make ρ bijective, we have found a diffeomorphism ρ making S spacelike. Since ρ is a diffeomorphism, the restriction of the projection π: Mc M to S is also a diffeomorphism. → Now S is a smooth hypersurface in Mc, which intersects Hb properly transver- sally in the sense that if q S Hb and the tangent space T Hb exists then T S is ∈ ∩ q q 26 CHAPTER 1. LORENTZIAN COBORDISMS

transverse to Tq Hb . Let

Sc0 {q S Hb q is an interior point of a generator of Hb }, = ∈ ∩ | S {q U H q is an interior point of a generator of H}. 0 = ∈ ∩ | Note that π(Sc0) S0 since if p is an interior point of a generator Γ then π(p) is an = interior point of the generator π(Γ) and vice versa. Note further that it holds that π(S Hb ) U H. Moreover, the projection π restricted to S is bijective and hence ∩ = ∩ π((S Hb )\ Sc0) (U H)\ S0. ∩ = ∩ 1 Since π restricted to S is a diffeomorphism, both π S and its inverse (π S)− are n | n | locally Lipschitz so that h ((S Hb )\ Sc0) 0 if and only if h ((U H)\ S0) 0. The ∩ = ∩ = latter set (U H)\S is the set of endpoints of generators of H contained in U. It is ∩ 0 shown in [9] that this set has zero hn-measure. This means that we can conclude n ³ ´ that h (S Hb )\ Sc0 0. In other words, Sc0 has full measure in S Hb . ∩ = ∩ Let

Sc1 {q Sc0 all interior points of the generator = ∈ | through q are Alexandrov points of Hb }. n n By Lemma 1.3.28 Sc1 has full h measure in S0. Hence it also has full h -measure in n S Hb . Since π is bi-Lipschitz, π(Sc1) has full h -measure in U H. ∩ ∩ The projection π: Mc M maps generators to generators, and Alexandrov → points of Hb to Alexandrov points of H, so each point of Sc1 belongs to A0. We have then shown that A0 U contains a subset π(Sc1) which has full measure in H U. ∩ ∩ Hence A itself has full measure in H U. As noted in the beginning of the proof, 0 ∩ H may be covered by countably many such sets U, so we have shown that A0 has full hn-measure in H. This completes the proof.

Definition 1.3.30. Let H be a horizon in a spacetime (M,g), and let A0 be its total flow set. Let σ be a Riemannian metric on M. Since H is differentiable at all points in A0, there is a unique σ-unit past-directed null vector tangent to H at each point in A0. This defines a vector field T on A0, which is tangent to the generators of H. Recall that A0 contains full generators, and hence full integral curves of T . We will call the flow of T , whenever it is defined, the generator flow of H with respect to σ, and denote it by (t,p) β (p). 7→ t The choice of A was made so that A flows into itself, in the sense that if p A 0 0 ∈ 0 and t 0 are such that β (p) is defined, then β (p) A . ≥ t t ∈ 0

1.3.5.3 Generator flow is area-preserving

The purpose of this section is to prove that the generator flow on a horizon with respect to a certain family of Riemannian metrics preserves the associated Haus- dorff measure if the null mean curvature is non-positive. Our immediate goal is 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 27 to construct a C 1,1 approximation of the horizon to be able to express the volume change. We do this in Lemma 1.3.32 and Lemma 1.3.33. We then construct a C 2 approximation of the horizon to compute the volume change in Proposition 1.3.35 and Proposition 1.3.38. The complicated constructions necessary are contained in Lemma 1.3.37. We begin by stating an extension theorem, which is proved in [8, Proposition 6.6].

Lemma 1.3.31. Let B Rn be an arbitrary subset and f : B R be an arbitrary ⊆ → function. Suppose that there is some constant C 0, and some function B Rn, > → p a (not necessarily continuous) such that the following two conditions hold: 7→ p 1. f has global upper and lower support paraboloids of opening C. Explicitly, for all x,p B, ∈ ¯ ¯ ¯f (x) f (p) x p,a ¯ C x p 2. − − 〈 − p 〉 ≤ || − ||

2. The upper and lower support paraboloids of f are disjoint. Explicitly, for all p,q B and all x Rn, ∈ ∈ f (p) x p,a C x p 2 f (q) x q,a C x q 2. + 〈 − p 〉 − || − || ≤ + 〈 − q 〉 − || − ||

Then there is a function F : Rn R of class C 1,1 such that f is the restriction of F to → loc B.

Using this lemma, we may prove the following.

Lemma 1.3.32. Let H be a horizon in an (n 1)-dimensional spacetime M. Let σ + be any Riemannian metric on M, let δ 0 and let A be the δ-flow set of H with > δ respect to σ. Let p A . Then there is some open globally hyperbolic neighborhood ∈ δ V M of p and a C 1,1 hypersurface N V in M such that A V N. ⊆ ⊆ δ ∩ ⊆

Proof. For each point q A , let q+ denote the point a σ-distance δ to the future ∈ δ along the unique generator through q. Similarly, let q− denote the point along the generator a distance δ to the past. By one of the defining properties of Aδ, we have q+,q− H. ∈ By the same reasoning as is used in the proof of Lemma 6.9 in [8], one may obtain a globally hyperbolic neighborhood V W of p and a constant C 0 with ⊆ > the following properties:

• V is diffeomorphic to ( a,a) B n(r ) with the slices {t} B n(r ) spacelike and − × × the curves ( a,a) {x} timelike and future-directed for all t ( a,a) and all − × ∈ − x B n(r ). ∈ 28 CHAPTER 1. LORENTZIAN COBORDISMS

• Let f denote the graphing function of the horizon over B n(r ), i.e. the func- tion such that V H {(f (x),x) q B n(r )}. For each q (f (x ),x ) ∩ = | ∈ = q q ∈ V A , the graph of the function ∩ δ 2 f −(x) f (x ) d f (x )(x x ) C x x , q = q + q − q − || − q || with the exception of the point q (f (x ),x ) itself, lies in the timelike past = q q I −(q+,V ) of q+.

• For each q (f (x ),x ) V A , the graph of the function = q q ∈ ∩ δ 2 f +(x) f (x ) d f (x )(x x ) C x x , q = q + q − q + || − q || with the exception of the point q (f (x ),x ) itself, lies in the timelike fu- = q q ture I +(q−,V ) of q−.

Note that if this holds for some value of C, it holds for all larger values of C as well. We will show that these conditions imply the first hypothesis of Lemma 1.3.31.

Suppose that the condition is violated. Then either f (x) f +(x) or f (x) f −(x) > q < q for some q (f (x ),x ) A and x B n(r ). The argument is the same for both = q q ∈ δ ∈ cases, so suppose without loss of generality that the first is the case. Since f (x ) q = f +(x ) we must have x x . Then (f +(x),x) belongs to the timelike future of q−, q q 6= q q by the choice of C. However, since f (x) f +(x), the point (f (x),x) lies to the > q timelike future of (fq+(x),x). This means that we can connect q− to (fq+(x),x) to (f (x),x) by a timelike curve. Hence (f (x),x) belongs to the timelike future of q−. Since both points belong to the horizon, this violates achronality of the horizon. This proves the first hypothesis of Lemma 1.3.31. For the second hypothesis, note that the first continues to hold if we increase C. By making sure that C is sufficiently large compared to the Lipschitz constant of f and the values of f , one may conclude as in the proof of Lemma 6.9 in [8] that the second hypothesis is satisfied as well. Let B denote the projection of A V on B n(r ). We can then apply the ex- δ ∩ tension theorem described in Lemma 1.3.31 to obtain a C 1,1 extension Rn R ¯ → of f ¯ : B ( a,a). Let F : Rn R denote the restriction to B n(r ) of this exten- B → − → sion. By definition F agrees with f on B. In particular, the graph of F contains p (f (x ),x ), so F (x ) ( a,a). Since F is continuous, there is some neigh- = p p p ∈ − borhood B n(²) B n(r ) of x such that F (B(²)) ( a,a). Hence by shrinking the ⊆ p ⊆ − neighborhood V to ( a,a) B n(²) and letting N be the graph of F there, we have − × obtained a C 1,1 hypersurface containing A V . δ ∩ Lemma 1.3.33. Let H be a horizon in an (n 1)-dimensional spacetime (M,g) + equipped with a Riemannian metric σ, let δ 0, let A be the δ-flow set of H with > δ 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 29

respect to σ, let A˜δ be the full-density subset (in the sense of Definition A.4.9) of Aδ, let V be a globally hyperbolic open neighborhood of p and let N V be a C 1,1 hy- ⊆ persurface containing A V , which can be represented by a graphing function in δ ∩ V. Fix t 0. Let β : A V A be the restriction of the generator flow (with ≥ t δ ∩ → 0 respect to σ) to A V , and suppose that this flow is defined on all of A V . Then δ ∩ δ ∩ there is a neighborhood U V of p such that the restriction of β to A˜ U is the ⊆ t δ ∩ restriction of a locally Lipschitz function βct : N U M. ∩ → Proof. Let (a,b) Σ be a decomposition in space and time of the globally hyper- × bolic neighborhood V of p, and let f denote the graphing function of N with re- spect to this decomposition. We wish to construct a Lipschitz normal vector field to N in a neighborhood of p (f (x),x). Choose a frame (e )n close to p consisting of the pushforward of a = i i 1 frame of Σ close to x under the map= y (f (y), y). This frame is Lipschitz since f is 7→ C 1,1. By shrinking V we may assume that the frame covers all of N. The condition that a vector field n along N is normal to N with respect to the spacetime metric g, and consists of unit vectors with respect to the Riemannian metric σ can be expressed by saying that n satisfies the n 1 equations + σ(n,n) 1 0, − = g(e ,n) 0 1 = g(e ,n) 0 2 = . . g(e ,n) 0. n = n n 1 n 1 Trivialize TN using the frame (e ) . Define F : N R + R + by the above i i 0 × → equations. Explicitly =

F (n) (σ(n,n) 1,g(e ,n),...,g(e ,n)). = − 1 n n 1 Let n be a zero of F . The tangent map of F at n with respect to the R + component is k (2σ(n,k),g(e ,k),...,g(e ,k)). 7→ 1 n We wish to show that this tangent map has full rank. For dimensional reasons, this is equivalent to its kernel being trivial. If k belongs to the kernel, then g(e ,k) 0 i = for all 1 i n. Hence k is a normal vector to N, and hence parallel to n. If k ≤ ≤ belongs to the kernel of the tangent map then it also holds that σ(n,k) 0. When k = is parallel to n, this can only happen when k 0. This shows that the tangent map n 1 = of the R + component of F has full rank. Clearly N has a tangent vector n which is 30 CHAPTER 1. LORENTZIAN COBORDISMS a zero of F at the point p. This means that we can apply Clarke’s Implicit Function Theorem (Corollary, p. 256 in [11]) to conclude that there is a Lipschitz function n satisfying F (q,n(q)) 0 in a neighborhood of p. By shrinking V if necessary, we = have then found a Lipschitz normal (with respect to g) vector field to N which is of unit length (with respect to σ). By considering graphing functions of N and H and applying the result about tangent spaces at full-density points described in Proposition A.4.12 we see that the tangent spaces T N and T H agree at all q A˜ . Consider now a point q A˜ . q q ∈ δ ∈ δ By Proposition 1.3.11 the tangent space T H is a null hyperplane. Since q A˜ we q ∈ δ have T N T H, so T N is also a null hyperplane. The normal vector n to the q = q q q null hyperplane Tq N is then null, and any two null vectors in a null hyperplane are parallel, so the vector n for a point q A˜ is parallel to any tangent vector of q ∈ δ the null geodesic generator through q. Once again, consider some point q A˜ N. Since β (q) is a point along the ∈ δ ⊆ t geodesic (with respect to the spacetime metric g) with initial velocity parallel to n (which by definition is nonzero) it holds that there is some function r : A˜ R q δ → such that β (q) expg (r (q)n ). t = q For each q N there is a unique real number rˆ(q) such that the σ-distance ∈ between q and expg (rˆ(q)n ) along the curve τ expg (τn ) is precisely t. By defi- q → q nition r and rˆ coincide on A˜δ. We now want to use Clarke’s implicit function theorem (Corollary, p. 256 in [11]) to conclude that rˆ is locally Lipschitz. By definition, the choice ξ rˆ solves = the equation Z 1 ¯¯ ¯¯ ¯¯ ∂ g ¯¯ q N ¯¯ exp (τξ(q)nq )¯¯ dτ 1, ∀ ∈ 0 ¯¯∂τ ¯¯σ = and this solution is of course unique. In other words ξ rˆ is the function satisfying = F (q,ξ(q)) 0 where F : N R R is defined by = × → µZ 1 ¯¯ ¯¯ ¶ ¯¯ ∂ g ¯¯ F (q,t) ¯¯ exp (τtnq )¯¯ dτ 1. = 0 ¯¯∂τ ¯¯σ − Since n is locally Lipschitz, so is F . Note that F has a partial derivative with respect to t and that ∂F 0 everywhere since n is nowhere zero. Clarke’s implicit function ∂t 6= theorem now tells us that there is a solution ξ of F (q,ξ(q)) 0 with ξ(p) r (p) = = which is Lipschitz in a neighborhood of p. Since we already know that the only solution of this equation is rˆ, this shows that rˆ is locally Lipschitz in some neigh- borhood U of p.

Since rˆ is Lipschitz on U, the function βct : N U M defined by ∩ → g βct (q) exp (rˆ(q)nq ) = 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 31 is also Lipschitz. The restriction of this function to A˜ U agrees with β , complet- δ ∩ t ing the proof.

For future reference, we note the following corollary.

Corollary 1.3.34. Let A˜δ be the full-density subset of a δ-flow set Aδ, and let U be any set such that the generator flow with respect to some Riemannian metric σ is defined on all of U A . Let hn be the n-dimensional Hausdorff measure associated ∩ δ to σ. Then β (A˜ U) has full hn-measure in β (A U). t δ ∩ t δ ∩ Proof. When U is contained in a sufficiently small open set, Lemma 1.3.33 tells us that βt is the restriction of a Lipschitz function. Hence βt (Aδ U)\βt (A˜δ U) has n n ∩ ∩ h -measure zero, since Aδ\A˜δ has h -measure zero. Otherwise, U may be covered by countably many such small open sets since it is second-countable, giving the same conclusion.

Proposition 1.3.35. Let H be a horizon in an (n 1)-dimensional spacetime (M,g) + equipped with a Riemannian metric σ of the form

σ(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ) for some timelike g-unit vector field V . = + Let hn be the n-dimensional Hausdorff measure associated to σ, let δ 0, let A be > δ the δ-flow set, let A˜ be the full-density subset of A , let t 0, let β be the restriction δ δ > t to Aδ of the generator flow with respect to σ, let U be a neighborhood of p which is open in M, suppose that β is defined on all of A U, let N U be a C 1,1 hypersur- t δ ∩ ⊆ face containing A and let βct : N M be a locally Lipschitz function which agrees δ → with βt on A˜ U. Suppose that θA 0 on all of HA . δ ∩ l ≤ l Then every p A has a neighborhood Z which is open in A such that there is ∈ δ δ a measurable function Ψ with Ψ 1 almost everywhere such that if ϕ: A R is ≥ δ → hn-integrable and supported in Z A˜ then ∩ δ Z Z n 1 n ϕΨdh ϕ(β−t (y))dh (y). A˜δ = βt (A˜δ)

Moreover, if Ψ 1 almost everywhere on Z , then θA 0 almost everywhere on Z . = l =

Remark 1.3.36. A sufficient condition for having θA 0 is that the generators of l ≤ H are complete in the past direction, together with the null energy condition. (See Proposition 1.3.23.)

Proof. Let Z A U. By hypothesis βct is Lipschitz on N. Theorem 3.1 of [12] tells = δ ∩ us that   Z Z n X n ϕJ(βct )dh  ϕ(x)dh (y). N = N 1 x βct − (y) ∈ 32 CHAPTER 1. LORENTZIAN COBORDISMS

Here J(βct ) is the Jacobian determinant of βct with respect to σ at points where βct is differentiable. Since βct is Lipschitz, J(βct ) is thus almost everywhere defined on N, which is sufficient for the integral to make sense.

Note that ϕ is zero outside of A˜δ, so X X ϕ(x) ϕ(x). 1 = 1 x βct − (y) x βct − (y) A˜ ∈ ∈ ∩ δ

Since βt is an bijection from Aδ to βt (Aδ), its restriction to A˜δ is injective. Its image is of course βt (A˜δ). This means that  1 ˜ 1 {β−t (y)} if y βt (Aδ), βct − (y) A˜δ ∈ ∩ =  otherwise. ; Hence Z Z n 1 n ϕJ(βct )dh ϕ(β−t (y))dh (y). A˜δ = βt (A˜δ)

To complete the proof of the theorem, we need to show that J(βct ) 1 almost ≥ everywhere on A˜ U, and that J(βct ) 1 almost everywhere on A˜ U only if δ ∩ = δ ∩ θA 0 almost everywhere on A U, possibly after shrinking U. We can then l = δ ∩ choose Ψ to be J(βct ). Since the argument for proving these statements is quite long, we prove them separately as Lemma 1.3.37.

Lemma 1.3.37. Fix t. Let p, U, Aδ, A˜δ, N and βbt be as in Proposition 1.3.35. After possibly shrinking U to a smaller neighborhood of p, it holds that J(βct ) 1 almost n ≥ everywhere (with respect to h ) on A U. Moreover, if J(βct ) 1 almost everywhere δ ∩ = on A U then θA 0 almost everywhere on A U. δ ∩ l = δ ∩

Proof. Proof part I(Construction of the set Bb): After possibly shrinking U to a smaller neighborhood of p and decomposing it as U (a,b) Σ (with the curves = × (a,b) {q} timelike and the slices {τ} Σ spacelike) where Σ is an open subset of × × Rn, we may view H as the graph of a semi-convex function f , as noted in Sec- tion 1.3.2.1. Similarly, N is the graph of a C 1,1 function g. Let Ln denote Lebesgue measure on Σ. By [13, Theorem 3.1.15], for each positive integer k there is a C 2 function g : Σ R such that k → Ln ¡{x Σ g (x) g(x)}¢ 1/k. ∈ | k 6= <

Let pr A˜δ be the projection of A˜δ on Σ. Let B be the full-density subset of pr A˜δ. n By Proposition A.4.7 the set B has full L -measure in pr A˜δ. Letting

B B {x Σ g (x) g(x)} k = ∩ ∈ | k = 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 33 we then have Ln(B \ B ) 1/k. k < Once again, we discard low-density points: Let B˜k be the full-density subset of Bk . Then B˜k has full measure in Bk by Proposition A.4.7 so

Ln(B \ B˜ ) 1/k. k <

Let ΣRad denote the points of Σ where g is twice differentiable in the sense that it has second order expansions of the form

1 2 2 g(x) g(x0) dg(x0)(x x0) D g(x0)(x x0,x x0) o( x x0 ), = + − + 2 − − + | − | (1.4) dg(x) dg(x ) D2g(x )(x x , ) o( x x ). = 0 + 0 − 0 · + | − 0| 1,1 Rademacher’s theorem tells us that since g is C , the set ΣRad has full measure in Σ. Defining

Bbk : B˜k ΣRad , = ∩ [ [ Bb : Bbk ΣRad B˜k = k N = ∩ k N ∈ ∈ n we then know that Bb has full L -measure in pr A˜δ. Since g is Lipschitz, the graph n of g over Bb has full h -measure in A˜δ U by Proposition A.4.3. It is now sufficient ∩ to show that J(βct )(p0) 1 whenever p0 (g(x0),x0) is such that x0 Bb. ≥ = ∈ Choose some x0 Bb. Since Bb pr A˜δ, the functions f and g agree at x0, and x0 ∈ ⊆ is an Alexandrov point of f . Hence we have the expansion

1 2 2 f (x) f (x0) d f (x0)(x x0) D f (x0)(x x0,x x0) o( x x0 ). = + − + 2 − − + | − |

Since x0 Bb, we know that g is twice differentiable at x0 so that we have the ex- ∈ pansions of equation (1.4). Moreover, x B for some i N so g (x ) g(x ). Fix 0 ∈ i ∈ i 0 = 0 this value of i for the remainder of the proof. Moreover, by definition of Bb, the point x0 is a full-density point of pr Aδ, and g and f agree on pr Aδ. Similarly, x0 is a full-density point of Bi , and g and gi agree on Bi . This means that we can use Proposition A.4.12 to conclude that

d f (x ) dg(x ) dg (x ) (1.5) 0 = 0 = i 0 and D2g (x ) D2g(x ) D2 f (x ). i 0 = 0 = 0

2 Proof part II(Construction of the C null hypersurface Hi ): Let Ni denote the 2 graph of gi over Σ. Let Nbi denote a C spacelike hypersurface in Ni containing 34 CHAPTER 1. LORENTZIAN COBORDISMS

p0. Thus Nbi has codimension 2 in the spacetime M. Let ni denote the past- directed σ-unit normal null vector field of Nbi such that ni (p0) is the σ-unit tan- gent of the (unique since p0 Aδ) generator of H passing through p0. Note that 1 ∈ ni is C . Let Gi be the union of the geodesics starting from Nbi with initial veloci- ties given by n . Let γ: [0,t] G denote the curve s β (p ). We wish to choose i → i 7→ s 0 a subset of G which is a C 2 hypersurface containing γ. Define exp: Ω M by i → exp(τ,q) exp (τni (q)), where Ω R Nbi is the largest subset on which exp may = q ⊆ × be defined. Proposition A.3 of [8] says that if exp is injective at (τ,q) then there is ∗ an open neighborhood O of (τ,q) such that exp(O) is a C 2 submanifold of M. This, together with the fact that exp is injective when restricted to [0,t] {p0}, shows 2 × that some neighborhood of γ in Gi is a C hypersurface in M. Hence we need to 2 show that exp is injective at (s,p0) for each s [0,t]. Note that Nbi is a C spacelike ∗ ∈ submanifold of M of codimension 2, and that Nbi is second order tangent to H at p0 in the sense of [8, Section 4.2] since D2g (x ) D2 f (x ). By [8, Lemma 4.15] there i 0 = 0 can then be no focal point of Nbi along γ. By [28, Proposition 10.30] this means that exp is injective at (s,p0) for all s [0,t]. As pointed out previously, [8, Proposition ∗ ∈ 2 A.3] then tells us that some open neighborhood of γ in Gi is a C submanifold. Since exp is injective at each point on γ, it is injective in a neighborhood of each ∗ such point. Since γ([0,t]) is compact, finitely many such neighborhoods suffice to 2 cover γ. Hence there is a neighborhood of γ where Gi is C and exp is injective. ∗ Denote this neighborhood by Hi . By an application of Proposition B.2.1 we see that Hi is a null hypersurface. i Proof part III(Definition of a map βct : Hi M): By definition of Hi , the vector → field ni (where defined) is tangent to Hi . Since Hi is a null hypersurface, it has a unique σ-unit normal null vector field, which must then be an extension of ni . Call this extension ni as well. Note that Hi contains both Nbi and the generator passing through p0, these being submanifolds of Ni transverse to each other. Hence the first and second derivatives of the graphing functions of Hi and Ni must agree at p0. i Define the map βct : Hi M by → i g βct (q) exp (r (q)ni ) = where r (q) is the unique nonnegative real number such that the σ-distance from g g q to exp (r (q)ni ) along the g-geodesic exp (τni ) is precisely t. Then by definition

i βct (p0) βct (p0). = i g Note that βct and βct are defined by the formula exp (r (q)k) where k is the normal g vector field of N and Hi , respectively. The derivative of exp (r (q)k) is determined by the first derivatives of r and k. These in turn are determined by the second 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 35

derivatives of the graphing functions of N and Hi . Since dg(p0) dgi (p0) and 2 2 i = D g(p0) D gi (p0) this means that the tangent maps of βct and βct agree at p0. = Hence i J(βct )(p0) J(βct )(p0). = i We have now reduced the problem to showing that J(βct )(p0) 1. ≥ i Proof part IV(Computation of J(βct )(p0)): Let bHi be the one-parameter family of Weingarten maps (defined in Section 1.3.1.1) along the generator of Hi through p0 with its affine parametrization, and let b0 denote the covariant derivative of b Hi along this generator. Equation (1.1) in Section 1.3.1.1 tells us that bHi satisfies the equation 2 b0 b Re 0. (1.6) + + = Recall from Remark 1.3.22 that H has a null Weingarten map bAl , defined in terms of Alexandrov derivatives, on all points to the past of p0 on the generator of H through p0, and that this map also satisfies equation (1.6). Since the null Wein- garten map of a null hypersurface can be expressed in the first two derivatives of a graphing function, and H shares these derivatives with Ni which in turns shares them with Hi . Hence bH (p0) bA (p0). By the uniqueness of solutions to the i = l ordinary differential equation (1.6), these two maps must agree on all of the past of p0 along the generator through p0. Let θHi denote the null mean curvature of Hi , as defined in Section 1.3.1.1. Then we have

θH trbH trbA θA . i = i = l = l Lemma 1.3.4 implies that

t i µ Z i ¶ J(βct )(p0) exp θHi (βcs (p0))ds . = − 0

i i Since J(βct )(p0) J(βct )(p0) and θA θH along the curve s βcs (p0) βcs(p0), = l = i 7→ = we then know that µ Z t ¶ J(βct )(p0) exp θAl (βcs(p0))ds . = − 0

n Recall that g(Bb) has full h -measure in N U and that p0 was an arbitrary point of ∩ g(Bb). Since we have assumed that θAl 0, we can conclude that J(βct ) 1 almost ≤ ≥ everywhere on the neighborhood U of p. This completes the proof of the first part of the lemma. To prove the last part of the lemma, some measure theoretical technicalities remain.

Claim V: J(βct ) 1 almost everywhere on A˜ U only if θA 0 almost everywhere = δ∩ l = on A˜ U δ ∩ 36 CHAPTER 1. LORENTZIAN COBORDISMS

After possibly shrinking U, we may choose C 1 coordinates on N U (a,b) Σ in ∩ =∼ × which

βct ((q,s)) (q,s t) for all q Σ and s,t such that s (a,b) and s t (a,b). = + ∈ ∈ + ∈ n Suppose that J(βct ) 1 almost everywhere on A˜ U with respect to h . Then in = δ ∩ particular (since g(Bb) has full measure in N U) it holds that J(βct ) 1 almost ∩ = everywhere on A˜δ U g(Bb). We have seen that on this set, with coordinates ∩ ∩ (q,s) (a,b) Σ, ∈ × µ Z t ¶ J(βct )((q,s)) exp θAl (βcs((q,s)))ds . = − 0

n If J(βct ) 1 almost everywhere, then it holds for h -almost every pair (q,s) that = θAl (q,s0) 0 for almost all s0 s with respect to Lebesgue measure. By shrinking U = > n further, we may assume that it holds for h -almost every q Σ that θA (q,s) 0 for ∈ l = almost all s 0 with respect to Lebesgue measure. Note that the measure hn differs > from the Lebesgue measure in coordinates of (a,b) Σ only by pointwise scaling by × a C 1 function. Note also that Lebesgue measure in coordinates on (a,b) Σ agrees × with the product of Lebesgue measure on Σ and Lebesgue measure on (a,b) (on sets which are measurable in both measures). Hence Fubini’s theorem (see [13, Theorem 2.6.2]) implies that Z n ¡© ª¢ 1 ¡ ¢ n 1 L (s,q) (a,b) Σ θAl (q,s) 0 L {s (a,b) θAl (q,s) 0 dL − (q) ∈ × | 6= = Σ ∈ | 6= Z n 1 0dL − (q) 0. = Σ =

n This shows that θA 0 on almost all of U A with respect to h , completing the l = ∩ δ proof.

Proposition 1.3.38. Let H be a horizon in an (n 1)-dimensional spacetime M + equipped with a Riemannian metric σ of the form

σ(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ) for some timelike g-unit vector field V . = + Let δ 0 and let A be the δ-flow set of H. Let hn denote the n-dimensional Haus- > δ dorff measure induced by σ. Let Hb be a past null geodesically ruled open subset of H. Suppose that θA 0 on all of HA Hb . Let t 0 be such that the generator flow l ≤ l ∩ > β is defined on all of Hb A . t ∩ δ Then hn(Hb A ) hn(β (Hb A )). ∩ δ = t ∩ δ Moreover, θA 0 almost everywhere on Hb A . l = ∩ δ 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 37

Proof. Note that the generators of Hb are the intersections of the generators of H with Hb since Hb is an open subset of the horizon H. For each p, let Zp and Ψp be the neighborhoods and functions given by Proposition 1.3.35. Since Aδ is second countable, a countable number of neighborhoods Z1, Z2,... suffice to cover Aδ. For each i 1, let ≥ [ Yi Zi \ Z j . = 1 j i ≤ < Then

• each Yi is measurable,

• Y Z for each i, i ⊆ i

• the Yi are pairwise disjoint, S S • i 1 Yi i 1 Zi Aδ. ≥ = ≥ ⊇ For each i 1, let ϕ be the indicator function of Y Hb A˜ . Then Proposi- ≥ i i ∩ ∩ δ tion 1.3.35 says that Z Z n 1 n ϕi Ψi dh ϕi (β−t (y))dh (y) A˜δ = βt (A˜δ) for each i 1. Since each ϕ is zero outside of Hb , this means that ≥ i Z Z n 1 n ϕi Ψi dh ϕi (β−t (y))dh (y) Hb A˜ = β (Hb A˜ ) ∩ δ t ∩ δ Taking a sum over i, we see that Z Z X n X 1 n ϕi Ψi dh ϕi (β−t (y))dh (y). ˜ ˜ Hb Aδ i 1 = βt (Hb Aδ) i 0 ∩ ≥ ∩ ≥

Since precisely one of the functions ϕi is nonzero at any point p Hb A˜δ, and P ∈ ∩ takes the value 1 there, we have i 1 ϕi Ψi 1 almost everywhere on Hb A˜δ. More- ≥ ≥ ∩ P 1 ˜ over, we have i 0 ϕi (β−t (y)) 1 almost everywhere on βt (Hb Aδ). Hence ≥ = ∩ Z n X n n h (Hb A˜δ) ϕi Ψi dh h (βt (Hb A˜δ)). ˜ ∩ ≤ Hb Aδ i 1 = ∩ ∩ ≥

Since A˜δ has full measure in Aδ and βt (A˜δ) has full measure in βt (Aδ) by Corol- lary 1.3.34, this means that

hn(Hb A ) hn(β (Hb A )). ∩ δ ≤ t ∩ δ However, β (Hb A ) Hb A since the generators of Hb agree with the generators t ∩ δ ⊆ ∩ δ of H, so we also know that

hn(Hb A ) hn(β (Hb A )) ∩ δ ≥ t ∩ δ 38 CHAPTER 1. LORENTZIAN COBORDISMS by additivity of the measure. Hence equality must hold, and the proof of the first statement is complete. P Equality can hold only if i 1 ϕi Ψi 1 almost everywhere on Hb Aδ. This ≥ = ∩ means that each function Ψ must be equal to 1 almost everywhere on Y Hb A . i i ∩ ∩ δ By Proposition 1.3.35 this implies that θA 0 almost everywhere on Yi Hb A . l = ∩ ∩ δ Since these sets cover Hb A , we have shown that θA 0 almost everywhere on ∩ δ l = Hb A with respect to the measure hn. This completes the proof. ∩ δ

1.3.5.4 Smoothness from area-preserving generator flow

Lemma 1.3.39. Let H be a horizon in a spacetime of dimension n 1 equipped with + a Riemannian metric σ and the corresponding Hausdorff measure hn. Let Ω be an open subset of H. Let Aδ denote the δ-flow set of H with respect to σ. Suppose that

hn(Ω A ) hn(β (Ω A )) ∩ δ = t ∩ δ for all t 0 and all δ 0, and that hn(Ω) . > > < ∞ Then the following two statements hold.

1. There is a dense subset of Ω consisting of points on maximal null geodesics contained in Ω.

2. No generator of H has any endpoint on Ω.

Proof. Let Aδ denote the δ-flow set of H with respect to σ. Recall that the total flow set of H is the set [ A0 Aδ. = δ 0 > Note that if δ δ0 then A A . Hence < δ ⊇ δ0 [ [ A0 Aδ A1/k . = δ 0 = k Z > ∈ + Since the family Ω A is increasing, ∩ 1/k n n n n h (βt (Ω A0)) lim h (βt (Ω A1/k )) lim h (Ω A1/k ) h (Ω A0). ∩ = k ∩ = k ∩ = ∩ →∞ →∞ The limits are finite, since by hypothesis hn(Ω A ) is uniformly bounded with ∩ 1/k respect to k by hn(Ω). Introduce the sets C and D defined by \ C βt (Ω A0), = t Z ∩ ∈ + D {p Ω there is a unique generator through p, = ∈ | and this generator has no future endpoint}. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 39

We first show that if p C then the generator Γ through p is a maximal geodesic ∈ p contained in Ω. By choice of Ω, the part of the generator to the past of p belongs to Ω, and is maximal in the past direction. Parameterize Γp by an affine parameter such that Γ (0) p. Suppose that the maximal future extension of the genera- p = tor were to leave Ω at some point q Γ (s). Since Γ is smooth and the interval = p p [0,s] is compact, the curve segment Γp ([0,s]) has finite length in the Riemannian metric σ. This means that p β (Ω A ) whenever t is greater than this length, ∉ t ∩ 0 contradicting the assumption that p C. This shows that the set C satisfies the ∈ conditions for the first statement in the conclusion, so that it is sufficient to show that C is dense to complete the proof of that statement. Note that the fact that generators through points of C are maximal means that they can have no endpoints. Hence C D. Since (βt (Ω A0))∞t 1 is a countable ⊆ ∩ = decreasing family of sets of equal measure it holds that hn(C) hn(Ω A ). Since = ∩ 0 A H, and hn(H \ A ) 0 by Proposition 1.3.29, this means that hn(Ω) hn(Ω 0 ⊆ 0 = = ∩ A0) so that hn(C) hn(Ω). = In particular, C is dense in Ω. Since C D, it follows that D is also dense in Ω. ⊆ We will now show that D is closed in Ω. Suppose that a sequence (pk )k N in D ∈ converges to p Ω. Let X denote the (unique, by definition of D) future-directed ∈ k σ-unit tangent of a generator at pk . The σ-unit over the com- pact countable set {p} {p ,p ,...} is compact, so by passing to a subsequence we ∪ 1 2 may assume that the Xk converge to some unit vector X at p. By [8, Lemma 6.4] the space of past-directed σ-unit generator tangents is closed in the unit tangent bundle, so we know that X is tangent to a generator. Let γ denote the inextendible geodesic with initial velocity X , and let γk denote the inextendible geodesic with initial velocity Xk . By definition of D, each geodesic γk avoids the open set I +(H). By continuous dependence on initial conditions for ordinary differential equa- tions, γ must also avoid the open set I +(H). Suppose to get a contradiction that γ leaves H at some point p. Choose coordinates around p of the form ( a,a) Σ, − × such that each curve ( a,a) {q} is timelike and each slice {t} Σ is spacelike. Let − × × Σ0 denote the projection of imγ (( a,a) Σ) on Σ. Recall that H is an achronal ∩ − × hypersurface, so by possibly shrinking Σ we may represent H (( a,a) Σ0) as ∩ − × the graph of a function fH : Σ0 ( a,a). Let fγ : Σ0 ( a,a) be the function the → − → − graph of which is the image of γ (on both sides of the point p). Recall that γ is a null curve, so if fH(x) fγ(x) at some point x Σ0 then there is a timelike curve > ∈ from (fH(x),x) to p, contradicting achronality of H. If fH(x) fγ(x) at some point < x Σ0 then γ intersects I +(H) which we saw earlier is impossible. Hence γ can- ∈ not leave H to the future, and so there is a generator through p without future endpoint. Moreover, p is then an interior point of a generator so this generator is 40 CHAPTER 1. LORENTZIAN COBORDISMS unique. Hence p D and we have shown that D is closed in Ω. ∈ We have now shown that D is a closed dense subset of Ω. Hence D Ω. Since = no point in D lies on a generator with a future endpoint, no point in D can be a future endpoint of a generator. This shows that no generator of H can have a future endpoint on Ω. Recall that no generator has any past endpoint either, since H is a horizon. This completes the proof.

The condition of H containing no endpoints is very strong, as the next theo- rem (see [8, Theorem 6.18]) illustrates.

Theorem 1.3.40. Suppose that Ω is an open subset of a horizon H in a spacetime M, such that Ω contains no endpoints of generators of H. Suppose moreover that n θA 0 almost everywhere with respect to the n-dimensional Hausdorff measure h l = induced by a Riemannian metric σ on M. Then Ω is a smooth submanifold of M.

We are now in a position to prove the main theorem of this section. It was shown in [6, Section 4] that not all compact horizons are smooth. Our theorem shows that the additional hypothesis of the null energy condition is sufficient to guarantee smoothness.

Theorem 1.3.41. Let M be a spacetime of dimension n 1 satisfying the null energy + condition. Let S M be an achronal set with edge(S) . Let Hb be an open subset ⊂ = ; of H +(S) with compact closure. Suppose that Hb is past null geodesically ruled. Then Hb is a smooth totally geodesic null hypersurface.

Proof. First note that Corollary 1.3.14 tells us that H +(S) is a horizon in the sense of Definition 1.3.7. Further note that Hb is a Lipschitz hypersurface, since it is an open subset of the Lipschitz hypersurface H +(S). It is also assumed to be past null geodesically ruled, so the generators of Hb are the intersection of the generators of H +(S) with Hb . Note that Alexandrov points of H +(S) are Alexandrov points of Hb . Each generator of Hb is a part of a null geodesic contained in H +(S). By definition, each generator of Hb is completely contained in, and hence totally past imprisoned in, the compact set Hb . This means that the generator flow βt is defined for all t on all of the part of the total flow set of H +(S) which lies in Hb . Moreover, Lemma B.1.1 shows that each generator of Hb is complete in the past direction. Since the null energy condition holds and we have shown that the generators of Hb are complete to the past, Proposition 1.3.23 tells us that θA 0 at all Alexandrov points of Hb . l ≤ Let V be an arbitrary timelike vector field on M, introduce the Riemannian metric σ on M defined by

σ(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ), = + 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 41

and let hn be the corresponding n-dimensional Hausdorff measure. By Proposi- tion A.4.4, the set Hb and all its measurable subsets have finite hn-measure since

Hb is compact. For each δ 0, let A denote the δ-flow set of H +(S). By Proposi- > δ tion 1.3.38 we know that

hn(Hb A ) hn(β (Hb A )) ∩ δ = t ∩ δ and θA 0 almost everywhere on Hb A l = ∩ δ

for all t 0 and all δ 0. Lemma 1.3.39 then tells us that no generator of H +(S) > > S has any endpoint on Hb . Moreover, since the total flow set A0 δ 0 Aδ has full n = > h -measure by Proposition 1.3.29 we see that θA 0 almost everywhere on Hb . l = Theorem 1.3.40 then says that Hb is a smooth submanifold of M. Since Hb is an open subset of H +(S) and the tangent space of H +(S) is a null hyperplane when- ever it exists, Hb is a null hypersurface. Let K be the tangent vector field of the generators of Hb , scaled to an affine parametrization. Since Hb is smooth, its null mean curvature θ with respect to K is a smooth function and its sign agrees with that of the Alexandrov null mean cur- vature θA . We saw previously that θA 0 almost everywhere, so by continuity l l = θ 0 everywhere. Let b denote the null Weingarten map with respect to K , and = let S b θ . Since S is self-adjoint, tr(S2) 0. Since θ 0 and Ric(K ,K ) 0 ev- = − n 2 ≥ = ≥ erywhere, equation− (1.2) tells us that tr(S2) 0. Since S is self-adjoint this implies = that S 0. Hence b 0 everywhere, so that Hb has vanishing null second funda- = = mental form. Theorem 1.3.1 then implies that Hb is totally geodesic, completing the proof.

The following corollary is immediate.

Corollary 1.3.42. Let M be a spacetime satisfying the null energy condition. Let

S M be an achronal set with edge(S) . Suppose that H +(S) is compact. Then ⊂ = ; H +(S) is a smooth totally geodesic null hypersurface.

Proof. Let Hb H +(S) and apply Theorem 1.3.41. =

1.3.6 Cobordisms satisfying energy conditions are trivial

In this section, we will apply Theorem 1.3.41 to prove Tipler’s theorem.

1.3.6.1 Cylinder extensions

We begin by describing a construction embedding a Lorentzian cobordism M into a manifold without boundary Ξ(M). Informally, we will glue two cylinders onto M. 42 CHAPTER 1. LORENTZIAN COBORDISMS

This construction is a technicality needed only to apply theorems about manifolds without boundary to a cobordism.

Definition 1.3.43. Let M be a Lorentzian cobordism between S1 and S2. Let

M [0, ) S . i = ∞ × i The cylinder extension of M is the spacetime without boundary

Ξ(M) M M M / = 1 t t 2 ∼ where is the smallest equivalence relation such that ∼ (0,p) p for all p S and (0,p) M . ∼ ∈ i ∈ i The metric of Ξ(M) is taken to be an arbitrary smooth extension of the metric on M.

Remark 1.3.44. Note that adding cylinders in this manner does not affect the causal relations of points in the original manifold M, since the time-orientation of the boundary does not allow causal curves to pass from M2 to M or from M to M1.

Lemma 1.3.45. Let M be a Lorentzian cobordism between S1 and S2. Then S1 is a closed and achronal subset of Ξ(M).

Proof. Let M denote the open submanifold (0, ) S of Ξ(M) and let M˚ de- i ∞ × i note the interior of M. The set S1 is a closed subset of Ξ(M) by the definition of the quotient topology on Ξ(M), since its inverse image under the projection map M M M Ξ(M) is the closed set {0} S S . 1 t t 2 → × 1 t 1 To see that S is achronal, suppose that γ:[a,b] Ξ(M) is a future-directed 1 → timelike curve (with a b) such that γ(a),γ(b) S1. Since S1 is closed, so is the 1 1 < ∈ set γ− (S1). If γ− (S1) were dense in [a,b], then by continuity γ would be com- pletely contained in S1. This is impossible since S1 is spacelike and γ timelike. Hence there is some maximal open interval J (a,b) such that γ(J) S . By ⊆ ∩ 1 = ; replacing a and b with inf(J) and sup(J), we see that we may without loss of gen- erality assume that γ(t) S for t (a,b). Since γ˙(a) is future-directed, it must be ∉ 1 ∈ inward-directed as a vector at the boundary S of M. Hence γ(a ²) M for all 1 + ∉ 1 sufficiently small ² 0. Since the boundary of M in Ξ(M) is S , continuity of γ > 1 1 implies that γ(t) M for all t (a,b). However, γ˙(b) is also future-directed and ∉ 1 ∈ hence inward-directed with respect to M, and so γ(b ²) M for all sufficiently − ∈ 1 small ² 0. This is a contradiction, so γ cannot intersect S more than once. Hence > 1 S1 is achronal. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 43

1.3.6.2 Tipler’s theorem without smoothness hypothesis

We will prove the generalization of Tipler’s theorem, suggested in [4], to causally compact Lorentzian pseudocobordisms.

Theorem 1.3.46. Let n 2. Let S ,S be n-dimensional manifolds and let (M,g) be ≥ 1 2 a causally compact Lorentzian pseudocobordism between S1 and S2 which satisfies the lightlike convergence condition and the lightlike generic condition.

Then M is globally hyperbolic. In particular M S1 [0,1] so that S1 and S2 are =∼ × diffeomorphic.

Proof. We will work with the cylinder extension Ξ(M) of M, since we need a man- ifold without boundary. Let M for i {1,2} denote the embedded submanifolds i ∈ (0, ) S of Ξ(M). Then Ξ(M) is the union of M ,S ,M˚ ,S and M , and these ∞ × i 1 1 2 2 sets are disjoint. Consider the future Cauchy horizon H +(S1) of S1 in Ξ(M). If it is empty, then S would be a Cauchy surface for M M and hence also for M. 1 ∪ 2 This would mean that M is globally hyperbolic. Hence it is sufficient to show that

H +(S1) is empty. Suppose for contradiction that this is not the case.

Claim I: H +(S1) is a horizon S1 is a smooth hypersurface in Ξ(M). Lemma 1.3.45 tells us that S1 is closed and achronal. Hence we can use Lemma 1.3.15 and Corollary 1.3.14 to conclude that

H +(S1) is a horizon in the sense of Definition 1.3.7. Choose a point p I +(H +(S ),M). To see that such a point exists, note that ∈ 1 I +(q,M) is nonempty if q S , and that H +(S )\ S is nonempty since H +(S ) is ∉ 2 1 2 1 past null geodesically ruled and hence cannot be completely contained in a space- like hypersurface. With this choice, H +(S ) I −(p,M) is nonempty. 1 ∩ Claim II: Every generator of H +(S1) which intersects I −(p,M) stays in I −(p,M) when followed to the past

Let γ be a generator of H +(S ). We will show that if γ(t) I −(p,M) for some t 1 ∈ then γ(( ,t]) I −(p,M). This will mean that γ is totally past imprisoned in −∞ ⊆ the set I −(p,M) which is compact since M is causally compact. To this end, let γ:(a,t] Ξ(M) be the maximal past geodesic extension of a generator of H +(S ), → 1 with γ(t) I −(p). By [28, Proposition 14.53], H +(S ) and S are disjoint, so γ does ∈ 1 1 not intersect S1. Moreover, when followed to the past γ cannot intersect S2 since it would need to do so with the wrong time orientation. Hence γ stays in M. Let s t. Then there is a causal curve in M from γ(s) to p formed by concatenating < γ with a timelike curve from γ(t) to p. Such a curve exists since we assumed that

γ(t) I −(p,M). Since this curve is not everywhere lightlike, there is a timelike ∈ curve from γ(s) to p. Hence γ(s) I −(p), proving the claim. ∈ Claim III: H : H +(S ) I −(p,M) is past null geodesically ruled = 1 ∩ 44 CHAPTER 1. LORENTZIAN COBORDISMS

Let H denote the set H +(S ) I −(p,M). To find a past complete null geodesic 1 ∩ segment through a point q H, consider the intersection of H and the generator ∈ of H +(S1) through q. This curve is a geodesic segment, and it is connected and past complete by the previous claim. Hence H is past null geodesically ruled.

Claim IV: The existence of H is contradictory

Since I −(p,M) is open, the set H is an open subset of H +(S1). Since M is causally compact, the set I −(p,M) is compact. Hence H is contained in a compact set and so has compact closure. Theorem 1.3.41 tells us that H is a totally geodesic smooth null hypersurface. The inequality (1.3) from Section 1.3.1.1 then reads

Ric(K ,K ) 0 ≤ for all null tangent vectors K to H. Combining this with the lightlike convergence condition, we can conclude that

Ric(K ,K ) 0. = We will now derive a contradiction from this. If the spacetime were to satisfy the strict lightlike convergence condition, then the contradiction is immediate. When we assume that the lightlike generic con- dition holds, a further argument is needed. By Lemma 1.3.39 there is a dense, and in particular nonempty, subset of H consisting of points on maximal null geodesics which are contained in H. Choose one such maximal geodesic γ. Let b denote the null Weingarten map with respect to a null vector field K which agrees with the tangent vector field of γ with an affine parametrization. Since H is to- tally geodesic, b 0 along γ. Equation (1.1) then implies that R(X ,K )K is par- = allel to the null vector K for every vector X , where R denotes the curvature ten- sor of the spacetime. Given a vector X , denote by λX the real number such that R(X ,K )K λ K . Let k g(K , ) denote the covector dual to K . Consider the ten- = X = · sor G k( )g(R( ,K )K , )k( ). Evaluating G on (X ,Y , Z ,W ) yields = · · · · k(X )g(R(Y ,K )K , Z )k(W ) k(X )g(λ K , Z )k(W ) k(X )λ k(Z )k(W ). = Y = Y From this we see that the antisymmetrization of G(X ,Y , Z ,W ) in Z and W is zero. Further antisymmetrizing in X and Y then also yields zero. In coordinates, this tells us that K e K f K R K 0 [a b]e f [c d] = on γ. However, the lightlike generic condition says that this antisymmetrized ten- sor is nonzero at some point along each maximal lightlike geodesic. We have now obtained a contradiction, so the assumption that H +(S1) is nonempty must be false. Hence M is globally hyperbolic. 1.3. LORENTZIAN COBORDISMS AND ENERGY CONDITIONS 45

The above theorem is in one sense the strongest result one may hope for: The null energy condition is the weakest of the commonly used energy conditions, and in the setting of cobordisms it implies global hyperbolicity, which is the strongest of the commonly used causality conditions. What we have proved is that topologically nontrivial Lorentzian cobordism cannot satisfy certain energy conditions. This is in sharp contrast to the result by Nardmann [26, Theorem 8.4] which allows the construction of Lorentzian metrics satisfying even stronger energy conditions on a large class of topologically non- trivial cobordisms. The difference is that the metrics so constructed only make the spacetime into a weak Lorentzian cobordism. This means that the notions of Lorentzian cobordism and weak Lorentzian cobordism are very different when one considers energy conditions. For a further discussion of this point, see [26, Section 9]. Note that a weak Lorentzian cobordism may be transformed into a Lorentzian cobordism by a homotopy of the metric making the light cones nar- rower. The theorems of Tipler (Theorem 1.3.46) and Nardmann ([26, Theorem 8.4]) then show that the imposed energy conditions must cease to hold at some point during this homotopy. Generic metrics satisfy the generic 2condition

In singularity theorems (for instance Theorem 2, Chapter 8, p. 266 of [19]) a con- dition called the generic condition is imposed on the spacetime. This is the con- dition that for each timelike or null geodesic γ there is some point γ(t) at which γ˙e γ˙ f γ˙ R γ˙ 0. (See Definition A.3.5.) We follow the terminology of [1] in [a b]e f [c d] 6= saying that a vector K for which K e K f K R K 0 is generic. [a b]e f [c d] 6= It has been proposed (see for instance [19] p. 101 and [4] p. 21) that this ”generic condition” should in some sense be ”generic” among Lorentzian met- rics on a given manifold. As pointed out by Lerner [23], the condition is generic in the set of metrics which also satisfies the strong energy condition. We will show that it is generic in the class of all metrics. The result we are going to prove is The- orem 2.6.3, which states the following.

Theorem 2.6.3. Let M be a smooth manifold of dimension n 4, possibly with ≥ boundary. Let L denote the bundle of Lorentzian metrics on M and endow the set r Γ∞(L) of smooth sections with the Whitney C topology for any 4 r . Then ≤ ≤ ∞ there is a residual (and hence dense) set G Γ∞(L) such that if g G and γ is a e f ⊆ ∈ geodesic then the points where γ˙ γ˙ γ˙[aRb]e f [c γ˙d] is nonzero along γ form a dense set. In particular,

• the lightlike generic condition holds for each metric g G, ∈ • the timelike generic condition holds for each metric g G, ∈ • the spacelike generic condition holds for each metric g G. ∈ An outline of the proof is provided in Section 2.1. The remaining sections con- tain a more detailed version of the proof. The method of proof is an adaptation of the method used by Rendall in [30].

2.1 Outline of the argument

In this section, we will outline the idea of the proof of Theorem 2.6.3, ignoring several technical complications. 2.1. OUTLINE OF THE ARGUMENT 47

Given a curvature tensor R, define F : TM T 4(TM) by → F (K ) : K e K f K R K . abcd = [a b]e f [c d] We wish to show that the complement of the zero set of the function t F (γ˙(t)) 7→ is dense for each geodesic γ. This means that it will be bad to have a large set of vectors for which the function vanishes to both zeroth and first order, i.e. vectors X which are zeros of F and for which the derivative of t F (γ˙(t)) along the geodesic 7→ γ with initial velocity X is also zero. For if X does not belong to this set, then either X or some vector close to X on the same geodesic is generic. Our task is then to find metrics where this set of bad vectors is as small as possible, preferably empty. It is fruitful to turn the question around. Given a vector X , which metrics are the ones for which X is a bad vector? (We outline here the argument only for non- lightlike vectors; the case of lightlike vectors is marginally more complicated.) By a computation by Beem and Harris [1] the condition that a non-lightlike vector Xp is nongeneric for some certain metric g is equivalent to that the curvature tensor R of g satisfies that

Rb(Ap Xp ),Bp Xp 0 for all vectors Ap and Bp . 〈 ∧ ∧ 〉 = 2 V2 Here Rb is the tensor in S ( (TM)) given by the curvature tensor. (See Section 2.3 for details.) The condition that the derivative of F along the geodesic with initial velocity X is zero reduces to

X R(Ap Xp ),Bp Xp 0 for all vectors Ap and Bp . 〈∇ ∧ ∧ 〉 =

Again, X R is a tensor constructed from X R as described in Section 2.3. The first ∇ ∇ equation determines a submanifold of codimension n(n 1)/2 of the space R0 of − tensors with some of the symmetries of curvature tensors. The second equation determines a submanifold of codimension n(n 1)/2 of the space R1 of tensors − with some of the symmetries of covariant derivatives of curvature tensors. This means that the set of pairs (R, R) induced by metrics which are bad for X is a ∇ codimension n(n 1) submanifold of R0 R1. − ⊕ The previous paragraph considered a single vector. What we want is an argu- ment which takes care of all vectors at once. We want to consider the set of pairs (g, X ) where X is a bad vector for g. As an intermediate step, we consider the subset of TM R0 R1 consisting of triples (X ,R, R) satisfying the above equa- ⊕ ⊕ ∇ tions, i.e. such that X is bad for the curvature tensor R with derivative R. This set ∇ is also a submanifold of codimension n(n 1). Denote this submanifold NG (for − ”non-generic”). We wish to use this set to determine which the bad pairs (g, X ) are. To determine if X is bad for g, what is needed is enough information about g to compute the value of the curvature tensor and its first covariant derivative 48 CHAPTER 2. THE GENERIC CONDITION

at the basepoint of X . In other words, the derivatives of g up to third order are needed. This information is encoded in the bundle J 3L of 3-jets of Lorentzian metrics. What we want to find is then the subset of J 3L TM consisting of pairs ⊕ (j 3g, X ) such that (X ,R, R) NG, where R and R are the pointwise curvature ∇ ∈ ∇ tensor and its derivative computed from g at the basepoint of X . To this end, let R0 be the map computing R from j 3g, and R1 the map computing R from j 3g. 0 1 1 3 ∇ Then W : (idTM ,R ,R )− (NG) is the set of bad pairs (j g, X ). It turns out that = 0 1 the map (idTM ,R ,R ) is a submersion, so W is a submanifold of the same codi- mension as NG, i.e. of codimension n(n 1). − Now (j 3g, X ) W means that (g, X ) is a bad pair, so if we can arrange so that x ∈ N W where ∩ = ; N ©(j 3g, X ) J 3L TM x M, X T M \ {0}ª, = x ∈ ⊕ | ∈ ∈ x then there would be no bad vectors for g at all. Now N has dimension 2n and W has codimension n(n 1). If n(n 1) 2n, i.e. if n 3, then saying that the − − > > intersection N W is empty is the same thing as saying that it is transverse. More ∩ specifically, it is the same thing as saying that the map {X TM X 0} J 3L x ∈ | x 6= → ⊕ TM defined by X (j 3g, X ) is transverse to W . x 7→ x x Finally, an application of a transversality theorem tells us that the set of such

metrics is dense in the set Γ∞(L) of all metrics, completing the proof. See Sec- tion 2.6 for details. In the outline just given, we have ignored many technical complications which are treated in more detail in the following sections.

2.2 Construction of the bundles Ri and Bi

Given a real inner product space V , let V2 V denote the vector space of skew- symmetric bivectors constructed from V with the inner product

A B,C D A,C B,D A,D B,C . 〈 ∧ ∧ 〉 = 〈 〉〈 〉 − 〈 〉〈 〉 Let S2(V2 V ) denote the set of symmetric 2-tensors on V2 V . Clearly, S2(V2 V ) is a subspace of the space of (2,0) tensors on V2 V . Let M be a manifold. We will define four tensor bundles over M. The first two are R0 and R1. We will think of R0 as a bundle of tensors with some of the symmetries of Riemannian curvature tensors R. Similarly, we will think of R1 as a bundle of tensors with some of the symmetries of the covariant derivative R of a ∇ curvature tensor. Explicitly, we use the following definition.

Definition 2.2.1. Let M be a smooth manifold. Then R0 and R1 are defined by ³^2 ´ R0 S2 TM , = 2.3. VIEWING A CURVATURE TENSOR R AS AN ENDOMORPHISM Rb 49

1 2 ³^2 ´ R S TM T ∗M. = ⊗ We will now construct bundles B0 and B1 with the help of two maps b0 and b1.

Definition 2.2.2. Define a map b0 : R0 V4 TM (in each fiber) by antisymmetriza- → tion over the last three components:

b0(R) R . abcd = a[bcd] 1 1 V4 5 Similarly, define a map b : R ( TM T ∗M) T (TM) in each fiber by → ⊗ ⊕ b1(R) (R ,R ). abcde = a[bcd]e ab[cde] Let 0 0 1 B (b )− ({0}) = and 1 1 1 B (b )− ({0}). = Remark 2.2.3. The idea is that the zero set B0 of b0 consists of those tensors which satisfy the first Bianchi identity. Similarly, the zero set B1 of b1 is the set of tensors which satisfy both the first and second Bianchi identities.

Remark 2.2.4. The maps b0 and b1 are vector bundle maps of constant rank, so B0 and B1 are vector bundles over M.

Remark 2.2.5. We will denote elements of B1 by S. This notation is purely for- ∇ mal; the does not signify an actual covariant derivative. The purpose of the no- ∇ tation is to remind us that S is a tensor with the same symmetries as a covariant ∇ derivative of a curvature tensor. Given a vector X , we will let S denote the ten- ∇X sor S X e , where S are the components of S. Note that S B0. abcde abcde ∇ ∇X ∈

2.3 Viewing a curvature tensor R as an endomorphism Rb

In this section, we will describe a way of interpreting a tensor R B0 as an endo- ∈ morphism of V2(TM). The isomorphism used for this will depend on a choice of metric. Moreover, given a vector X T M and a tensor S B1, the tensor S is ∈ p ∇ ∈ ∇X an element of B0 and as such may be viewed as an endomorphism of V2(TM) in the same way.

Definition 2.3.1. Given a metric g, define a map g : B0 End(V2(TM)) by letting b → gb(R) be the endomorphism such that

g(R)(A B),C D R(A,B,C,D) for all A,B,C,D T M and all p M. 〈b ∧ ∧ 〉 = ∈ p ∈ 50 CHAPTER 2. THE GENERIC CONDITION

When the metric is implicitly understood, we will use the notation Rb : gb(R). For 1 = S B and given a vector field X , we will use the notation X S g( X S). ∇ ∈ ∇ = b ∇

2.4 Construction of NGi,t and NGi,`

2.4.1 Definitions

Let PTM denote the projectivization of the tangent bundle TM of M. We will now construct certain subspaces of L PTM B0 B1. We will denote an element of ⊕ ⊕ ⊕ L PTM B0 B1 by (g ,[X ],R, S), i.e. including the basepoint p only in the ⊕ ⊕ ⊕ p ∇ notation for g . We emphasize again that the notation S is purely formal. p ∇ The following are the spaces we are interested in:

Definition 2.4.1. Let NG0,t L PTM B0 B1 be the set of (g ,[X ],R, S) such ⊆ ⊕ ⊕ ⊕ p ∇ that

• g (X , X ) 0, p <

• Rb(A X ),B X 0 for all A,B Tp M. 〈 ∧ ∧ 〉 = ∈ Definition 2.4.2. Let NG1,t L PTM B0 B1 be the set of (g ,[X ],R, S) such ⊆ ⊕ ⊕ ⊕ p ∇ that

• g (X , X ) 0, p <

• (ƒX S)(A X ),B X 0 for all A,B Tp M. 〈 ∇ ∧ ∧ 〉 = ∈ Definition 2.4.3. Let NG0,` L PTM B0 B1 be the set of (g ,[X ],R, S) such ⊆ ⊕ ⊕ ⊕ p ∇ that

• g (X , X ) 0, p =

• Rb(A X ),B X 0 for all A,B Tp M orthogonal to X (with respect to gp ). 〈 ∧ ∧ 〉 = ∈ Definition 2.4.4. Let NG1,t L PTM B0 B1 be the set of (g ,[X ],R, S) such ⊆ ⊕ ⊕ ⊕ p ∇ that

• g (X , X ) 0, p =

• (ƒX S)(A X ),B X 0 for all A,B Tp M orthogonal to X (with respect to 〈 ∇ ∧ ∧ 〉 = ∈ gp ).

Some words about these definitions are in order. First note that the conditions g (X , X ) 0 and g (X , X ) 0 are independent of the choice of representative X p < p = of [X ] PTM. Similarly, the conditions of the forms Rb(A X ),B X 0 and ∈ 〈 ∧ ∧ 〉 = 2.4. CONSTRUCTION OF NGI,T AND NGI,` 51

X S(A X ),B X 0 are also independent of the choice of representative. Note 〈∇ ∧ ∧ 〉 = also that an element of L PTM B0 B1 specifies a metric so that the construc- ⊕ ⊕ ⊕ tion of Rb and X S as in Section 2.3 is well-defined. ∇

2.4.2 Manifoldness and codimension

We now wish to show that the spaces defined above are submanifolds, and com- pute their codimensions.

Lemma 2.4.5. NG0,t is a submanifold of L PTM B0 B1 of codimension n(n ⊕ ⊕ ⊕ − 1)/2.

Proof. It is sufficient to show this locally, in a local trivialization of the bundle L PTM B0 B1. We will choose a suitable such trivialization shortly. ⊕ ⊕ ⊕ First let TLV ©(g,[X ]) L PTM [X ] is timelike for gª. = ∈ ⊕ | This is a subbundle of L PTM, of codimension 0, so it is sufficient to show that ⊕ NG0,t is a submanifold of TLV B0 B1 of the desired codimension n(n 1)/2. ⊕ ⊕ − Note that by definition, NG0,t TLV B0 B1. ⊆ ⊕ ⊕ Proof part I(Choice of trivialization): Choose an arbitrary point p M, and let ∈ U be an open neighborhood of p such that the bundles TLV,B0,B1 and TM are all trivializable over U. Let Ψ: U TLV TLV be an arbitrary local trivializa- × p → tion over U of the bundle TLV. We wish to extend this to a local trivialization of the full bundle in a clever way. Since B0 and B1 are vector subbundles of tensor bundles on TM, extending Ψ to a trivialization of TLV B0 B1 may be done ⊕ ⊕ by, for each (g,[X ]) TLV , choosing a frame V g,[X ] on U, and expressing each ∈ p fiber in the basis supplied by this frame. The frame V g,[X ] must depend smoothly on (g,[X ]) for the extended trivialization to be smooth. We propose to choose V g,[X ] {X , X ,... X } in the following manner: = 1 2 n • At the point p, choose representatives X [X ] for each [X ], depending n ∈ smoothly on [X ].

• At the point p, choose (smoothly in [X ]) n 1 additional vectors Y1,...,Yn 1 − − which together with Xn form a basis.

• Extend these vectors to frames, by use of a local trivialization of TM over U.

• Perform the Gram-Schmidt orthogonalization procedure to obtain the re- quired orthogonality relations. This algorithm is smooth in the components

of the metric g and in the starting vector Xn. 52 CHAPTER 2. THE GENERIC CONDITION

Using this frame, we can form the ordered frame for V2 TM given by

X1 X2; X1 X3, X2 X3;...; X1 Xn, X2 Xn,... Xn 1 Xn. ∧ ∧ ∧ ∧ ∧ − ∧ For our purposes, the frame needs to be ordered so that the last n 1 basis vectors − are X X for 1 i n. i ∧ n ≤ < Proof part II(Concrete description of NG0,t in the chosen frame): We now ex- press the condition Rb(A X ),B X 0 in the basis provided by our local trivial- 〈 ∧ ∧ 〉 = ization, where we have fixed p U and (gp ,[X ]) TLVp . Since Rb is a self-adjoint ∈ ∈ linear operator on V2 TM, it can be represented by a symmetric matrix in the cho- sen basis. We write this on block form where the lower right-hand block has size (n 1) (n 1): − × − Ã ! AB BC

0,t The condition defining NG (that Rb(A X ),B X 0 for all A,B Tp M) is equiv- 〈 ∧ ∧ 〉 = ∈ alent to the condition that Rb(Xi X ), Xi X 0 for all i and j. We ordered the 〈 ∧ ∧ 〉 = basis of V2 TM so that the vectors X X are the last n 1 vectors in the basis. i ∧ − Hence the condition is equivalent to C 0, when Rb is expressed in this basis. = Proof part III(Definition of a map c): For each 0 i j n, we define a map ¯ ≤ ≤ < c : TLV B0 B1¯ R by letting c (g ,[X ],R, S) be the (i, j) component of i,j ⊕ ⊕ U → i,j p ∇ the part C of the matrix representation of Rb at the point p using the basis from the fixed frame at p. In other words, let

ci,j (gp ,[X ],R, S) Rb(Xi Xn), X j Xn for 0 i j n. ∇ = 〈 ∧ ∧ 〉 ≤ ≤ < g,[X ] Note that we have fixed a frame V so that these functions ci j are well-defined. In particular, they depend up to a multiplicative scaling on the representative

Xn [X ] chosen, and we have already fixed such a representative. Define the map ∈ 0 1¯ n(n 1)/2 0,t ¯ c : TLV B B ¯U R − by c (c1,1,c1,2,...,cn 1,n 1). Clearly, NG ¯U is 1 ⊕ ⊕ → = − − 0,t ¯ c− (0). If we can show that c is a submersion, then we know that NG ¯U is a codimension n(n 1)/2 submanifold of TLV B0 B1, since {0} is a codimension − n(n 1)/2 ⊕ ⊕ n(n 1)/2 submanifold of R − . − Proof part IV(The map c is a submersion): For c to be a submersion at some point

(gp ,[X ]p ,Rp , Sp ), it is sufficient that the restricted map c {g } {[X ] } B 0 { S } is a ∇ | p × p × p × ∇ p submersion. This map is linear, so it is enough to show that it is surjective. The relations defining B0 as a subspace of all 2-tensors translate (via Rb(A B),C D 〈 ∧ ∧ 〉 = R(A,B)C,D ) to 〈 〉 Rb(A B) Rb(B A), ∧ = − ∧ Rb(A B),C D Rb(C D), A B , 〈 ∧ ∧ 〉 = 〈 ∧ ∧ 〉 2.4. CONSTRUCTION OF NGI,T AND NGI,` 53

Rb(A B),C D Rb(A C),D B Rb(A D),B C 0. 〈 ∧ ∧ 〉 + 〈 ∧ ∧ 〉 + 〈 ∧ ∧ 〉 = We are interested in the ways in which these relations impose relations among the ci,j ; in fact our goal is to show that no such relations exist. The first relation V2 follows immediately from Rb being a tensor on TM, and so gives no restriction on the ci,j . The second tells us that the matrix representation of Rb is symmetric, but since we required that i j, this puts no restriction on the c . The third ≤ i,j condition comes from the Bianchi identity, i.e. the definition of B0 as the kernel 0 of the map b . The components ci,j all refer to entries corresponding to basis bivectors X X for k n. Explicitly k ∧ n <

ci,j Rb(X j Xn), Xi Xn . = 〈 ∧ ∧ 〉 For such vectors

Rb(X j Xn), Xi Xn Rb(X j Xi ), Xn Xn Rb(X j Xn), Xn Xi 0, 〈 ∧ ∧ 〉 + 〈 ∧ ∧ 〉 + 〈 ∧ ∧ 〉 = so the third condition gives no relations between the ci,j . This means that given any choice of components c (with 1 i j n) there is an element of B0 hav- i,j ≤ ≤ < ing these components in its lower right block. In other words, the linear map

c {g } {[X ] } B0 { S } is surjective and hence a submersion, so the full map c is also | p × p × p × ∇ p a submersion. 0,t 1 n(n 1)/2 As noted above, NG c− ({0}). Since c is a submersion with image R − , = this means that NG0,t is a submanifold of TLV B0 B1 of codimension n(n 1)/2, ⊕ ⊕ − and hence also a submanifold of L PTM B0 B1 of codimension n(n 1)/2. ⊕ ⊕ ⊕ − Lemma 2.4.6. The set NG1,t is a submanifold of L PTM B0 B1 of codimension ⊕ ⊕ ⊕ n(n 1)/2. − Proof. The proof of this lemma is similar to that of Lemma 2.4.5. Choose the family of frames V g,[X ] in the same way. With this choice, S B1 may be ∇Xn ∈ expressed as a block matrix, the same way R B0 way expressed as a block ma- ∈ trix in the proof of Lemma 2.4.5. Just as in the preceding proof, define the func- tions ci,j to be the components of the lower right-hand block. We need the map 0 1 n(n 1)/2 c (c1,1,c1,2,...,cn 1,n 1): TLV B B R − to be a submersion. The = − − ⊕ ⊕ → same argument as in the proof of Lemma 2.4.5 shows that the first Bianchi identity does not impose any relations between the ci j . Similarly, the second Bianchi iden- tity only imposes relations between Sabcde for different e, never relations between components with the same e. In particular, since the ci j are defined using Sabcdn, no relations between them are imposed by the second Bianchi identity. Hence c is a submersion, and we conclude that NG1,t has codimension n(n 1)/2. − Proposition 2.4.7. NG0,t NG1,t is a submanifold of L PTM B0 B1 of codi- ∩ ⊕ ⊕ ⊕ mension n(n 1). − 54 CHAPTER 2. THE GENERIC CONDITION

Proof. Since NG0,t is of the form E B1 for some fiber bundle E L PTM, and 0 ⊕ 0 → ⊕ NG1,t is of the form E B0 for some fiber bundle E L PTM, their intersec- 1 ⊕ 1 → ⊕ tion NGt NG0,t NG1,t is the product bundle E E , which is a submanifold of = ∩ 0 ⊕ 1 L PTM B0 B1 of codimension codimNG0,t codimNG1,t n(n 1). ⊕ ⊕ ⊕ + = − To compute the codimensions of NG0,` and NG1,` we proceed similarly. How- ever, we may obtain a sharper bounds on the codimensions by making use of the fact that these spaces are defined in terms of lightlike vectors. Let LLV denote the bundle LLV {(g,[X ]) L PTM [X ] is lightlike for g}. = ∈ ⊕ | Then LLV has codimension 1 in L PTM. ⊕ Lemma 2.4.8. The set NG0,` is a submanifold of LLV B0 B1 of codimension ⊕ ⊕ (n 1)(n 2)/2. − − Proof. The proof is analogous to that of Lemma 2.4.5, with the exception that the g,[X ] frame V is chosen so that X2,..., Xn 1 are all orthogonal to Xn, but X1 need − not be. This means that when Rb is written in block form à ! AB BC where lower right-hand block C is a (n 2) (n 2) matrix, the equation describing − × − NG0,` is precisely that C is zero. As in the proof of Lemma 2.4.5, define functions ci,j by extracting the (i, j) component of C. This time, since C is smaller, we get (n 1)(n 2)/2 different functions ci,j (after considering the symmetry of Rb). As − − before, these are independent in the sense that c (c1,1,c1,2,...,cn 1,n 1) is a sub- = − − mersion. Hence NG0,` has codimension (n 1)(n 2)/2 in LLV B0 B1. − − ⊕ ⊕ Lemma 2.4.9. The set NG1,` is a submanifold of LLV B0 B1 of codimension ⊕ ⊕ (n 1)(n 2)/2. − − Proof. To compute the codimension of NG1,`, do the same thing as for NG1,t , but in the bundle LLV B0 B1 and with a block C of size (n 2) (n 2). This yields ⊕ ⊕ − × − a codimension of (n 1)(n 2)/2. − − Proposition 2.4.10. NG0,` NG1,` is a submanifold of L PTM B0 B1 of codi- ∩ ⊕ ⊕ ⊕ mension (n 1)(n 2) 1. − − + Proof. Since NG0,` and NG1,` impose restrictions in different components of the product LLV B0 B1, their intersection NG` NG0,` NG1,` is a submanifold of ⊕ ⊕ = ∩ codimension codimNG0,` codimNG1,` (n 1)(n 2) in LLV B0 B1. Since + = − − ⊕ ⊕ LLV B0 B1 has codimension 1 in L PTM B0 B1, we conclude that NG` ⊕ ⊕ ⊕ ⊕ ⊕ has codimension (n 1)(n 2) 1 in L PTM B0 B1. − − − ⊕ ⊕ ⊕ 2.5. THE MAP α 55

2.5 The map α

The purpose of this section is to define a bundle map

α: PTM J 3L PTM L B0 B1 ⊕ → ⊕ ⊕ ⊕

and show that it is a submersion. Recall that the jet bundle J 3L has canonical projections π : J 3L J i L for i 0,1,2. These are described in Section A.5. i → = Definition 2.5.1. Let α be the bundle map

(id ,π ,R0,R1) PTM J 3L PTM 0 PTM L B0 B1. ⊕ −−−−−−−−−−−→ ⊕ ⊕ ⊕

Here J 3L denotes the bundle of 3-jets of metrics on M, and PTM denotes the pro-

jectivized tangent bundle of M. The map idPTM is the identity on PTM. The map π : J 3L L is the projection onto the 0-jet part of J 3L. (Note that we use the 0 → canonical isomorphism between L and J 0L to view π as a map J 3L L.) The 0 → map R0 : J 3L B0 computes the Riemann curvature tensor of a metric. More → specifically, if g J 3 L then R0(g) is the curvature tensor at p of any metric corre- ∈ p sponding to the equivalence class g. Similarly, the map R1 : J 3L B1 computes → the first covariant derivative R1(g) R of the curvature tensor R of g. The 3-jet = ∇ of a metric is precisely the information needed to compute these tensors.

0 1 We wish to show that α : (idPTM ,π0,R ,R ) is a submersion. Since the map 0 1 = 0 1 3 0 1 (idP ,π ,R ,R ) is simply the product of the map (π ,R ,R ): J L L B B TM 0 0 → ⊕ ⊕ with the identity idPTM : PTM PTM, and the identity is a submersion, it is suf- 0 1 → ficient to show that (π0,R ,R ) is a submersion. 0 1 To show that (π0,R ,R ) is a submersion, we will use the following lemma.

π π0 Lemma 2.5.2. Let B, F and F 0 be smooth manifolds. Let E B and E 0 B be −→ −→ smooth fiber bundles with fibers F and F 0. Let ϕ: E E 0 be a smooth bundle map. → Suppose that for each p B the fiberwise restriction ϕ: E E 0 is a submersion. ∈ p → p Then ϕ is itself a submersion.

Proof. Choose some point e E, and let b π(e). Choose an open neighborhood ∈ = U B of b such that E and E 0 both have trivializations over U. Use this trivial- ⊆ ization to express e (b, f ). For each c U, the bundle map ϕ restricts to some = ∈ map ϕ :{c} F {c} F 0, i.e. to a map ϕ : F F 0. By hypothesis, this map c × → × c → is a submersion, so there are coordinates on a neighborhood V F of f and its c ⊆ image ϕ (V ) F 0 such that ϕ is the projection onto the first k coordinates (where c c ⊆ k dimF dimF 0). We may choose these coordinate systems to depend smoothly = − 56 CHAPTER 2. THE GENERIC CONDITION

T on c, and (possibly after shrinking U) such that V c U Vc . In coordinates in = ∈ 6= ; the neighborhood U V of e, we then have × ϕ(b, f ) (b,ψ(f )) = where ψ denotes projection onto the first k coordinates. Hence we can locally express ϕ as a product of the identity and a projection, both of which are submer- sions, so ϕ is a submersion.

Proposition 2.5.3. The map (π ,R0,R1): J 3L L B0 B1 is a submersion. 0 → ⊕ ⊕ Proof. Note that J 3L and L B0 B1 are bundles over M, so the lemma tells us ⊕ ⊕ that (π ,R0,R1) is a submersion if its restriction to the fiber over any point p M 0 ∈ is a submersion. We then need to show that

(π ,R0,R1): J 3 L L B0 B1 0 p → p ⊕ p ⊕ p is a submersion. Note now that both J 3 L and L B0 B1 are bundles over L . p p ⊕ p ⊕ p p This means that the lemma tells us that it is sufficient to show, for each h L , ∈ p that the restriction to the fiber over h is a submersion. Thus we have reduced the problem to showing that

(R0,R1):(J 3 L) B0 B1 p h → p ⊕ p is a submersion. Choose some arbitrary point [g] (J 3 L) . We will show that (R0,R1) is a sub- ∈ p h mersion at [g]. In fact, we will show that the restriction of (R0,R1) to the set n o J : [k] (J 3 L) : π (k) π (g) is a submersion. The reason for this restriction = ∈ p h 1 = 1 is to make the 1-jets of the elements of the domain agree, so that we may choose normal coordinates: Choose coordinates centered p which are normal with re- spect to π1(g). In these coordinates, an element [g] of J is characterized by the 0 tensors gab,cd and gab,cde . Similarly, an element of B is a tensor Rabcd and an 1 element of B is a tensor Rabcde . (We use implicitly the isomorphisms mentioned in Section 2.3 to view B0 and B1 as spaces of 4- and 5-tensors, respectively. Since we have restricted our attention to the set J, these isomorphisms are canonical.) In the center of normal coordinates it holds that the Riemann tensor of a met- ric g is 1 R Γ Γ ¡g g g g ¢ abcd = abd,c − abc,d = 2 ad,bc + bc,ad − ac,bd − bd,ac and that its first derivative is 1 R R Γ Γ ¡g g g g ¢. abcd;e = abcd,e = abd,ce − abc,de = 2 ad,bce + bc,ade − ac,bde − bd,ace 2.6. TRANSVERSALITY 57

These are then the coordinate expressions of R0 and R1 in normal coordinates at p. Note in particular that they are linear. This means that R0 and R1 agree with 3 their differential maps at [g]. Note that the tangent space of (Jp L)h is the product 4 5 4 TQ TQ where TQ is the vector space of 4-tensors with the symmetries of second × 5 derivatives of a metric, and TQ is the vector space of 4-tensors with the symmetries of third derivatives of a metric. Explicitly, the tangent maps of R0 and R1 are then

0 1 R (Qabcd ) (Qadbc Qbcad Qacbd Qbdac ), ∗ = 2 + − −

0 1 ¡ ¢ R (Qabcd,e ) Qadbc,e Qbcad,e Qacbd,e Qbdac,e . ∗ = 2 + − − It is now sufficient to show that (R0,R1) is a surjective linear map. The two maps ∗ ∗ are surjective (in fact isomorphisms) since the following maps are their inverses [27]. 1 g (R R ) ab,cd = 3 acbd + adbc 1 X gab,cde Raib j,k = 6 (i jk) The sum ranges over all permutations i jk of cde. Note that these two maps are independent of each other: The map à ! 1 1 X (Rαβγδ,Rabcde ) (Rαγβδ Rαδβγ), Raib j,k 7→ 3 + 6 (i jk)

is an inverse of (R0,R1) and we have shown that it is a submersion. ∗ ∗ As argued earlier, this implies the following corollary.

Corollary 2.5.4. The bundle map

0 1 3 0 1 (idP ,π ,R ,R ): PTM J L PTM L B B . TM 0 ⊕ → ⊕ ⊕ ⊕ is a submersion.

2.6 Transversality

An application of Proposition B.3.3 yields the following result.

Proposition 2.6.1. Let M be a manifold without boundary. Let W J 3L PT M be ⊆ ⊕ a submanifold of codimension q. Then the set of metrics g Γ∞(L) such that the 3 3 ∈ image of the map PTM J L PT M defined by Vx (j gx ,Vx ) is transverse to W → 4 ⊕ 7→ is residual in the Whitney C topology on the space Γ∞(L) of all metrics. 58 CHAPTER 2. THE GENERIC CONDITION

Proof. This is simply an application of Proposition B.3.3 with r 3 and K PTM. = =

Proposition 2.6.2. Let (m,g) be a Lorentzian manifold with curvature tensor R. Let γ be a geodesic with initial velocity X : γ˙(0). Let ρ(g): PTM J 3L PT M denote = → ⊕ the map defined by ρ(g)(V ) (j 3g ,V ). If either x = x x 1 t • X is timelike and ρ(g)([X ]) α− (NG ), or ∉ 1 ` • X is lightlike and ρ(g)([X ]) α− (NG ) ∉ then for every δ 0 there is a point t ( δ,δ) such that γ˙(t) is a generic vector. > ∈ − 1 ` Proof. Consider the case when X is a lightlike vector. That ρ(g)([X ]) α− (NG ) ∉ means that either α(ρ(g)([X ])) NG0,` or α(ρ(g)([X ])) NG1,`. In the first case, ∉ ∉ we know by definition of α and NG0,` that there are vectors A and B orthogonal to X such that Rb(A X ),B X 0 〈 ∧ ∧ 〉 6= where R is the curvature tensor corresponding to g. By Proposition 2.2 of [1], this means that X is a generic vector, and so we may choose t 0. = The remaining case is when α(ρ(g)([X ])) NG0,` but α(ρ(g)([X ])) NG1,`. By ∈ ∉ definition of NG1,` we then know that there are vectors E and F orthogonal to X such that

X R(E X ),F X 0. 〈∇ ∧ ∧ 〉 6= By the same argument as is used in [1], this is equivalent to that there are vectors A, B, C and D such that

0 X e X f X ( R) X AaB bC c Dd . 6= [a ∇X b]e f [c d] Extend these four vectors by parallel transport to vector fields along γ in a neigh- borhood of γ(0). Then A B C D 0 so the above is equivalent to ∇X = ∇X = ∇X = ∇X = that 0 X (k) 6= where k is the function along γ defined by

k X e X f X R X AaB bC c Dd . = [a b]e f [c d] By definition, it holds that k(t) 0 only if γ˙(t) is generic. Since we have now shown 6= that dk (0) X (k) 0 and we assumed that k(0) 0, there must be some point t in dt = 6= = any neighborhood of 0 such that k(t) 0. Hence γ˙(t) is generic for arbitrarily small 6= t. The proof in the case when X is timelike is completely analogous. This com- pletes the proof. 2.6. TRANSVERSALITY 59

We are now finally in a position to prove our main theorem.

Theorem 2.6.3. Let M be a smooth manifold of dimension n 4, possibly with ≥ boundary. Let L denote the bundle of Lorentzian metrics on M and endow the set r Γ∞(L) of smooth sections with the Whitney C topology for any 4 r . Then ≤ ≤ ∞ there is a residual (and hence dense) set G Γ∞(L) such that if g G and γ is a e f ⊆ ∈ geodesic then the points where γ˙ γ˙ γ˙[aRb]e f [c γ˙d] is nonzero along γ form a dense set. In particular,

• the lightlike generic condition holds for each metric g G, ∈ • the timelike generic condition holds for each metric g G, ∈ • the spacelike generic condition holds for each metric g G. ∈ Proof. We note that all of the constructions for the timelike case work equally well for the spacelike case, so we prove the result only for timelike and lightlike geodesics. 0 1 Consider the following diagram, where α (idP ,π ,R ,R ) is the map con- = TM 0 sidered in Section 2.5.

α J 3L PTM L PTM B0 B1 ⊕ ⊕ ⊕ ⊕

1 α α− (NG) NG

Here NG denotes the union NGt NG`, so by the results of Section 2.4 we ∪ know that NG is a union of submanifolds of codimensions (at least) n(n 1) and − (n 1)(n 2) 1. Since α is a submersion, as was shown in the previous sec- − − + 1 tion, and inverse submersions preserve codimension α− (NG) is a union of two submanifolds of codimension at least n(n 1) and (n 1)(n 2) 1. Applying − − − + Proposition 2.6.1 to each of these submanifolds in turn, we conclude that there is a residual set U ` of metrics g for which the map PTM J 3L PTM defined 3 1 ` → t ⊕ by Vx (j gx ,Vx ) is transverse to α− (NG ), and a residual set U of metrics such 7→ 1 t that the map is transverse to α− (NG ). The intersection of two residual sets is again residual, so U U ` U t is residual. It now remains to see what it means that = ∩ a metric belongs to U. Choose some g U. The image of ρ(g) (V (j 3g ,V )) has dimension ∈ = x 7→ x x 2n 1, so in particular − 2n 1 n(n 1) codim(NGt ) when n 4. − < − ≤ ≥ 60 CHAPTER 2. THE GENERIC CONDITION

1 t Transversality then means that the image of ρ(g) is disjoint from α− (NG ). Simi- larly, 2n 1 codim(NG`) 2n 1 (n 1)(n 2) 1 0 if n 4, − − ≤ − − − − − ≤ ≥ so im(ρ(g)) can intersect NG` only in submanifolds of dimension 0. In particular, along every geodesic γ there is a dense set of points t where 1 γ˙(t) α− (NG). By Proposition 2.6.2, this means that there is also a dense set of ∉ points where γ˙ is generic. In particular the generic condition holds. BackgroundA

A.1 Preliminaries

A.1.1 Manifolds and fiber bundles

Convention A.1.1. If nothing to the contrary is explicitly stated, all manifolds will be smooth, finite-dimensional, second-countable and paracompact without bound- ary. We will also work with manifolds with boundary, and in these instances this will be stated explicitly. We will sometimes write ”manifold without boundary” for emphasis.

Convention A.1.2. We will use the word submanifold to mean the same thing as embedded submanifold. Similarly, ”hypersurface” will mean ”embedded hyper- surface”.

Convention A.1.3. The word closed when used about subsets, hypersurfaces or submanifolds will mean closed in the topological sense, not in the sense of com- pact without boundary.

Definition A.1.4. A subset of a smooth (n 1)-manifold is a topological hypersur- + face if it is locally Euclidean of dimension n.

Remark A.1.5. Note that all subsets of a smooth manifold are Hausdorff, para- compact and second-countable, so these properties hold for all topological hy- persurfaces.

Convention A.1.6. By fiber bundle, we will mean a smooth fiber bundle, i.e. a fiber bundle in which the base space, fiber and total space are smooth manifolds, and all relevant maps are smooth. π π Convention A.1.7. When E 1 M and E 2 M are fiber bundles over a common 1 −→ 2 −→ base space we will denote the fiber bundle product of E and E by E E . In 1 2 1 ⊕ 2 other words, E E {(e ,e ) E E π (e ) π (e )} 1 ⊕ 2 = 1 2 ∈ 1 × 2 | 1 1 = 2 2 as sets. 62 APPENDIX A. BACKGROUND

A.1.2 Cobordisms

Definition A.1.8. Let S and S0 be smooth compact manifolds of dimension n. A cobordism between S and S0 is a compact (n 1)-dimensional manifold-with- + boundary M, the boundary of which is the disjoint union S S0. If there is a cobor- t dism between S and S0, we say that they are cobordant.

The following theorem (see [24, Corollary 4.11]) characterizes the cobordism relation in terms of Stiefel-Whitney numbers.

Theorem A.1.9. Two compact manifolds without boundary are cobordant if and only if their Stiefel-Whitney numbers agree.

The cobordism relation is particularly simple in dimension 3. (See [24, p. 203].)

Theorem A.1.10. Any two compact three-dimensional manifolds without bound- ary are cobordant.

A.2 Lorentzian geometry

For a more detailed discussion of the definitions in this section, see [28] and [7].

Definition A.2.1. A spacetime is a time-oriented Lorentzian manifold of dimen- sion at least 2.

Definition A.2.2. A C 1 submanifold N of a spacetime (M,g) is spacelike if the re- striction of g to TN is positive definite.

Definition A.2.3. A C 1 hypersurface H in a spacetime (M,g) is a null hypersurface if the restriction of g to T H is degenerate for all p H. p ∈ Definition A.2.4. Let (M,g) be a spacetime. A vector X TM is ∈ • timelike if g(X , X ) 0, < • lightlike or null if g(X , X ) 0, = • causal if g(X , X ) 0, ≤ • spacelike if g(X , X ) 0. > Definition A.2.5. A curve γ in a spacetime is timelike if for all t in the domain of γ the vector γ˙(t) is timelike. A curve is causal if each vector γ˙(t) is causal. A.2. LORENTZIAN GEOMETRY 63

Definition A.2.6. Let M be a spacetime, and let γ:[a,b) M be a future-directed → causal curve. A point p is a future endpoint of γ if p lim γ(t). = t b → Definition A.2.7. A future-directed causal curve γ:[a,b) M is future extendible → ¯ if there is a future-directed causal curve γ0 :[a,c) M with c b and γ0¯ γ. → > [a,b) = Definition A.2.8. Let M be a spacetime with or without boundary. For any point p M, we define the chronological future of p to be the set ∈ I +(p) {q M : there is a (non-constant) future-directed = ∈ timelike curve from p to q}. S The chronological future of a set A M is defined to be I +(A) p A I +(A). Anal- ⊆ = ∈ ogously, the chronological past of a point p M is the set ∈ I −(p) {q M : there is a (non-constant) future-directed = ∈ timelike curve from q to p}, S and the chronological past of a set A is I −(A) p A I −(p). We define = ∈ I(A) I +(A) I −(A). = ∪ We define the causal future and causal past of a point p to be

J +(p) {q M there is a future-directed causal curve from p to q} = ∈ | and

J −(p) {q M there is a future-directed causal curve from q to p}, = ∈ | respectively.

We also define pasts and futures relative to a subset U M. The set I +(p,U) ⊆ is the set of points which can be reached by a timelike curve which is completely contained in U. We define I −(p,U), I ±(A,U), J ±(p,U) and J ±(A,U) similarly.

Note that I +(A) and I −(A) are open sets for each subset A of the spacetime. (See [7, Proposition 2.4.12].)

Definition A.2.9. Let M be a spacetime without boundary. A set A is achronal if no two points of A can be connected by a timelike curve. It is acausal if no two points can be connected by a causal curve.

Definition A.2.10. Let A be an achronal set. Then edge(A) denotes the set of points p A such that every neighborhood U of p contains a timelike curve from ∈ I −(p,U) to I +(p,U) which does not intersect A.

Definition A.2.11. Let M be a spacetime and let A M be an arbitrary subset. A ⊆ future-directed causal curve γ: I M is totally past imprisoned in A if there is → some t I such that γ(s) A for all s I with s t. ∈ ∈ ∈ ≤ 64 APPENDIX A. BACKGROUND

A.2.1 Cauchy developments and Cauchy horizons

Definition A.2.12. Let A be an arbitrary subset of a spacetime M. The future Cauchy development of A is the set

D+(A) {p M every past inextendible causal curve = ∈ | through p intersects A}.

Informally, D+(A) consists of the points about which one has complete knowledge given initial data on A. The past Cauchy development D−(A) may be defined anal- ogously, but we will not have use for it.

Note that D+(A) is defined in terms of causal curves. Using an analogous def- inition with timelike curves yields the topological closure of D+(A):

Proposition A.2.13. Let M be a spacetime without boundary, and let A be an arbi- trary subset of M. Then

D (A) {p M every past inextendible timelike curve through p intersects A}. + = ∈ | Proof. See [21, Lemma 8.3.8].

The ”future boundary” of D+(A) will be of particular interest:

Definition A.2.14. Let A be an arbitrary subset of a spacetime M. The future Cauchy horizon of A is the set

H +(S) D (A)\ I −(D+(A)). = + Definition A.2.15. A topological hypersurface S in a spacetime M is a Cauchy sur-

face if M D−(S) D+(S). = ∪

A.2.2 Causality conditions

Definition A.2.16. A spacetime is chronological (or ”satisfies the chronology con- dition”) if it contains no closed timelike curves.

Definition A.2.17. A spacetime is causal (or ”satisfies the causality condition”) if there are no closed causal curves.

Definition A.2.18. A spacetime is strongly causal (or ”satisfies the strong causality condition”) if each point has a neighborhood which intersects no timelike curve more than once.

Definition A.2.19. A spacetime is stably causal if it admits a global time function. A.3. ENERGY CONDITIONS 65

Definition A.2.20. A spacetime is globally hyperbolic if it contains a Cauchy sur- face.

Proposition A.2.21. Let S be an achronal hypersurface in a spacetime M. Then the interior of D (S) D (S) is globally hyperbolic (provided it is nonempty). + ∪ − Proof. See [7, Proposition 2.9.9].

Proposition A.2.22. Let (M,g) be a spacetime. Then

M is globally hyperbolic

⇓ M is stably causal

⇓ M is strongly causal

⇓ M is causal

⇓ M is chronological.

A.3 Energy conditions

Throughout this section (M,g) will denote a spacetime. R will denote its curvature tensor, Ric its Ricci curvature, S its scalar curvature, and T the stress-energy tensor. We define T through Einstein’s equation using a cosmological constant Λ.

1 8πT Ric Sg Λg. = +2 +

We are mainly concerned with lightlike vectors, so the cosmological constant plays no role.

Definition A.3.1. We define the following energy conditions. Name of energy condition Definition Lightlike convergence condition Ric(v,v) 0 for all lightlike vectors v ≥ Null energy condition T (v,v) 0 for all lightlike vectors v ≥ Weak energy condition T (v,v) 0 for all causal vectors v ≥ Strict lightlike convergence condition Ric(v,v) 0 for all lightlike vectors v > Strict null energy condition T (v,v) 0 for all lightlike vectors v > Strict weak energy condition T (v,v) 0 for all causal vectors v > 66 APPENDIX A. BACKGROUND

Remark A.3.2. Note that the lightlike convergence condition is equivalent to the null energy condition. Both of them are implied by the weak energy condition.

Definition A.3.3. Let M be a spacetime. We say that a vector X TM is generic if e f ∈ the tensor X X X[aRb]e f [c Xd] is not equal to zero.

Definition A.3.4. Let M be a spacetime with curvature tensor R. We say that M satisfies the timelike/lightlike/spacelike generic condition if it holds that each in- extendible timelike/lightlike/spacelike geodesic γ has some point at which γ˙(t) is generic.

Definition A.3.5. We say that a spacetime satisfies the generic condition if it satis- fies the timelike and lightlike generic conditions.

Remark A.3.6. The generic condition is also known as the ”genericity condition” (see [21, p. 388]) and the ”generality condition” (see [23, p. 30]). The term ”generic condition” is used in [19, p. 101] and [1].

A.4 Geometric measure theory

References for this section are [13] and [25].

A.4.1 Regularity of functions

Definition A.4.1. A function f : Rn Rm is C 1,1 if it is C 1 and its differential d f is → Lipschitz. A submanifold of a smooth manifold is C 1,1 if it is locally the graph of a C 1,1 function in coordinates.

Definition A.4.2. A function f : Rn R is semi-convex if it is the sum of a convex → function and a C 2 function. A submanifold of a smooth manifold is semi-convex if it is locally the graph of a semi-convex function in coordinates.

A.4.2 Measure zero

Let Σ be some smooth manifold of dimension m, and let M be a smooth mani- fold of dimension at least m. Let ψ: Σ M be a topological embedding. We will → consider two notions of ”measure zero”:

• Since Σ is a smooth manifold (and in particular second countable), any two Riemannian metrics on Σ give rise to the same family of sets having zero measure in the associated volume measure on Σ. A.4. GEOMETRIC MEASURE THEORY 67

• Let σ be an arbitrary Riemannian metric on M. This metric induces a dis- tance function, which in turn induces Hausdorff measures of any dimen- sion. Let hm denote the m-dimensional Hausdorff measure induced by σ. Then we say that A ψ(Σ) has measure zero if hm(A) 0. ⊆ = These two notions are connected in the following sense.

Proposition A.4.3. Let Σ be some smooth manifold of dimension m, and let M be a smooth manifold of dimension n with n m. Let ψ: Σ M be a topological ≥ → embedding. Suppose that ψ is locally Lipschitz. Then hm(ψ(A)) 0 if A has measure = zero viewed as a subset of Σ.

Proof. After representing ψ is coordinates by ψ: U V with open sets U Σ and → ⊆ V M identified with subsets of Rm and Rn it holds that ⊆ hm(ψ(A U)) Lµ(A U) ∩ ≤ ∩ where L is the Lipschitz constant of ψ over U, and µ denotes the m-dimensional Hausdorff measure on U. (This can be proved by bounding the volume change of images of unit balls using the Lipschitz constant, or it can be seen as a special case of the much more powerful [13, Theorem 2.10.25].) Since U is a subset of Rm, this Hausdorff measure agrees up to pointwise scaling by a smooth function with the Lebesgue measure in coordinates. In particular, if A U has zero measure in Σ, ∩ then hm(ψ(A U)) 0. By second countability, countably many charts suffice to ∩ = cover ψ(Σ), and so hm(ψ(A U)) 0 if A has measure zero. ∩ = In general, we will mostly be interested in the notions of ”measure zero” and ”finite measure”, so it will not matter precisely which Riemannian metric is used to induce a measure.

Proposition A.4.4. Let M be a smooth manifold of dimension at least n and let N be a Lipschitz submanifold of M with dimension n. Let K be a compact subset of N. Let σ be a Riemannian metric on M, and let hn be the corresponding n-dimensional Hausdorff measure. Then all measurable subsets of K have finite hn-measure.

Proof. It is sufficient to show that K has finite measure, since subsets of K have S smaller measure than K . Take a finite subcover of the cover K p H Bσ(p,1). ⊆ ∈ Each H B (p,1) has finite hn-measure since H is a Lipschitz hypersurface. Hence ∩ σ we can conclude that K has finite measure.

A.4.3 Density functions

A reference for density functions is [25, Chapter 2]. We will use the same idea, but with somewhat different notation. 68 APPENDIX A. BACKGROUND

Definition A.4.5. Let Ln denote Lebesgue measure on Rn. For each measurable subset A Rn define the density function of A (with respect to Ln) to be the func- ⊆ tion Θ(A, ): A [0,1], · → Ln(A B n(q,r )) Θ(A,q) lim ∩ . = r 0 Ln(B n(q,r ) → Here B n(q,r ) denotes the ball of radius r centered at q.

Definition A.4.6. Let A be a measurable subset of Rn. We will call the set

A˜ {a A Θ(A,a) 1} = ∈ | = the full-density subset of A.

Proposition A.4.7. Let A˜ be the full-density subset of some set A Rn. Then A˜ has ⊆ full Lebesgue measure in A.

Proof. By [25, Corollary 2.9] the density function Θ(A, ) is equal to the character- · istic function of A almost everywhere, yielding the conclusion.

We now generalize the notion of full-density subsets to hypersurfaces in Rie- mannian manifolds.

Lemma A.4.8. Let (M,σ) be a of dimension n 1 and let + H be a Lipschitz hypersurface. Let U be an open subset of H and let ϕ: U Rn → and ψ: U Rn be charts. Let A be a subset of U and let A˜ and A˜ denote the → ϕ ψ full-density subsets of ϕ(A) and ψ(A), respectively. Then

1 A˜ ψ(ϕ− (A˜ )). ψ = ϕ 1 Proof. Abbreviate ψ ϕ− by f . Since f is bi-Lipschitz, it holds for any measurable ◦ subset X of imϕ that

1 n n n n n L (X ) L (f (x)) L f L (X ) L 1 ≤ ≤ f −

1 where L f −1 and L f denote the Lipschitz constants of f − and f . In particular, n n n L n n L (B (q,r )\ ϕ(A)) f 1 L (f (B (q,r )) \ ψ(A)) − . Ln n n Ln n (B (q,r )) ≤ L f (f (B (q,r ))) By letting R(r ) be a positive real number such that f (B n(q,r )) B n(f (q),R(r )) and ⊆ ρ(r ) a positive real number such that f (B n(q,r )) B n(f (q),ρ(r )) we see that ⊇ n n n L n n L (B (q,r )\ ϕ(A)) f 1 L (B (f (q),R(r )) \ ψ(A)) − . Ln n n Ln n (B (q,r )) ≤ L f (B (f (q),ρ(r ))) A.4. GEOMETRIC MEASURE THEORY 69

1 Since f and f − are Lipschitz, we may choose R and ρ to be bounded from above and below by linear functions of positive derivative, so there are positive constants

D and D0 such that

n n n n n n DL (B (f (q),ρ(r ))) L (B (f (q),R(r ))) D0L (B (f (q),ρ(r ))). ≤ ≤ This together with the previous inequality means that there is a positive real num- ber C independent of r such that

Ln(B n(q,r )\ ϕ(A)) Ln(B n(f (q),R(r )) \ ψ(A)) C . Ln(B n(q,r )) ≤ Ln(B n(f (q),R(r )))

In particular, if Θ(ψ(A), f (q)) 1 so that = Ln(B n(f (q),R(r )) \ ψ(A)) lim 0 r 0 Ln(B n(f (q),R(r ))) = → then Ln(B n(q,r )\ ϕ(A)) lim 0 r 0 Ln(B n(q,r )) = → so that Θ(φ(A),q) 1. Hence = Θ(ψ(A), f (q)) 1 Θ(φ(A),q) 1. = =⇒ = Repeating this argument for the inverse of f we see that

Θ(φ(A),q) 1 Θ(ψ(A), f (q)) 1. = =⇒ = This proves that A˜ f (A˜ ), completing the proof. ψ = ϕ Definition A.4.9. Consider a Riemannian manifold (M,σ) of dimension n 1 and + let H be a Lipschitz hypersurface. In light of the previous proposition, we may define the full-density subset of a set A H to be the set A˜ such that if ϕ: U Rn ⊆ → is a chart on H then ϕ(A˜ U) is the full-density subset of ϕ(A U). ∩ ∩ Definition A.4.10. If q belongs to the full-density subset of a set A, we say that q is a full-density point of A.

Proposition A.4.11. Let (M,σ) be a Riemannian manifold of dimension n 1 and + let H be a Lipschitz hypersurface. Let hn be the n-dimensional Hausdorff measure constructed from σ. Let A H be any subset and let A˜ be its full-density subset. ⊆ Then hn(A \ A˜) 0. = Proof. Since H is second-countable, it is sufficient to prove this locally. This can be done by the use of charts and Proposition A.4.7. 70 APPENDIX A. BACKGROUND

Proposition A.4.12. Let Ω be an open subset of Rn and let f ,g : Ω R be Lipschitz → functions. Let A Ω be a measurable subset of Ω and suppose that f and g agree on ⊂ A. Let q be a full-density point of A and suppose that f and g are both differentiable at q. Then d f (q) dg(q). = If, moreover, q is a point where f and g have second order expansions of the form 1 f (q ξ) f (q) d f (q)(ξ) D2 f (q)(ξ,ξ), + = + + 2 1 g(q ξ) g(q) dg(q)(ξ) D2g(q)(ξ,ξ), + = + + 2 then D2 f (q) D2g(q). = Proof. Let h f g and note that h is differentiable at q and zero on A. Suppose = − that dh(q)(V ) 0 for some vector V at q. By continuity dh(q)(W ) for all W in 6= some open neighborhood U of V in T Rn. Then for all sufficiently small ² 0 q > and all W U it holds that h(q ²U) 0. This means that h is nonzero on some ∈ + 6= small open cone in the direction of V , which in turn means that Θ(A,q) cannot be equal to 1, contradicting the fact that q is a full-density point of A. This shows that dh(q) 0, proving the first part of the proposition. = Suppose now that q is a point where f and g have second order expansions in the sense that 1 f (q ξ) f (q) d f (q)(ξ) D2 f (q)(ξ,ξ), + = + + 2 1 g(q ξ) g(q) dg(q)(ξ) D2g(q)(ξ,ξ). + = + + 2 Then, since f (q) g(q) and d f (q) dg(q), = = 1 f (q ξ) g(q ξ) ¡D2 f (q) D2g(q)¢(ξ,ξ). + − + = 2 − If ¡D2 f (q) D2g(q)¢(ξ,ξ) 0 for every vector ξ then D2 f (q) D2g(q) 0 and we − = − = are done (since D2 f (q) D2g(q) is symmetric), so suppose that there is some ξ − with ¡D2 f (q) D2g(q)¢(ξ,ξ) 0. By continuity, this then holds for all ν in a neigh- − 6= borhood of ξ, and this means that f g is nonzero on a small open cone from − q. This means that Θ(A,q) cannot be equal to 1 contradicting the fact that q is a full-density point of A, showing that indeed D2 f (q) D2g(q). =

A.5 Jet bundles

In Chapter 2 we need the concept of ”jets”. In this section, we give the necessary definitions. For the most part, we use the notation of [31]. For jets of arbitrary maps rather than jets of sections of a smooth fiber bundle, we follow [17, §2 Chap- ter II]. The idea of jets is to capture the concept of an ”abstract Taylor polynomial” A.5. JET BUNDLES 71

of a function between manifolds, by considering equivalence classes of functions which have the same Taylor polynomials in coordinates.

A.5.1 Jets of maps between manifolds

Definition A.5.1. Let X and Y be smooth manifolds, and choose a nonnegative integer k. Let P X P(X ) denote the set of those pairs (x,U) such that x U. k ⊆ × ∈ Let J (X ,Y ) be the set of equivalence classes on the set P C ∞(X ,Y ) under the × relation defined by (x ,U , f ) (x ,U , f ) if 1 1 1 ∼ 2 2 2 • x x , and 1 = 2 ¯ ¯ • f1¯ f2¯ , and U1 = U2 • after choosing coordinates around x x and f (x ) f (x ), the partial 1 = 2 1 1 = 2 2 derivatives of f and f at x x agree up to order k. 1 2 1 = 2 This is independent of the coordinates chosen. The set J k (X ,Y ) is called the set of k-jets of maps from X to Y .

Locally, J k (X ,Y ) looks like a product U V Rm (where U X , V Y ), and × × ⊆ ⊆ this local structure can be patched together to make J k (X ,Y ) into a smooth fiber bundle over X Y (see [17, Theorem 2.7, Chapter II]). × One can define several useful maps on jet bundles.

Definition A.5.2. Let X and Y be smooth manifolds and let k be a nonnegative integer. Define the source map πs : J k (X ,Y ) X by → πs([x, f ]) x = and the target map πt : J k (X ,Y ) Y by → πt ([x, f ]) f (x). = Let 0 l k and define the l-jet projection π : J k (X ,Y ) J l (X ,Y ) by ≤ ≤ l → π ([x, f ]) [x, f ]. l = Note that two different equivalence relations are used in the preceding equation, l and that πl is well-defined since the equivalence relation defining J (X ,Y ) is coarser than the one defining J k (X ,Y ). For each point x X , there is a canonical map k k ∈ k j : C ∞(X ,Y ) J (X ,Y ) defined by f [x, f ]. We will call j f the 1-jet of f at x. x → 7→ x We will sometimes denote this by j k f (x): j k f , and call j k f the 1-jet of f . = x 72 APPENDIX A. BACKGROUND

A.5.2 Jets of sections of a smooth fiber bundle

To define jets of sections of a smooth fiber bundle one proceeds similarly.

Definition A.5.3. Let E B be a smooth fiber bundle, and let k be a nonnegative k → integer. Let J E be the set of equivalence classes on the set X Γ∞(E) of sections, × under the relation defined by (x ,σ ) (x ,σ ) if 1 1 ∼ 2 2 • x x , and 1 = 2 • σ (x ) σ (x ), and 1 1 = 2 2 • after choosing coordinates for a local trivialization of E around x x and 1 = 2 σ (x ) σ (x ), and locally viewing the sections σ as functions f from the 1 1 = 2 2 i i trivializing neighborhoods to the fiber, the partial derivatives of f1 and f2 at x x agree up to order k. 1 = 2 We may define the source, target and l-jet projection maps as in the previous section. Note that for sections of a bundle, there is an identification of J 0E with E itself, and under this identification the 0-jet projection π : J k E J 0E agrees with 0 → the target map πt : J k E E. As was the case for the space of jets of functions, the → space J k E has a natural structure as a fiber bundle over B.

A.6 Whitney topologies

Definition A.6.1. A subset of a topological space is residual if it is a countable in- tersection of open dense sets. A topological space is a Baire space if every residual set is dense.

Let X and Y be smooth manifolds. We will define a family of topologies on k C ∞(X ,Y ) called the Whitney C topologies. We will follow [17, §3 Chapter II].

Definition A.6.2. Let X and Y be smooth manifolds. Fix some nonnegative integer k. For each open subset U of J k (X ,Y ), denote

k n k o M (U) g C ∞(X ,Y ) j f (X ) U . = ∈ | ⊆ The collection n o M k (U) U J k (X ,Y ) is open | ⊆

now forms a basis for a topology on C ∞(X ,Y ). We call this topology the Whitney C k topology.

Remark A.6.3. This topology is also known at the strong or fine C k topology. A.6. WHITNEY TOPOLOGIES 73

Definition A.6.4. Let X and Y be smooth manifolds. The collection n o M k (U) k 0 and U J k (X ,Y ) is open | ≥ ⊆ forms a basis for a topology on C ∞(X ,Y ). We call this topology the Whitney C ∞ topology.

We need the fact that the Whitney C ∞ makes C ∞(X ,Y ) a Baire space. A proof of this fact can be found in [17, Proposition 3.3 in Chapter II].

Proposition A.6.5. Let X and Y be smooth manifolds. The space C ∞(X ,Y ) en- dowed with the Whitney C ∞ topology is a Baire space.

Corollary A.6.6. Let X and Y be smooth manifolds. The space C ∞(X ,Y ) endowed with the Whitney C r topology for any r 1 is a Baire space. ≥ AssortedB proofs

B.1 Complete generators

The following is a straightforward generalization of Lemma 8.5.5 in [19], and the proof follows that of [19] but contains significantly more details.

Lemma B.1.1. Let S be an achronal hypersurface in a spacetime (M,g). Let γ be a

null geodesic segment contained H +(S). Suppose that γ has no past endpoint and is totally past imprisoned in some compact set K . Then γ is complete in the past direction.

Proof. Let γ have an affine parametrization. Suppose to get a contradiction that γ

is incomplete to the past, i.e. that the domain of γ has some infimum v0. We may without loss of generality (by translation of the parameter of γ and restriction of γ to a smaller domain) assume that γ has domain (v ,0] and that γ(t) K for all 0 ∈ t (v ,0]. ∈ 0 The idea is now to show that if γ is past incomplete, then a small perturbation of it yields a past inextendible timelike curve with contradictory properties. To help with this, we introduce an auxiliary Riemannian metric: Let W be an open

neighborhood of H +(S) K with compact closure. Let V be a future-directed unit ∩ timelike vector field on M such that V 0 everywhere on W. Define a metric ∇V = g 0 by

g 0(X ,Y ) g(X ,Y ) 2g(X ,V )g(Y ,V ). = +

This metric is positive definite. To see this, let X be nonzero and compute g 0(X , X ) in a basis (V,e1,e2,e3) orthonormal for g:

2 0 2 1 2 2 2 3 2 0 2 g 0(X , X ) g(X , X ) 2(g(X ,V )) ( (X ) (X ) (X ) (X ) ) 2(X ) 0. = + = − + + + + >

Let α (t) γ(v(t)) be a reparametrization of γ such that g(α˙ ,V ) 1/p2. 0 = 0 = − This implies that v is strictly increasing. For convenience, suppose also that v(0) = B.1. COMPLETE GENERATORS 75

0. With this choice, the curve length in g 0 of α0 on the parameter interval [a,b] is

Z b q Z b q 2 g 0(α˙0(t),α˙0(t))dt g(α˙0(t),α˙0(t)) 2(g(α0(t),V )) dt a = a + Z b r 1 0 2 dt b a = a + 2 = − so α0 is parameterized by arc length in the Riemannian metric g 0. Since γ has no past endpoint, α0 does not have one either. We will later construct a variation α of α0, and the computations will be done along the two-parameter map α.

Claim I: The domain of α0 is not bounded from below Suppose for contradiction that the domain of α were bounded below. Let a 0 > −∞ be the infimum of the domain of α0. Recall that a Riemannian metric induces a distance function defined as the infimum of the lengths of curves from one point to the other. Then for any sequence a a it holds that α (a ) is a Cauchy se- n → 0 n quence with respect to the distance function induced by g 0 (for α0 is a curve of length a a a a 0 connecting α (a ) to α (a )). The sequence | n − m| < | min(m,n) − | → 0 n 0 m α0(an) is also contained in the compact set K , and so has a convergent subse- quence. These two statements together imply that α0(an) is convergent for any sequence an a so the limit limt a α0(t) exists contradicting the fact that α0 → → + has no past endpoint. Hence the domain of α0 is not bounded from below.

Proof part II(Relations between α0 and γ): Since α0 is a reparametrization of a geodesic, ˙ α˙ is parallel to α˙ . In other words, there is a function f :( ,0) R ∇α0 0 0 −∞ → such that

˙ α˙ (t) f (t)α˙ (t), t ( ,0). ∇α0(t) 0 = 0 ∀ ∈ −∞ It also holds that

v0(t)γ˙(v(t)) α˙ (t), t ( ,0). = 0 ∀ ∈ −∞ Now v00(t) f (t)α˙0(t) α˙ 0 α˙0 α˙ 0 (v0γ˙) α0(v0)γ˙ v0 γ˙γ˙ α˙0(t) = ∇ = ∇ = + ∇ = v0(t) so v00 f . = v0 Note also that f is bounded. This can be seen by computing the norm of f α˙0 in the metric g 0.

¡ ¢ f p2g(f α˙ ,V ) p2g( ˙ α˙ ,V ) p2 ˙ g(α˙ ,V ) g(α˙ , ˙ V ) = − 0 = − ∇α0 0 = − ∇α0 0 − 0 ∇α0 ³ ´ p2 α˙ ( 1/p2) g(α˙ , ˙ V ) p2g(α˙ , ˙ V ). = − 0 − − 0 ∇α0 = 0 ∇α0 76 APPENDIX B. ASSORTED PROOFS

This shows that f can be defined in terms of g, α˙ and V . The coefficients of these in coordinate patches are all bounded since α˙ is a unit vector field in g 0. Since H +(S) K is compact it can be covered by finitely many coordinate patches, and ∩ hence f is bounded.

Claim III: v0 is bounded Since γ is incomplete to the past, v is bounded below. In other words, the integral Z t v(t) v0(τ)dτ = 0 is bounded. This of course implies that liminft v0(t) 0, since v is strictly →−∞ = increasing. We will now show that boundedness of v on ( ,0] together with −∞ v 00 boundedness of f implies that v0 is bounded. Suppose not. Since v0 is con- = v 0 tinuous, it can only be unbounded on ( ,0] if limsupt v0(t) . Since we −∞ →−∞ = ∞ also know that liminft v0(t) 0 and that v0 is continuous there are, for arbi- →−∞ = trarily large C 0, sequences t ,s such that > n n → −∞

tn 1 sn tn for all n, + < <

v0(t ) 2C, n = v0(s ) C n = and

C v0(t) 2C if t (s ,t ). ≤ ≤ ∈ n n By the mean value theorem of calculus, there is for each n some τ [s ,t ] such n ∈ n n that v0(tn) v0(sn) C v00(τn) − . = t s = t s n − n n − n However ¯Z ¯ X∞ ¯ −∞ ¯ C(tn sn) ¯ v0(τ)dτ¯ n 0 − ≤ ¯ 0 ¯ < ∞ = so (t s ) 0 as n . Hence n − n → → ∞

lim f (τn)v0(τn) lim v00(τn) . n n →∞ = →∞ = ∞ Since v0(τ ) [C,2C] for all n, this implies that f (τ ) , contradicting the fact n ∈ n → ∞ that f is bounded. Hence v0 must be bounded.

Proof part IV(Construction of a variation α of α0): We will now construct a varia- tion α of α0. The idea is to push α0 to the past and make it timelike, and then de- rive a contradiction from the resulting curve. Let x :( ,0) R denote a smooth −∞ → positive function which will be fixed later. Let

α:( δ,δ) ( ,0) H +(S) − × −∞ → (u,t) α(u,t) 7→ B.1. COMPLETE GENERATORS 77 be a smooth map such that

∂α α(0, ) α0 and (u,t) x(t)Vα(u,t). (B.1) · = ∂u = −

To see that such a variation exists, note that the conditions can be viewed as a family of ordinary differential equations in u, parameterized by t. This means that a smooth solution exists for each t for some small δ 0 (depending on t) by t > the existence theorem and theorem about smooth dependence on initial values for ordinary differential equations. To claim that the necessary variation exists, we need to show that the existence times δt can be uniformly bounded from below by some δ 0 independent of t. However, we know that a solution to the differential > equation exists as long as it stays in the compact set W. Since H +(S) K is compact ∩ and W open, the g 0 distance between H +(S) K and M \ W is positive. Since V is ∩ bounded, and x will be bounded when we choose it, there is a positive uniform lower bound for the time after which a solution may leave W. This means that there is a uniform lower bound for the existence times of the solutions of the family of ordinary differential equations defining the variation. Hence a variation with the desired properties exists. When we later fix the function x, it will be bounded.

Since H +(S) K is compact and contained in the open set W, the distance between ∩ H +(S) K and M \ W in the metric g 0 is positive. This means (since g 0(V,V ) is ∩ bounded) that there is some δ 0 such that the image of α is contained in W. We > will choose such a value for δ. For ease of notation, let α denote the curve α(u, ). We now wish to choose the u · (positive) function x in such a way that some curve α² is timelike. In other words, we want there to be some ² 0 such that the function >

y(u,t) g(α˙ (t),α˙ (t)) = u u is negative for u ² and all t ( ,0). To show that this is the case, we will ¯ = ∈ −∞2 ∂y ¯ ∂ y compute and a bound for 2 , and from this obtain an upper bound for ∂u ¯u 0 ∂u y. Choosing a= suitable function x will make this upper bound negative for small values of u. ∂y Proof part V(Computation of ): Let U denote the pushforward through α of ∂u the coordinate vector field ∂ on ( δ,δ) ( ,0). We will not always write out the ∂u − × −∞ dependence on t and u. The first partial derivative of y can be computed as

∂y ∂ (u,t) g (α˙u(t),α˙u(t)) Ug(α˙u,α˙u) 2g( U α˙u,α˙u) ∂u = ∂u = = ∇

¡ ¢ 2g( ˙ U,α˙ ) 2 ˙ g(U,α˙ ) g(U, ˙ α˙ ) = ∇αu u = ∇αu u − ∇αu u 78 APPENDIX B. ASSORTED PROOFS

where α˙ ˙ U since U and α˙ are pushforwards of coordinate vector fields. ∇U u = ∇αu u Evaluating this at u 0 we see that = ∂y ¡ ¢ (0,t) 2 α˙ g( xV,α˙0) g( xV, α˙ α˙0) ∂u = ∇ 0 − − − ∇ 0 ¡ ¢ 2 ˙ (xg(V,α˙ )) xg(V, ˙(v0γ˙)) = −∇α0 0 + ∇v 0γ µ 1 ¶ 2 α˙0(x) xv0g(V, γ˙(v0γ˙)) = p2 + ∇ µ ¶ 1 2 x 2 α˙0(x) x(v0) g(V, γ˙(γ˙)) α˙0(v0)g(V,α˙0) = p2 + ∇ + v0 x(t)v00(t) p2x0(t) p2 = − v0(t) d µ x(t) ¶ p2v0(t) , = dt v0(t) where we have used that α˙ v0γ˙, g(V,α˙ ) 1/p2 and ˙γ˙ 0. 0 = 0 = − ∇γ = ∂2 y Proof part VI(An upper bound for ): We now compute an upper bound for ∂u2 the second partial derivative of y with respect to u. For convenient notation, we use the vector fields µ ∂ ¶ T α∗ , = ∂t µ ∂ ¶ U α∗ . = ∂u Note that T α˙ (t) α(u,t) = u and U x(t)V . α(u,t) = − α(u,t) Now 1 ∂2 1 ∂2 1 ∂2 ∂ y(u,t) g (α˙u(t),α˙u(t)) g (T,T ) g( U T,T ) 2 ∂u2 = 2 ∂u2 = 2 ∂u2 = ∂u ∇ g( T, T ) g( T,T ) g( U, U) g( U,T ) = ∇U ∇U + ∇U ∇U = ∇T ∇T + ∇U ∇T g( U, U) g( U,T ) g(R(U,T )U,T ) = ∇T ∇T + ∇T ∇U + where we have used that T U since U and T are coordinate vector fields ∇U = ∇T and R(U,K ). ∇U ∇K = ∇K ∇U + We now compute each term separately.

Evaluating the first term at α(0,t) and using that T (x) α (x) x0 we get = 0 = g( U, U) g( (xV ), (xV )) g(T (x)V x V,T (x)V x V ) ∇T ∇T = ∇T ∇T = + ∇T + ∇T B.1. COMPLETE GENERATORS 79

¡ ¢2 2 x0(t) g(V,V ) 2x(t)x0(t)g(V, V ) x (t)g( V, V ) = + ∇T + ∇T ∇T ¡ ¢2 2 x0(t) (x(t)) g( V, V ). = − + ∇T ∇T We have used that g(V, V ) 0. That this is true is seen by noting that ∇T = 1/2 g(V, V ) T g(V,V ) g( V,V ) T ( 2− ) g(V, V ) g(V, V ) ∇T = − ∇T = − − ∇T = − ∇T so that g(V, V ) g(V, V ). For the second term, note that ∇T = − ∇T

2 ∂x UU U ( xV ) xU(x)V x V V x(t) V 0 0 ∇ = ∇ − = + ∇ = ∂u + = (since V 0 in the neighborhood W by choice of V , and x is independent of u) ∇V = so that g( U,T ) g( 0,T ) 0. ∇T ∇U = ∇T = The third term is simply

g(R(U,T )U,T ) g(R( xV,T )( xV ),T ) x2(t)g(R(V,T )V,T ). = − − = Hence

2 1 ∂ ¡ ¢2 2 ¡ ¢ g (α˙u(t),α˙u(t)) x0(t) (x(t)) g( T V, T V ) g(R(V,T )V,T ) 2 ∂u2 = − + ∇ ∇ +

x2 ¡g( V, V ) g(R(V,T )V,T )¢. ≤ ∇T ∇T + 2 2 We wish to bound this by C x g 0(T,T ) for some constant C on the neighborhood W of H +(S), which we chose to have compact closure. To see that this is possi- ble, view g( V, V ) g(R(V,T )V,T ) as a quadratic form in T . Its components ∇T ∇T + in coordinates depend on g, V and R, all of which are bounded in coordinate neighborhoods, and H +(S) K can be covered by finitely many such neighbor- ∩ hoods. Since the quadratic form g 0 is positive definite, there is some C such that g( V, V ) g(R(V,T )V,T ) g 0(T,T ). Hence ∇T ∇T + ≤ 2 ∂ 2 2 g (α˙u(t),α˙u(t)) C x g 0(T,T ) ∂u2 ≤ for some constant C. We want a bound in terms of g (α˙u(t),α˙u(t)) instead, so we compute ¡ ¢2 g 0(T,T ) g(T,T ) 2 g(V,T ) . = + Since ∂ g(V,T ) Ug(V,T ) g( x V V,T ) g(V, U T ) 0 g(V, T U) ∂u = = − ∇ + ∇ = + ∇

g(V,T (x)V x V ) T (x) xg(V, V ) x0(t) = − − ∇T = + ∇T = 80 APPENDIX B. ASSORTED PROOFS

(where as earlier g(V, V ) 0) we know that ∇T = ¯ 1 g(V,T ) ux0(t) g(V,T )¯u 0 ux0(t) . = + = = − p2

dx ¡ ¢2 When we choose x, we will make sure that dt is bounded, and then 2 g(V,T ) is bounded by some constant d for all small u. Hence we can convert our bound in terms of g 0(T,T ) to a bound in terms of g(T,T ):

2 ∂ 2 2 2 2 g (α˙u(t),α˙u(t)) C x g 0(T,T ) C x (g(T,T ) d). ∂u2 ≤ ≤ + In the notation of the function y, we now know that

∂2 y (u,t) (y(u,t) d)C 2(x(t))2. ∂u2 ≤ +

Claim VII: For all sufficiently small ² 0, the curve α is timelike > ² From our previous calculation we know that µ ¶ ∂y v0(t) d x(t) (0,t) . ∂u = p2 dt v0(t) Moreover, y(0,t) 0 since α is a lightlike curve. For each fixed t, this is a differen- = 0 tial inequality in the variable u. Let z be the solution of the differential equation resulting from replacing the inequality with equality:

∂2z (u,t) C 2x2(0,t)(z(u,t) d), ∂u2 = + ∂z ∂y (0,t) (0,t), ∂u = ∂u z(0,t) y(0,t) 0. = = 2 2 Integrating the inequality ∂ y (u,t) ∂ z (u,t) we see that ∂u2 ≤ ∂u2 ∂y ∂y ∂z ∂z (u,t) (0,t) (u,t) (0,t) ∂u − ∂u ≤ ∂u − ∂u so that ∂y ∂z (u,t) (u,t). ∂u ≤ ∂u Integrating once again and using the fact that z(0,t) y(0,t) 0 we have = = y(u,t) z(u,t). ≤ Solving the differential equation for z we see that

z(u,t) d cosh(C x(t)u) a(t)sinh(C x(t)u) d = + − B.1. COMPLETE GENERATORS 81 where µ ¶ v0(t) d x(t) a(t) . = p2C x(t) dt v0(t) Since d is nonnegative, an upper bound for z is

z(u,t) d cosh(C x(t)u) a(t)sinh(C x(t)u) d = + − (d tanh(C x(t)u) a(t))sinh(C x(t)u) d (d tanh(C x(t)u) a(t))sinh(C x(t)u). = + − ≤ + Hence y(u,t) (d tanh(C x(t)u) a(t))sinh(C x(t)u). ≤ + Recall that the idea was to choose the function x in such a way that there exists some ² 0 such that y(²,t) 0 for all t. We claim that a possible choice of x is > < v0(t) x(t) . = v(t) 2v − 0 Recall that

v0 lim v(t) = t →−∞ and that v is increasing so that

v v(t) 0 t ( ,0]. 0 ≤ ≤ ∀ ∈ −∞ We begin by making good on the promises we made about the function x: It should be positive, bounded, and have bounded derivative. The denominator in the def- inition of x is bounded from below by v and from above by 2v , and v is − 0 − 0 − 0 positive, so boundedness and positivity of x follow from boundedness and posi- tivity of v0. Computing the derivative of x we see that

2 v00(t) (v0(t)) v0(t)f (t) 2 x0(t) x (t) = v(t) 2v − (v(t) 2v )2 = v(t) 2v − − 0 − 0 − 0 Since x, v0 and f v00/v0 are bounded, so is x0. Having chosen x, we can now fix = the number δ 0 defining the domain of α such that the image of α is contained > in W. Recall that

y(u,t) (d tanh(C x(t)u) a(t))sinh(C x(t)u) ≤ + where µ ¶ v0(t) d x(t) a(t) . = p2C x(t) dt v0(t) With our present choice of x, x(t) a(t) . = −p2C v0(t) 82 APPENDIX B. ASSORTED PROOFS

The objective is to ensure that y(u,t) 0 for some positive u and for all t. Since < sinh(C xu) 0 for positive u, a sufficient condition is that ≥ x(t) d tanh(C x(t)u) 0 − p2C v0(t) < for some u 0 and all t. A series expansion tells us that > tanh(C x(t)u) C x(t)u O((ux(t))3) = + so that µ 1 ¶ d tanh(C x(t)u) a(t) dCu x(t) O(u3x3(t)). + = − p2C v0(t) +

Since v0 is bounded, there is some positive lower bound for 1/v0. Hence it holds for all sufficiently small u such that dCu 1/(p2C v0(t)) is negative for all t. Since − x is bounded, it further holds for all sufficiently small u that the O(u3x3(t)) term does not affect the sign: With such a choice of u, it holds that d tanh(C xu) a is + negative for all t, and hence

y(²,t) (d tanh(C x(t)²) a(t))sinh(C x(t)²) 0 ≤ + < for all values of t and all sufficiently small ² 0. Since > y(²,t) g(α˙ (t),α˙ (t)) = ² ² this shows that the curve α² is timelike.

Claim VIII: For all sufficiently small ² 0, the curve α has infinite g 0-length > ² For each (negative) integer k, let Lk (u) be the g 0-length of the restriction of αu to [k,k 1]. By the formula for the first variation of arc length ([28, Proposition 10.2]) + Z k 1 + ¯k 1 + L0k (0) g 0( α˙ 0 α˙0,V )dt g 0(α˙0,V )¯k = − k ∇ + Z k 1 + ¯k 1 f (t)g 0(α˙0,V )dt g 0(α˙0,V )¯k+ = − k + Z k 1 f (t) + dt. = − k p2

We here used that g 0(α ,V ) 1/p2 by definition of g 0 and α . Since f is bounded, 0 = 0 we know that L0k (0) is bounded uniformly in k. This means that for all sufficiently small ² 0 it holds that L (²) 1/2 for all k. (Recall that L (0) 1 since α is > k > k = 0 parameterized by arc length.) This means that the length of α² is X X Lk (²) 1/2 . k 0 ≥ k 0 = ∞ < < B.1. COMPLETE GENERATORS 83

Claim IX: For all sufficiently small ² 0, the curve α belongs to the interior of > ² D+(S0) Since the curve α for ² 0 is a variation to the past of α , it belongs to the open set ² > 0 I −(H +(S )). We will first show that I −(H +(S )) I +(S ) D+(S ), and then show 0 0 ∩ 0 ⊆ 0 that α belongs to I −(H +(S )) I +(S ) for all sufficiently small ² 0. We will then ² 0 ∩ 0 > have shown that α² belongs to an open set contained in D+(S0), and hence it must belong to the interior of D+(S0). Let p I −(H +(S )) I +(S ). We will first show that p D (S ), and then that ∈ 0 ∩ 0 ∈ + 0 p D+(S ). That p I −(H +(S )) means that there is some future-directed timelike ∈ 0 ∈ 0 curve λ from p to H +(S0). This curve cannot pass S0, since p lies to the future of S0 and S0 is achronal. Suppose now that κ is a future-directed past inextendible time- like curve with future endpoint p. By concatenating κ and λ and smoothing (in a neighborhood of p which is disjoint from S , which exists since p I +(S )) we 0 ∈ 0 obtain a past-inextendible timelike curve with future endpoint in H +(S0). Since H +(S ) D (S ), this combined curve must intersect S . Since the curve λ does 0 ⊆ + 0 0 not intersect S0, the curve κ must do so. This proves that every past-inextendible timelike curve κ through p must intersect S , so that p D (S ). By the same ar- 0 ∈ + 0 gument, all points in the interior of λ belong to D+(S0). Let q be some point in the interior of λ. Since λ is timelike, q I +(p). Since I +(p) is open, it is a neigh- ∈ borhood of q. Since q D (S ) it is a limit point of D+(S ). This means that the ∈ + 0 0 neighborhood I +(p) of q must contain some point r D+(S0) I +(p). Let λb be ∈ ∩ a future-directed timelike curve from p to r . Now let κ be a future-directed past inextendible causal curve with future endpoint p. Concatenating κ with λb and smoothing (again in a neighborhood of p which is disjoint from S0) we obtain a past-inextendible causal curve with future endpoint r . Since r D+(S ), this ∈ 0 curve must intersect S0. Since r I +(p) I +(S0) and S0 is achronal, the curve λb ∈ ⊆ cannot intersect S0. This means that κ must intersect S0. This proves that every past-inextendible causal curve κ through p must intersect S , so that p D+(S ). 0 ∈ 0 Since α is a curve in H +(S ) and α is a variation to the past for ² 0 we know 0 0 ² > that α belongs to I −(H +(S )). Since g 0(V,V ) 1 and x is bounded by 2v , equa- ² 0 = | 0| tion (B.1) implies that the distance from a point on α to α cannot exceed 2v ² . ² 0 | 0 | Since H +(S ) K is compact and disjoint from the closed set S , the g 0-distance 0 ∩ 0 from H +(S ) to S is positive. Choosing ² 0 so small that 2v ² is smaller than 0 0 > | 0 | this distance, we know that α does not intersect S . To see that α (t) I +(S ) ² 0 ² ∈ 0 for some t, note that no curve α with 0 u ² can intersect S so that the u ≤ ≤ 0 timelike curve λ:[ ²,0] M defined by λ(u) α u(t) does not intersect S0. Ex- − → = − tend λ to some past inextendible timelike curve. Then λ is a past inextendible timelike curve with future endpoint λ(0) α (t) H +(S ), so λ must intersect = 0 ∈ 0 84 APPENDIX B. ASSORTED PROOFS

S . Since λ passes through α (t), we know that α (t) I +(S ). Since t was ar- 0 ² ² ∈ 0 bitrary, we have now shown that the image of α belongs to I −(H +(S )) I +(S ) ² 0 ∩ 0 for all sufficiently small ² 0. As noted previously, this together with the fact that > I −(H +(S )) I +(S ) D+(S ) shows that the image of α belongs to the interior of 0 ∩ 0 ⊆ 0 ² D+(S0). Proof part X(Contradiction ensues): We have now shown that if we choose ² 0 > small enough, then α² is a timelike curve of infinite g 0-length, contained in the interior of D+(S0). Since it has infinite g 0-length and belongs to the compact set W, it must have a limit point. Since it is timelike, the existence of this limit point implies that the strong causality condition cannot hold in any open neighborhood

of α². However, the interior of D+(S0) is an open neighborhood of α² satisfying the strong causality condition (by Proposition A.2.21), so we have arrived at a contra- diction. Hence γ cannot be incomplete in the past direction.

B.2 Geodesically spanned null hypersurfaces

The following is a version of [28, Lemma 8.6] for geodesics normal to a submani- fold. We also note which regularity is necessary for the result to hold.

Proposition B.2.1. Let (M,g) be a spacetime of dimension n 1 and let N M be a + ⊂ C 2 submanifold of codimension 2. Let n denote a C 1 normal null vector field along N. Consider the normal exponential map exp: R N M defined by × → exp(t,p) exp (tn ) = p p where exp is the exponential map at the point p. Suppose that O R N is an p ⊂ × open subset such that H : exp(O) is an embedded C 1 hypersurface in M and the = pushforward exp is injective on O. Then H is a null hypersurface. ∗ Proof. Choose a point q exp(t,p) H and let γ denote the null geodesic s = ∈ 7→ exp(s,p). Our goal is to show that every vector W T H is orthogonal to γ˙(t), ∈ q thereby proving that Tq H is a null hyperplane. Since exp is injective at exp(t,p), it is also surjective for dimensional rea- ∗ sons. This means that W has some preimage in T (R N). Denote this preim- (t,p) × age by (ζ, Z ), where we make use of the canonical isomorphism T(t,p) (R N) × =∼ Tt R Tp N. The pushforward is linear so exp (ζ, Z ) exp (ζ,0) exp (0, Z ). Note × ∗ = ∗ + ∗ that exp (ζ,0) is tangent to γ, so g(exp (ζ,0),γ˙(t)) 0. Hence ∗ ∗ = g(W,γ˙(t)) g(exp (ζ,0) exp (0, Z ),γ˙(t)) g(exp (0, Z ),γ˙(t)). = ∗ + ∗ = ∗ Let α:( 1,1) N be a curve with α(0) p and α˙(0) Z . Consider the two- − → = = parameter map x(s,u) exp(st,α(u)) = B.3. GENERALIZED JET TRANSVERSALITY 85

defined for s [0,1] and u (0,1). Let V be a vector field along γ defined by ∈ ∈ V (s) x (s,0). = u Each curve s x(s,u) is a geodesic, so the map x is a variation through geodesics. 7→ Hence V is a Jacobi vector field. The curve u x(0,u) is contained in N so V (0) is 7→ tangent to N. By assumption on n, the vector γ˙(0) is orthogonal to N, so

g(V (0),γ˙(0)) 0. =

Let T denote the vector field xs along the map x. Partial derivatives of two-parameter maps commute by [28, Proposition 4.44] so

V 0(0) x (0,0) x (0,0) T. = us = su = ∇Z Hence 1 g(V 0(0),T ) g( Z T,T ) Z g(T,T ) 0 = ∇ = 2 = since T is tangent to null curves. Since x (0,0) γ˙(0) we have shown that s =

g(V 0(0),γ˙(0)) 0. =

By [28, Lemma 8.7], the fact that V (0) and V 0(0) are both orthogonal to the geodesic γ, together with the fact that V is a Jacobi field along γ, implies that V (s) is orthog- onal to γ for all s. In particular,

g(V (1),γ˙(t)) 0. = Computing V (1) we see that

V (1) xu(1,0) exp (0,α˙(0)) exp (0, Z ). = = ∗ = ∗ Hence g(W,γ˙(t)) 0 = for all W T H. Since q exp(t,p) was arbitrary, this shows that each tangent ∈ q = plane of H is a null hyperplane, so that H is a null hypersurface.

B.3 Generalized Jet Transversality

The notation for jet bundles used in this section is defined in Section A.5. Recall in particular that each jet bundle has an associated source map and target map. We will call these maps πs and πt , respectively. We need a version of the Thom Transversality Theorem which applies to sec- tions of a bundle. To prove this adapted theorem, we need the following lemma, which can be found as [17, Lemma 4.6, Chapter II], or as a special case of [20, The- orem 2.7, Chapter 3]. 86 APPENDIX B. ASSORTED PROOFS

Lemma B.3.1. Let X , B and Z be smooth manifolds with V a submanifold of Z . Let

ρ : B C ∞(X ,Y ) be a mapping (not necessarily continuous) and define Φ: X B → × → Y by Φ(x,b) ρ(b)(x). Assume that Φ is smooth and transverse to V . Then the set = {b B ρ(b) V } is dense in B. ∈ | t We will also need the following straightforward generalization of Proposition 4.5 in Chapter II of [17].

Proposition B.3.2. Let X and Y be smooth manifolds and let W be a submanifold of Y . Let C be a compact subset of W . Let

T {f C ∞(X ,Y ) f W on C}. C = ∈ | t

1 Then TC is an open subset of C ∞(X ,Y ) in the Whitney C topology.

Proof. For any subset V J 1(X ,Y ), let ⊆ 1 M(V ) {f C ∞(X ,Y ) j f (X ) V }. = ∈ | ⊆

Recall that the sets M(V ) where V J 1(X ,Y ) is open form a basis of the Whitney 1 ⊆ C topology on C ∞(X ,Y ). 1 Note that if σ J (X ,Y ) has source x, then the tangent map (f )x : Tx X ∈ ∗ → T f (x)Y is independent of the choice of representative f of σ. This means that it makes sense to speak of a 1-jet (rather than a representative of it) being transverse to W at its source. Define the set Ω J 1(X ,Y ) by ⊆

Ω {σ J 1(X ,Y ) πt (σ) C or σ is transverse to W at πs(σ)}. = ∈ | ∉

Note that f W on C if and only if j 1 f (X ) Ω. Hence T M(Ω), so T is open tp ⊆ C = C if Ω is open. We will show that J 1(X ,Y )\ Ω is closed. By definition of Ω

J 1(X ,Y )\ Ω {σ J 1(X ,Y ) πt (σ) C and σ is not transverse to W at πs(σ)}. = ∈ | ∈

1 1 Consider a sequence (σ )∞ in J (X ,Y )\Ω, converging to some σ J (X ,Y ). Since i i 1 ∈ the target πt (σ ) of each σ= is in C and C is a closed set, πt (σ) C as well. The goal i i ∈ is now to show that σ is not transverse to W at πs(σ), for then σ J 1(X ,Y )\ Ω ∈ s and we are done. Choose closed coordinate neighborhoods UX and UY of π (σ) and πt (σ), respectively, with surjective charts ψ : U Dn and ψ : U Dm. X X → Y Y → Here Dk denotes the closed unit k-disk. We may choose ψ so that ψ (W U ) Y Y ∩ Y = Dm Rk where Rk Rm denotes a fixed choice of k-dimensional subspace. Let J ∩ ⊆ = {τ J 1(X ,Y ) πs(τ) U and πt (τ) U }. Since U and U are neighborhoods of ∈ | ∈ X ∈ Y X Y πs(σ) and πt (σ), all but finitely many of the σ belong to J. Each τ J determines i ∈ B.3. GENERALIZED JET TRANSVERSALITY 87 a linear map from Rn to Rm±Rk in the following way. Let π: Rm Rm±Rk denote → the projection. Define η: J Hom(Rn, Rm±Rk ) by → 1 η([g]) (π ψY g ψ−X ) . = ◦ ◦ ◦ ∗ This definition is independent of the particular representative g of [g] J, since ∈ it depends only on the tangent map of g. Note that η([g]) can be thought of as the coordinate representation of the tangent map of g, composed with the linear projection onto the tangent space of W . The nice thing about the map η is that η(τ) is surjective if and only if τ is transverse to W at πs(τ). By noting that η([g]) 1 = (π ψY ) g (ψ− ) we see that η([g] depends continuously on g , which in turn ◦ ∗ ◦ ∗ ◦ X ∗ ∗ depends continuously on [g], so η is continuous. Note now that the set

F {A Hom(Rn, Rm±Rk ) A is not surjective} = ∈ | is (after a choice of bases) the zero set of an algebraic equation involving determi- 1 1 nants, and as such closed. This means that η− (F ) is also closed. Since σ η− (F ) i ∈ for all i (except possibly for the finitely many which do not belong to J) by assump- 1 tion, it holds that σ η− (F ). This means that η(σ) is not surjective, so that σ is not ∈ transverse to W at πs(σ). This means that σ J 1(X ,Y )\ Ω, and so we have shown ∈ that Ω is open. As noted in the beginning of the proof, this means that TC is open, and so we are done.

We now state the transversality theorem we need.

Proposition B.3.3. Let L M and K M be a smooth fiber bundles over a smooth → → manifold. Suppose that W is a submanifold of the product bundle J r L K . Endow r 1 ⊕ Γ∞(L) with the Whitney C + topology and let

r T {φ Γ∞(L) (j φ π ,id ) W } W = ∈ | ◦ M K t where π : K M denotes the projection. Then T is a residual subset of Γ∞(L). M → W Proof. Our proof mimics the proof of the Thom Transversality Theorem given in [17, Theorem 4.9, Chapter II]. We begin by introducing some notation. The spaces L, K , J r L and all fiber bundle products of them are fiber bundles over M. We will denote the fiber bundle projection maps from each of these bundles onto M by a single symbol πM , relying on context to infer which map is intended. (Note that π : J r L M agrees with the source map πs : J r L M. We shall not use the M → → latter notation in the present proof.) We extend the target map πt : J r L L to → a map π˜ t : J r L K L K by letting it be the identity in the second component: ⊕ → ⊕ π˜ t (σ,k) (πt (σ),k). = 88 APPENDIX B. ASSORTED PROOFS

Proof part I(Localization): The goal is to express TW as a countable intersection of open dense subsets of Γ∞(L). To do this, we begin by covering W by countably many open subsets W1,W2,... such that each Wγ satisfies that

• the closure W of W in J r L K is contained in W , γ γ ⊕

• Wγ is compact,

• there is (where FL and FK denote the fibers of L and K ) an open trivializing coordinate neighborhood M L K M F F of the image of W γ × γ × γ ⊆ × L × K γ under the map π˜ t : (πt ,id ): J r L K L K , = K ⊕ → ⊕

• Mγ is compact.

To see that it is possible to construct such sets, note that it is sufficient to construct such a neighborhood of each point, since by second countability of W every open cover of W has a countable subcover. To choose a neighborhood Wp of a point p, begin by choosing an open trivializing coordinate neighborhood M L K p × p × p ⊆ M F F of (πt ,id )(p). Shrinking M if necessary, we may make M compact. × L × K K p p The inverse image of M L K under (πt ,id ) is an open neighborhood V of p × p × p K p, and V W is open in W since W is an embedded submanifold. We may then ∩ choose W W to be an open neighborhood of p such that W is compact and p ⊆ p contained in V W , since W is an embedded submanifold. This completes the ∩ construction. For each γ, let

r T {φ Γ∞(L) (j φ π ,id ) W on W }. γ = ∈ | ◦ M K t γ

If we can show that each Tγ is open and dense, then TW is residual since \ TW Tγ. = γ

Fix γ for the remainder of the proof. Showing that Tγ is open can be done by an application of Proposition B.3.2. We do this first, and then devote the rest of the proof to showing that Tγ is dense.

Claim II: Tγ is open We begin by showing that Tγ is open. Let

r T {f C ∞(K , J L K ) f W on W }. = ∈ ⊕ | t γ r Then Tγ is the inverse image of T under the map Γ∞(L) C ∞(M, J L) defined r → by φ (j φ πM ,idK ). This map is continuous when Γ∞(L) is given the Whit- 7→r 1 ◦ r 1 ney C + topology and C ∞(M, J L) is given the Whitney C topology. It is then B.3. GENERALIZED JET TRANSVERSALITY 89

1 r sufficient to show that T is open in the Whitney C topology on C ∞(M, J L). By Proposition B.3.2 (taking X K and Y J r L K and C W ) this is indeed the = = ⊕ = γ case.

Proof part III(Construction of local polynomial perturbations): We will now show that Tγ is dense in Γ∞(L). The strategy we employ involves local polynomial per- turbations of sections of L. We begin by describing the construction of such poly- nomial perturbations. They will be defined with respect to

• a choice of coordinate chart ψ : M Rn, M γ → • a choice of coordinate chart ψ : L Rm, L γ → • a choice of smooth cut-off function ζ : M [0,1] such that M → – ζ (x) 1 for all x in some neighborhood of π (W ), M = M γ – ζ (x) 0 if x M , M = ∉ γ • a choice of smooth cut-off function ζ : Rm [0,1] such that L → – ζ (x) 1 for all x in some neighborhood of ψ (p (π˜ t (W ))), where L = L L γ p : M L K L denotes the projection, L γ × γ × γ → γ – ζ (x) 0 if x ψ (L ). L = ∉ L γ

Fix such functions ψM , ψL, ζM and ζL. Note that a choice of cut-off function ζM exists since πM (Wγ) is compact and contained in the open set Mγ, and similarly for ζ . A section σ Γ∞(L) can locally on the trivializing coordinate neighborhood L ∈ M be viewed as a function f : M F . We would like to perturb it by a polyno- γ γ → L mial function p : Rn Rm. The straightforward way of doing this in coordinates → is forming θ : x ψ (f (x)) p(ψ (x)). However, there is no guarantee that the 7→ L + M image of this ends up in the image of ψL, so we will not be able to lift this function back to the manifold. Moreover, we would have perturbed f close to the boundary of Mγ and so would not be able to patch the perturbed function together with the original section. This is where the cut-off function come in. We define the pertur- bation of f : M F by a polynomial function p : Rn Rm (with respect to ψ , γ → L → M ψ , ζ and ζ ) to be the function f : M F defined by L M L p γ → L  f (x) if f (x) Lγ, fp (x) ∉ 1 ¡ ¢ = ψ− ψ (f (x)) τ(x)p(ψ (x)) otherwise, L L + M where the auxiliary function τ is defined by

τ(x) ζ (x)ζ (θ(x)), = M L 90 APPENDIX B. ASSORTED PROOFS

θ(x) ψ (f (x)) p(ψ (x)). = L + M The idea is that fp (x) is a convex combination of f (x) and what by abuse of nota- tion can be written f (x) p(x), where the interpolation parameter is determined + by the cut-off functions. The function fp has a couple of notable features:

• fp and f agree smoothly at the boundary of Mγ, in the sense that they have the same smooth extensions to neighborhoods of Mγ.

• Suppose that p is a polynomial. Then there is some neighborhood of the set π (W ) p (πt (W )) such that if (x, f (x)) belongs to this neighborhood M γ × L γ and p(ψM (x)) is sufficiently small then the coordinate representation of fp is precisely ψ (f (x)) τ(x)p(ψ (x)). In particular, for each jet ν J r (M ,F ) L + M ∈ γ L with source x π (W ) and target f (x) p (πt (W )) there is a polynomial ∈ M γ ∈ L γ (of degree at most r ) such that j r f (x) ν. = The first observation means that we may extend this construction to a local pertur- bation of a section, rather than of a local section. Let σ Γ∞(L) have the local rep- ∈ resentation f : M F in the trivializing neighborhood M . Let i : M F L γ → L γ γ × L → denote the local trivialization. Define the perturbation of σ by a polynomial func- tion p : Rn Rm (with respect to ψ , ψ , ζ and ζ ) to be → M L M L  σ(x) if x Mγ, σp (x) ∉ = i(x, fp (x)) otherwise.

Since f and fp agree smoothly at the boundary of Mγ (by the use of the cut-off function ζM in the definition of fp ) this is a smooth section. By the second ob- r servation about f it holds that any for a fixed section σ Γ∞(L), any jet ν J L p ∈ ∈ with source x π (W ) and target σ(x) p (πt (W )) may be realized as j r σ for ∈ M γ ∈ L γ p some polynomial p of degree at most r . This is the crucial observation; a finite- dimensional family of perturbations are sufficient to realize any r -jet.

Proof part IV(Application of Lemma B.3.1): We return to our goal of showing that

T is dense in Γ∞(L). To this end, we fix some σ Γ∞(L) and aim to show that this γ ∈ σ is a limit point of T . Let P be the set of polynomial functions Rn Rm of degree γ → at most r . The set P has a conventional structure of a finite-dimensional vector space, and so has a canonical topology. Endow P with this topology. With P 0 P r ⊆ being a subset to be described later, define ρ : P 0 C ∞(K , J L K ) by → ⊕ ρ(p)(x) (j r σ (π (x)),x). = p M

We will later show that it is possible to choose P 0 P such that ⊆

• P 0 is open, B.3. GENERALIZED JET TRANSVERSALITY 91

•0 P 0, ∈

r • each point of W is a regular value of the evaluation map Φ: K P 0 J L K γ × → ⊕ defined by Φ(x,p) ρ(p)(x). In particular, Φ is transverse to W on some = open neighborhood of Wγ.

To see that we may not simply choose P 0 P, consider polynomial p which is so = large that σ σ for all q in a neighborhood of p. (The reason this can happen is q = the use of the cut-off function ζL in the definition of σq .) Then (p,t) will typically not be a regular point of Φ. Suppose for the moment that we have chosen a set

P 0 satisfying the stated hypotheses. Then P 0 inherits a smooth manifold structure from P. The map Φ is smooth since πM is smooth and the partial derivatives in co- ordinates (in particular those up to order r ) of σp (m) depend smoothly on m and p. Let V be an open neighborhood of Wγ on which Φ is transverse to W . (We will show in the next paragraph that such a V exists.) Then V is an open submanifold of W , so Φ being transverse to W on V is equivalent to Φ being transverse to V . We can now apply Lemma B.3.1 to find that there is a dense subset of P 0 such that if p belongs to this subset then ρ(p) V . In particular (since 0 P 0) there is a sequence t ∈ (pi )i∞ 1 in P 0 converging to 0 such that ρ(pi ) t V for each i. That ρ(pi ) t V means = that ρ(pi ) t W on V which in turn means that ρ(pi ) t W on Wγ. By definition of r 1 T this means that σ T for each i. Since σ σ in the Whitney C + topol- γ pi ∈ γ pi → ogy (in fact in most reasonable topologies on Γ∞(L)), we have shown that σ is a limit point of Tγ. Since σ was arbitrary, this shows that Tγ is dense. Once we have constructed the set P 0 with the desired properties, this will complete the proof.

Claim V: There is a set P 0 P with the desired properties ⊆ Recall that we have fixed a section σ, which is locally represented by the function t m f : Mγ FL. Choose an open neighborhoodU of ψL(pL(π˜ (Wγ))) R where ζL is → ⊆ t identically equal to 1. Choose a smaller open neighborhood V of ψL(pL(π˜ (Wγ))), such that V U. Let δ denote the distance between ψ (p (π˜ t (W ))) andU \V , and ⊂ L L γ let ∆ denote the distance between V and Rm \U. It is possible to choose U and V in such a way that both δ and ∆ are positive. We make such a choice. Let P 0 P be ⊆ the set of polynomials p such that p(x) min(δ,∆) whenever x ψ (M ). Since | | < ∈ M γ ψM (Mγ) is open with compact closure, P 0 is an open subset of P. It also contains the zero polynomial since δ,∆ 0. It remains to be shown that each point of Wγ is > r a regular value of the map Φ: K P 0 J L K defined by × → ⊕

Φ(k,p) (j r σ (π (k)),k). = p M

(Recall that we have fixed a section σ.) Suppose that Φ(k,p) belongs to Wγ. Then π (k) M . Let x π (k). The section σ is represented locally by the function M ∈ γ = M p 92 APPENDIX B. ASSORTED PROOFS f : M F on M . The condition p(ψ (x)) δ ensures that p γ → L γ | M | < ψ (f (x)) ψ (p (π˜ t (W ))) ψ (f (x)) V. L p ∈ L L γ =⇒ L ∈ It holds that f (x) p (π˜ t (W )) L , so ψ (f (x)) V . The condition p(ψ (x)) p ∈ L γ ⊆ γ L ∈ | M | < ∆ ensures that ψ (f (x)) V ψ (f (x)) p(ψ (x)) U. L ∈ =⇒ L + M ∈ Hence 1 f (y) ψ− (ψ (f (y)) p(ψ (y))) p = L L + M in a neighborhood of x. By perturbing p slightly we may perturb the partial deriva- tives up to order k of fp arbitrarily. This means that (x,p) is a regular point of the k (locally defined) map (y,q) j fq (y), which in turn means that it is a regular k 7→ point of the map (y,q) j σq (y). This in turn means that (k,p) is a regular point k7→ of the map (`,q) (j σq (πM (`)),`). The latter map is the map Φ, so since (k,p) 7→ 1 was an arbitrary point of Φ− (Wγ) we have now shown that each point of Wγ is a regular value of Φ. This proves the claim, and hence the theorem.

Bibliography

[1] J. K. Beem and S. G. Harris. The generic condition is generic. Gen. Relativity Gravitation, 25(9):939–962, 1993.

[2] J. K. Beem and S. G. Harris. Nongeneric null vectors. Gen. Relativity Gravita- tion, 25(9):963–973, 1993.

[3] J. K. Beem and A. Królak. Cauchy horizon end points and differentiability. J. Math. Phys., 39(11):6001–6010, 1998.

[4] A. Borde. Topology change in classical general relativity. arXiv:gr- qc/9406053v1, 1994.

[5] R. J. Budzy´nski, W. Kondracki, and A. Królak. New properties of Cauchy and event horizons. In Proceedings of the Third World Congress of Nonlinear An- alysts, Part 5 (Catania, 2000), volume 47, pages 2983–2993, 2001.

[6] R. J. Budzy´nski, W. Kondracki, and A. Królak. On the differentiability of com- pact Cauchy horizons. Lett. Math. Phys., 63(1):1–4, 2003.

[7] P.T. Chru´sciel. Elements of causality theory. arXiv:1110.6706v1 [gr-qc], 2011.

[8] P.T. Chru´sciel, E. Delay, G. J. Galloway, and R. Howard. Regularity of horizons and the area theorem. Ann. Henri Poincaré, 2(1):109–178, 2001.

[9] P. T. Chru´sciel, J. H. G. Fu, G. J. Galloway, and R. Howard. On fine differen- tiability properties of horizons and applications to . J. Geom. Phys., 41(1-2):1–12, 2002.

[10] P.T. Chru´scieland G. J. Galloway. Horizons non-differentiable on a dense set. Comm. Math. Phys., 193(2):449–470, 1998.

[11] F. H. Clarke. Optimization and nonsmooth analysis, volume 5 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, second edition, 1990.

[12] H. Federer. Curvature measures. Trans. Amer. Math. Soc., 93:418–491, 1959. 94 BIBLIOGRAPHY

[13] H. Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.

[14] G. J. Galloway. Maximum principles for null hypersurfaces and null splitting theorems. Ann. Henri Poincaré, 1(3):543–567, 2000.

[15] G. J. Galloway. Null geometry and the Einstein equations. In The Einstein equations and the large scale behavior of gravitational fields, pages 379–400. Birkhäuser, Basel, 2004.

[16] R. P.Geroch. Topology in general relativity. J. Mathematical Phys., 8:782–786, 1967.

[17] M. Golubitsky and V. Guillemin. Stable mappings and their singularities. Springer-Verlag, New York-Heidelberg, 1973. Graduate Texts in Mathemat- ics, Vol. 14.

[18] A. Hatcher. Algebraic topology. Cambridge University Press, Cambridge, 2002.

[19] S. W. Hawking and G. F.R. Ellis. The large scale structure of space-time. Cam- bridge University Press, London, 1973. Cambridge Monographs on Mathe- matical Physics, No. 1.

[20] M. W. Hirsch. Differential topology, volume 33 of Graduate Texts in Mathe- matics. Springer-Verlag, New York, 1994. Corrected reprint of the 1976 origi- nal.

[21] M. Kriele. Spacetime, volume 59 of Lecture Notes in Physics. New Series m: Monographs. Springer-Verlag, Berlin, 1999. Foundations of general relativity and differential geometry.

[22] D. N. Kupeli. On null submanifolds in spacetimes. Geom. Dedicata, 23(1):33– 51, 1987.

[23] D. E. Lerner. The space of Lorentz metrics. Comm. Math. Phys., 32:19–38, 1973.

[24] J. W. Milnor and J. D. Stasheff. Characteristic classes. Princeton University Press, Princeton, N. J.; University of Tokyo Press, Tokyo, 1974. Annals of Math- ematics Studies, No. 76.

[25] F. Morgan. Geometric measure theory. Academic Press, Inc., San Diego, CA, third edition, 2000. A beginner’s guide. BIBLIOGRAPHY 95

[26] M. Nardmann. Nonexistence of spacelike foliations and the dominant energy condition in Lorentzian geometry. arXiv:math/0702311v1 [math.DG], 2007.

[27] W.-T. Ni. Geodesic triangles and expansion of metrics in normal coordinates. Chinese Journal of Physics, 16:223–227, 1978.

[28] B. O’Neill. Semi-Riemannian geometry, volume 103 of Pure and Applied Mathematics. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, 1983. With applications to relativity.

[29] B. L. Reinhart. Cobordism and the Euler number. Topology, 2:173–177, 1963.

[30] A. D. Rendall. The continuous determination of spacetime geometry by the Riemann curvature tensor. Classical Quantum Gravity, 5(5):695–705, 1988.

[31] D. J. Saunders. The geometry of jet bundles, volume 142 of London Mathe- matical Society Lecture Note Series. Cambridge University Press, Cambridge, 1989.

[32] F. J. Tipler. Singularities and causality violation. Ann. Physics, 108(1):1–36, 1977.

[33] P.Yodzis. Lorentz cobordism. Comm. Math. Phys., 26:39–52, 1972.

[34] P.Yodzis. Lorentz cobordism. II. General Relativity and Gravitation, 4(4):299– 307, 1973.

TRITA-MAT-E 2014:29 ISRN-KTH/MAT/E—14/29-SE

www.kth.se