Sie sind auf Seite 1von 159

Wormholes in 2+1 dimensions

Balt van Rees Supervisor: dr. K. Skenderis

June 2006

Faculteit Technische Natuurwetenschappen Kavli Institute of Nanoscience Instituut voor Theoretische Fysica

Abstract This Masters thesis investigates wormholes within the framework of general relativity in 2+1 dimensions. After introductory chapters devoted to Riemann surfaces, Teichmller spaces and three-dimensional gravity, we discusss how u these wormholes arise as quotients of a part of AdS3 . We show that all the Riemann surfaces with ideal boundaries have a static wormhole counterpart. The ideal boundaries become asymptotic regions with horizons and the nontrivial topology of the Riemann surface behind the horizon is unobservable from the outside. We give physical coordinate systems in the special case of a torus glued to a single outer region. We also discuss a new procedure for analytically continuing the metric. With this procedure, the corresponding Euclidean spaces are handlebodies. We nally treat some aspects of holography for the handlebodies and nd the vev of the stress-energy tensor of the dual eld theory.

Contents
Introduction Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Riemann surfaces 1.1 Introduction . . . . . . . . . . . . . . . . 1.2 Conformal structures . . . . . . . . . . . 1.3 Uniformization . . . . . . . . . . . . . . 1.3.1 Coverings . . . . . . . . . . . . . 1.3.2 The uniformization theorem . . . 1.3.3 Properties of universal coverings 1.4 The torus . . . . . . . . . . . . . . . . . 1.5 Uniformization in H . . . . . . . . . . . 2 Moduli and Teichm ller spaces u 2.1 Introduction . . . . . . . . . . . 2.2 The exceptional types . . . . . 2.3 Fricke coordinates . . . . . . . 2.4 Quasiconformal mappings . . . 2.5 Uniformization and Teichmller u 2.6 The Bers embedding . . . . . . . . . . . . . . . . . . . . . . theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 7 8 9 9 11 15 15 17 19 21 22 31 31 32 34 35 36 38 41 41 42 44 46 48 56 59 59 61 62 63 64 64 68

3 Three-dimensional spacetimes 3.1 Gravity in three dimensions . . . . . . . . 3.2 Identications and quotient spacetimes . . 3.3 Three-dimensional Anti de Sitter . . . . . 3.4 Coordinate systems on AdS3 . . . . . . . 3.4.1 Coordinate ranges . . . . . . . . . 3.4.2 Combining the coordinate systems 4 Three-dimensional wormholes 4.1 Introduction . . . . . . . . . . . . . 4.2 Extension of Aut(H) to AdS3 . . . 4.2.1 Fixed points and lightcones 4.3 Quotients . . . . . . . . . . . . . . 4.4 The BTZ black hole . . . . . . . . 4.4.1 Unit mass BTZ black hole . 4.4.2 Mass . . . . . . . . . . . . . 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.5

4.6

4.4.3 Angular momentum . . . . Outer regions . . . . . . . . . . . . 4.5.1 Riemann surfaces revisited 4.5.2 Extension to a spacetime . Non-fuchsian isometries . . . . . .

. . . . .

. . . . .

. . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69 71 71 75 77 79 79 80 81 82 82 84 89 90 90 91 91 94 96 96 98 98 99 100 102 102 104 106 109 111 112 112 114 116 119 120 121 122 124 125 127 127 128 129 129 131 135 135

5 Toroidal wormhole 5.1 Denition of the spacetime . . . . . . . 5.1.1 Initial data . . . . . . . . . . . . 5.1.2 Spacetime . . . . . . . . . . . . . 5.2 Patching the surface . . . . . . . . . . . 5.2.1 Denition . . . . . . . . . . . . . 5.2.2 The patches in H . . . . . . . . . 5.2.3 A nal isometry . . . . . . . . . 5.2.4 Summary . . . . . . . . . . . . . 5.3 Lifting the issues . . . . . . . . . . . . . 5.3.1 Objective . . . . . . . . . . . . . 5.3.2 Procedure . . . . . . . . . . . . . 5.3.3 Conclusion and a new condition 5.3.4 The denition of Si . . . . . . . 5.4 Patching the spacetime . . . . . . . . . . 5.5 Toy coordinates . . . . . . . . . . . . . . 5.5.1 Inner annulus . . . . . . . . . . . 5.5.2 Outer annulus . . . . . . . . . . 5.5.3 Transition functions . . . . . . . 5.6 New metric, old metric . . . . . . . . . . 5.6.1 Inner annulus . . . . . . . . . . . 5.6.2 Outer annulus . . . . . . . . . . 5.6.3 Transition functions . . . . . . . 5.7 Discussion . . . . . . . . . . . . . . . . .

6 Hyperbolic three-manifolds 6.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1.1 Hyperbolic three-space . . . . . . . . . . . . . . . 6.1.2 Schottky uniformization . . . . . . . . . . . . . . 6.2 The Euclidean BTZ black hole . . . . . . . . . . . . . . 6.3 A dierent analytic continuation . . . . . . . . . . . . . 6.3.1 Formalism . . . . . . . . . . . . . . . . . . . . . . 6.3.2 General isometries of H 2+1 . . . . . . . . . . . . 6.3.3 Isometries and their Killing vectors . . . . . . . . 6.3.4 An explicit map to Isom0 (H 3 ) . . . . . . . . . . 6.3.5 A dierent viewpoint . . . . . . . . . . . . . . . . 6.4 Handlebodies . . . . . . . . . . . . . . . . . . . . . . . . 6.4.1 Analytic continuation of BTZ . . . . . . . . . . . 6.4.2 Analytic continuation of static wormholes . . . . 6.5 Issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.1 Conjugation freedom and coordinate dependence 6.5.2 Analytic continuation of subgroups . . . . . . . . 6.5.3 Summary . . . . . . . . . . . . . . . . . . . . . . 6.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . 4

7 Holography 7.1 The asymptotic form of the metric . . 7.2 Complex coordinates . . . . . . . . . . 7.3 Further properties of the expansion . . 7.4 Fixing the global data . . . . . . . . . 7.5 Constant curvature metric results . . . 7.5.1 The almost trivial sphere . . . 7.5.2 Contracting loops on a torus . 7.5.3 Boundaries with negative Euler 7.6 Conclusion . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . number . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

137 138 139 141 143 145 146 146 148 149

Conclusion 151 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

Introduction
This Masters thesis considers general relativity in three spacetime dimensions. In particular, we analyze the geometry of certain three-dimensional spacetimes that we call 2+1-dimensional wormholes. These spacetimes have the topology of S R and carry a signature ( + +) metric satisfying the vacuum Einstein equations with < 0. Here S is a two-dimensional non-compact surface and R the timelike direction. The non-compactness of S ensures that the spacetimes have asymptotic regions and we will show how each asymptotic region has horizons associated to it, so the spacetimes indeed classify as wormholes.

Motivation
Of course, since our universe has four (macroscopic) dimensions, the applicability of three-dimensional gravity in physics is not completely straightforward. So let us start by giving three reasons for the relevance of our work. First, three-dimensional black holes are used as a playground for quantum gravity that might help us to understand the full, higher-dimensional problem. This is because the three-dimensional black holes exhibit many features that the four-dimensional black holes do, but the calculations are a lot less involved. In particular, after their discovery in 1992 [8], it was shown that three-dimensional black holes can be equipped with angular momentum and charge, that they emit thermal radiation and have an entropy equal to one quarter of the horizon area (the horizon length in this case) in Planck units [16]. On the other hand, the three-dimensional case is easier and more tractable, mainly because there are no local degrees of freedom (see chapter 3). This considerably simplies the calculations of general relativity. The wormhole spacetimes we study here are a relatively new and fairly uninvestigated class of solutions. Their quantum properties might be exotic but not completely uninteresting or impossibly hard to nd. Therefore, although the results of this thesis are mainly classical in nature, we like to see it rst and foremost as setting the stage for a hopefully very interesting quantum analysis. A second reason is the recently conjectured duality between gauge theories and (quantum) gravity in one dimension higher, known as the AdS/CFT correspondence [21, 31, 46]. The main idea of this correspondence is that computations in a d-dimensional conformal eld theory can be performed within d + 1-dimensional (quantum) gravity with < 0 and vice versa. A dictionary then translates the results between the theories. The d-dimensional space on which the conformal eld theory is dened is the conformal boundary of the d + 1-dimensional manifold. This correspondence is also not completely un7

derstood (let alone proved) and again restricting ourselves to three dimensions can be an interesting toy model to gain more insight in the AdS/CFT correspondence. In our case, of course, we should look for clues of a 2-dimensional conformal eld theory living on the boundary of our spacetimes. Last, we mention a result from string theory. It can be shown [24, 36, 31] that some of the big (higher-dimensional) brothers and sisters of the threedimensional black hole are, in a near-horizon limit, U-dual to a three-dimensional wormhole. In particular, microscopic derivations of the entropy also reduce to the three-dimensional case so from this perspective it is extremely worthwhile to investigate the three-dimensional black holes and wormholes in more detail.

Outline
Our goal in this report is to present a detailed account of the geometry of the 2+1-dimensional wormhole spacetimes we introduced above. We do this as outlined below. The rst two chapters serve as the mathematical background for the remainder of the report. Chapter 1 covers the theory of Riemann surfaces and chapter 2 the theory of moduli spaces, which are spaces of all Riemann surfaces of a certain type. We will see in later chapters how Riemann surfaces and their moduli spaces are related to our study of three-dimensional wormholes. Chapter 3 is also an introductory chapter where we consider the special features of gravity in three dimensions. We will see how solutions to the Einstein equations can be constructed by cutting pieces from a manifold and gluing the edges together. The particular manifold we will cut to pieces is called threedimensional Anti de Sitter; we discuss some properties of it as well. Chapter 4 discusses in detail the actual procedure to obtain a wormhole spacetime. We then discuss the canonical example, which is the three-dimensional equivalent of the extended Schwarzschild black hole. We also show that all the wormholes are indistinguishable from each other on the outside of their horizons. Afterwards, in chapter 5, we pick the particular example where S is a punctured torus. We introduce new, physical coordinates to cover the spacetime and work out the metric and the transition functions. Chapter 6 will deal with an analytic continuation procedure. We hope to nd a procedure to obtain signature (+ + +) manifolds as Euclidean equivalents of the wormhole spacetimes. We introduce and extend the procedure proposed in [27], but also show some ambiguities that need to be resolved. Finally, chapter 7 discusses some aspects of the AdS/CFT correspondence. In particular, in the case of the Euclidean manifolds of chapter 6, we show how geometric results can give some insight in the dual conformal eld theory.

Chapter 1

Riemann surfaces
In this chapter, we will consider some topics from the theory of Riemann surfaces. The introduction sketched below is far from complete but gives the necessary background for the remaining chapters. We will then also see how many of the features of the three-dimensional wormhole spacetimes can be described using the theory of Riemann surfaces. Riemann surfaces are complexied versions of ordinary surfaces. The classication of all possible Riemann surfaces is basically the domain of Teichmller u theory, which is the subject of the next chapter. This chapter is both an introduction to Riemann surfaces and a preparation for the upcoming treatment of Teichmller spaces. u Many books have appeared on the theory of Riemann surfaces and Teichmller spaces. Most of the material presented in this and the next chapu ter can be found in the introductory book by Imayoshi and Taniguchi [25]. A more thorough introduction to both Riemann surfaces and uniformization has been written by Farkas and Kra [19]. For the theory of quasiconformal mappings, the reader is referred to Nag [34]. The books by Abiko [1] and Lehto [30] are also useful references for a detailed treatment of covering surfaces and uniformization.

1.1

Introduction

We start with the introduction of several concepts related to the theory of complex manifolds. To begin with, a Riemann surface is a complex manifold of complex dimension one. More precisely, this means: Denition 1.1.1. S is a Riemann surface if the following is true: S is a topological space; S is covered with a family of open sets {Ui }; For every Ui , there exists a homeomorphism i : Ui i (Ui ) with i (Ui ) an open subset of C; For every set Ui Uj = , the transition functions ji = j 1 from i i (Ui Uj ) to j (Ui Uj ) are holomorphic. 9

We call a set {(Ui , i )} with holomorphic transition functions a holomorphic atlas for S. Every such an atlas is contained in a unique maximal holomorphic atlas which is the complex structure of S. Henceforth, when the chosen chart i is clear, we do not need to and will not distinguish between a point p on S and the corresponding point z = i (p) in C. Remark. It can be shown [40] that every Riemann surface is orientable. Having dened complex manifolds, a logical next step would be to introduce the corresponding complex (co)tangent bundles. Below, we do this by the convenient method of generalizing the notion of these bundles on real manifolds. We rst map a complex manifold of complex dimension one to a real manifold of dimension two, simply by setting z = x + iy in every chart. The transition functions of the real manifold, which we call X below, will then satisfy the Cauchy-Riemann equations. Consider now such an ordinary orientable real two-dimensional smooth manifold X with charts to R2 , tangent bundle T X and cotangent bundle T X. For every p X, we complexify the vector space Tp X and write the resulting space as Tp S C . An element of the complexied vector space can be written as U + iV with U and V elements of Tp X. Adding vectors and complex scalar multiplication are straightforward, remembering that i2 = 1. The complex dimension of Tp S C equals the real dimension of Tp X and the vectors 1 1 = i z 2 x 2 y 1 1 = +i z 2 x 2 y (1.1)

form a basis for the complex vector space Tp S C . We say that an element of Tp S C is holomorphic resp. anti-holomorphic if it can be written as c z resp c z , c C. A vector c z + c z is real. Similarly, the dual vector space can be complexied. The dual vectors of (1.1) are, respectively: dz = dx + idy d = dx idy z

and form a basis of Tp S C . It follows that a basis for the complexication of the C C space Tp S Tp S is: dz d = 2idx dy. z

The basis dz, d transforms under a holomorphic change of charts w = w(z) as z w dz z w w d = z d z dw = z z dw =

because of the Cauchy-Riemann equations w = 0. z The concepts introduced above can be generalized to tensors. Denition 1.1.2. Given two arbitrary integers p and q, a tensor of type (p, q) on a Riemann surface S is the assignment of a function (z) (z) to every chart z on S such that (z) (z)dz p dq z 10

is invariant under coordinate changes (i.e. a change of charts). This means that for such a change w = w(z). (z) (z) = (w) (w(z)) dw dz
p

dw dz

We will drop the subscripts from now on and just write (z). Clearly, a function on S is a (0,0) tensor; a eld of holomorphic or anti-holomorphic vectors can be directly obtained from a (-1,0) tensor or a (0,-1) tensor, respectively. A (1,0) tensor and a (0,1) tensor together dene a one-form (z)dz + (z)d. z A two-form is obtained from a (1,1) tensor and is dened as (z)dz d. We z say that a (p, 0) tensor with a holomorphic function (z) is holomorphic. Note that the holomorphicity of (z) is well-dened because the transition functions are holomorphic. We conclude this section by having a brief look at derivatives. Let f = f (z, z) be a function on S. We write fz = f and fz = f and introduce the and z z operators: f = fz dz For a one-form = (z)dz + (z)d: z = z (z)dz d z = z (z)dz d z f = fz d. z

and for a two-form, we dene = = 0. We also dene: d=+ = 2i

where the factor 2i comes from the ordinary denition (in real coordinates) of the Laplacian. Clearly, 2 = + = 2 = d2 = 0.

1.2

Conformal structures

We can extend the relation between a Riemann surface S and a real orientable surface X, introduced in the previous section. We will see that the complex structure of S is equivalent to a conformal structure on X. Denition 1.2.1. A Riemannian metric g on X is a positive denite section of S 2 T X, i.e. a real positive denite symmetric tensor eld in T X T X. This means that at every p X, the component functions gab (p) of g: g = gxx dx dx + gxy dx dy + gyx dy dx + gyy dy dy form a positive denite symmetric matrix. So for a Riemannian metric on X and a vector V Tp X, we have g(V, V ) 0 and g(V, V ) = 0 V = 0.

Remark. For notational simplicity, when the context is clear, we write dxdy 1 when we mean 2 (dx dy + dy dx) and for example (dx)2 instead of dx dx. We also write ds2 = g for a metric. 11

Suppose we have a Riemannian metric g on a surface X. Again setting z = x + iy, we obtain g = 2 |dz + d|2 z (1.2) with (x, y) = 1 2 (gxx + gyy + 2 gyy gxx gxy ) 4 1 (x, y) = 1 (gxx gyy + 2igxy ). 4

This is the most general form of a metric in complex coordinates. It is not dicult to show that |(x, y)| < 1 everywhere. Note that a diagonal Riemannian metric is given by dzd = (dx)2 + (dy)2 , so (x, y) = 0. z In any dierent chart w, where w = w(z) is holomorphic, the metric is given by g = 2 (w, w)| dz dz dw|2 = dw + (w, w) dw dw dz 2 dz 2 (w, w) |dw + dw dw

dz dw

dw|2 . (1.3)

From the above transformation property we can read o immediately that 2 (z, z) should be a (1,1) tensor and (z, z ) should be a (-1,1) tensor. Note that dz 2 2 (z, z) = 2 (w, w) dw so that there is no problem with the demand that 2 > 0. Denition 1.2.2. A conformal structure on X is an equivalence class of Riemannian metrics, where the equivalence relation is: g e2 g with = (p) any smooth real-valued function on X. A conformal structure can be used to determine angles between vectors but not lengths. The remainder of this section is dedicated to proving the following theorem: Theorem 1.2.3. On a two-dimensional manifold, a complex structure and a conformal structure are equivalent. We will do this rather extensively and in a mathematically rigorous way, which gives us the opportunity to introduce many important tools for the next chapter. Lemma 1.2.4. A complex structure on a surface gives a unique conformal structure. Proof. We pick a chart z on a Riemann surface S and dene a metric g by g = dzd. z 12

Then in any other chart w = w(z), we have, as follows from equation (1.3): g= dz dz dz 2 dwdw = dwdw. dw dw dw

which is in the same conformal class as the metric g = dwdw. So no matter what chart z i we started with, all metrics dened as dz i di +di dz i for some z z i are conformally equivalent. So if we do not prefer a certain chart over another (as we should not), this means the following: starting from only a complex structure, we have constructed a conformal equivalence class of metrics, thus a conformal structure. The converse, namely that a conformal structure implies a complex structure, is a little less obvious. We need theorem 1.2.6 and lemma 1.2.7 to prove lemma 1.2.8. Suppose we have a surface X with a conformal structure which is written in an arbitrary chart (x, y) as the equivalence class of |dz + (z, z )d| 2 with z z = x + iy. Note that z is an arbitrary complex coordinate and depends on the particular chart we chose. It is thus not in any way xed by the conformal structure on X, hence not the complex coordinate we are looking for. However, we still extend the coordinate z to a complex coordinate on X entirely with holomorphic transition functions (that this can be done, follows from the reasoning below with = 0). Starting now from the arbitrary coordinate z, we shall nd charts in which the conformal structure is diagonal and show that a unique complex structure is dened by complexifying in these charts. Extending z to X entirely implies that (z, z ) becomes a (-1,1) tensor in these charts, as we saw before. Also, this tensor completely determines the conformal structure since from equation (1.3), under a holomorphic change of charts u = u(z), the following holds: |dz + (z) (z, z )d|2 |du + (u) (u, u)d|2 . z u To diagonalize this conformal structure, we search a dieomorphism of X X of the form w = w(z, z ) such that the metric dwdw is in the same conformal class as the original metric, that is dwdw |dz +d|2 . This means that w(z, z ) z should satisfy the following dierential equation: wz (z, z ) = (z, z )wz (z, z ) (1.4)

which is the all-important Beltrami equation. Below, we will show that w exists and denes a unique complex structure associated with any conformal structure dened by a (-1,1) tensor . Denition 1.2.5. Consider a general orientation-preserving homeomorphism f between open domains U1 and U2 in C. We dene the complex dilatation f (z, z ) of such a homeomorphism to be f = f f z z
1

Then f is a quasiconformal homeomorphism if |f | < 1 a.e. (almost everywhere, which means everywhere with the exception of a set of zero measure) on U1 and a conformal homeomorphism if f = 0 a.e. 13

For a good introduction to quasiconformal mappings, see for example [34]. It is Weyls lemma that states that if gz = 0 a.e., then g is holomorphic and thus a conformal homeomorphism is biholomorphic. Theorem 1.2.6. For every function on a domain U C, || < 1 a.e., there exists a quasiconformal homeomorphism f of U with complex dilatation . This is the existence theorem for solutions to (1.4). It will not be proved here, but a proof can be found in [34]. Its meaning is that in any chart z i there exists at least one solution w i that diagonalizes the conformal structure, as follows directly from denition 1.2.5 and the Beltrami equation. The new charts are now exactly these solutions w i . Having established existence of a solution in every patch, we now turn to the demand of holomorphic transition functions: given two solutions w i and wj on overlapping patches, the transition function wi (wj )1 should be holomorphic. Lemma 1.2.7. Under the composition of two quasiconformal homeomorphisms f (z) and g(w) with respective complex dilatations f and g , the complex dilatation of g f is equal to: gf = a.e. on the domain of f . Proof. The result follows directly from the usual chain rules, which are in complex notation written as: (g f )z = (gw f )fz + (gw f )fz fz f + fz (g f ) fz + fz (g f )f (1.5)

(g f )z = (gw f )fz + (gw f )fz . We can rearrange (1.5) in order to solve for g f : g f = fz gf f fz 1 gf f

so if and only if two new charts w i and wj that are quasiconformal functions of z have the same complex dilatation, that is if and only if wi = wj = a.e., the complex dilatation of w i (wj )1 is zero a.e. This means that for two dierent solutions w i and wj of the same Beltrami equation (1.4) with complex dilatation , the function w i (wj )1 is conformal hence biholomorphic. We have proved our earlier statement: the transition functions between two new charts wi and wj on overlapping patches are indeed holomorphic. We nally need to show that the charts w i dened on all the patches covering X can be extended to a maximal complex structure on X. This means that we should be able to compose any solution (i.e. any new chart) w to the Beltrami 14

now setting g = hf 1 , it follows that for two quasiconformal homeomorphisms f and h: fz (h f ) hf 1 f = fz (1 h f )

equation (1.4) with an arbitrary biholomorphic mapping g and g w should still be a solution. This property also follows directly from lemma 1.2.7: if g is biholomorphic so g = 0, then gf = f . So for any given , any composition of a solution of (1.4) with a biholomorphic mapping is also a solution. We thus proved that the solutions w i extend to a maximal atlas on S. The existence of a solution to the Beltrami equation and its uniqueness up to biholomorphic mappings now imply the following: Lemma 1.2.8. On a two-dimensional manifold, any conformal equivalence class of metrics of the form |dz + d|2 (so any bounded (-1,1) tensor) implies z a unique complex structure with local charts w such that |dz + d|2 dwdw. z Proof. The complex structure is the restriction of the maximal atlas of X to all charts where the conformal structure is diagonal. We have just seen that, starting from a given chart, such a chart can always be obtained so these charts cover the entire surface. Furthermore, two solutions can only dier by composition with a biholomorphic mapping so the transition functions are always holomorphic. The inverse, that composition of a solution with a biholomorphic mapping remains a solution, means that the atlas {w} can be extended to a maximal atlas, which is a complex structure on X. Finally, starting from this complex structure w, it is easy to see that the conformal structure obtained in the proof of lemma 1.2.4 is exactly the conformal structure we started with. This closes the circle: not only does a complex structure imply a unique conformal structure (lemma 1.2.4) and vice versa (lemma 1.2.8), but the two are completely equivalent and theorem 1.2.3 is proved. A natural question is now whether one can classify all complex structures on a surface. It is clear that this has at least some relation to the set of all bounded (-1,1) tensors that can be dened on the surface. In the next chapter, we will investigate this subject in more detail, but rst we discuss some useful properties of Riemann surfaces.

1.3

Uniformization

A very useful tool in the classication of all possible Riemann surfaces is the theory of uniformization. In particular, this means that every Riemann surface can be obtained as a suitable quotient space (explained below) from one of only three simply connected Riemann surfaces with a certain group of biholomorphic self-mappings. Our goal is to give an overview of uniformization in this section and state the most important theorems.

1.3.1

Coverings

As a necessary introduction, we will work our way through coverings of surfaces and in particular universal coverings. Denition 1.3.1. A covering (S , f ) of a Riemann surface S is a Riemann surface S and a continuous surjective map f : S S which is locally biholomorphic, i.e. every point p S has an open neighborhood which is mapped biholomorphically to an open domain in S. 15

If the set f 1 (p) contains exactly n points for every p S, the covering is said to be n-sheeted. Furthermore, if f is everywhere locally biholomorphic but with the exception of a certain distinct number of points, these points are called branch points and the covering is said to be branched. In the right charts where a branch point is mapped to 0 on S and its image to 0 on S, the branched covering looks locally like (z )k = z for some positive integer k. Of course, an unbranched covering looks locally like the identity map in such charts. Denition 1.3.2. A biholomorphic map : S S such that f = f is called a covering transformation of (S , f ). The set of all covering transformations has a group structure and this group is called the covering group of the covering. A covering (S, f ) of S is universal if S is simply connected. We then say The associated covering group is the that S has universal covering surface S. universal covering group. One can actually construct a universal covering surface for every Riemann surface S. This is done by taking a base point p0 on S and looking at all inequivalent paths from this base point, where two paths are equivalent if they are homotopic to each other and have the same endpoint on S. The equivalence class of a path C from p0 to p is denoted [C, p]. It can be shown [25] that the set of all these classes indeed becomes a universal covering (S, f) with a locally biholomorphic covering map f : S S dened as f : [C, p] p. Theorem 1.3.3. For any Riemann surface, the universal covering surface as constructed above always exists and is unique up to biholomorphic mappings. For a proof, see [3]. With the existence and uniqueness of a universal covering surface in mind, we now turn to the universal covering group , to show ultimately that S is biholomorphic to S/. First of all, one has the following property of : Lemma 1.3.4. The universal covering group acts transitively on S, meaning p p with f (1 ) = f (2 ), there exists an element such that for any p1 , p2 S that (1 ) = p2 . p In fact, suppose that f (1 ) = f (2 ) = p and p1 = p2 . Then there exist p p two paths C1 and C2 on S from p0 to p that are not homotopic to each other. 1 The path C1 C2 is a nontrivial closed loop with basepoint p0 . A covering transformation : S S is now dened by adding this path to all points in S, 1 = f so is indeed a covering i.e. ([C, p]) = [C C1 C2 , p]. We see that f transformation that satises (1 ) = p2 . Again, it can be shown that this map p is a biholomorphism S S. For a full proof of lemma 1.3.4, see [30]. For the next theorem, we need the important notion of a properly discontinuous action, which we dene for a general manifold below. Denition 1.3.5. A group G of homeomorphic self-mappings of a manifold M is properly discontinuous if for any compact subset U M , there are at most nitely many elements g G satisfying g(U ) U = . We can take a quotient of M with G, denoted as M/G, by identifying all points that are equivalent under G, i.e. we identify all points pi , pj M if there exists a g G with g(pi ) = pj . If G is properly discontinuous and every 16

nontrivial element of G has no xed points in M , then the quotient space M/G with the quotient topology is a Hausdor manifold, see [23] or [30]. Example 1.3.6. As an example, take M = R2 and G the cyclic group generated by (x, y) (x + 2, y). In the quotient space M/G, we identify (x, y) and (x + 2, y), which we also write as (x, y) (x + 2, y). The quotient manifold M/G is homeomorphic to the (innite) cylinder. Returning now to the theory of Riemann surfaces, we have the following: Theorem 1.3.7. The universal covering group of a Riemann surface S with universal covering surface S satises the following properties: 1. 1 (S), with 1 (S) the fundamental group of S; = 2. Each element of \{Id} has no xed points in S; 3. acts properly discontinuously on S; 4. is unique up to conjugation with a biholomorphic automorphism of S. For the proof of these points, see [25]. The rst point follows from the construction of an element from a closed loop on S as sketched above. The second statement stems from the fact that f is locally biholomorphic, so injective in the neighborhood of possible xed points of an element i . It then follows that i = Id. The third point is a little harder to prove, see [25] or [30]. The last property follows from the lifting property of mappings, which are explained in section 1.3.3 below. Intuitively, suppose we have two covering groups and such that S/ and S/ are biholomorphic to S by theorem 1.3.8 below. Then there exists a biholomorphic automorphism of S/ to S/ , which lifts to a biholomorphic automorphism of S. Theorem 1.3.8. The quotient space S/ is biholomorphic to S. Proof. Properties (2) and (3) of theorem 1.3.7 ensure that the quotient space S/ is a Hausdor manifold because acts properly discontinuously. Denote p by [] an element of S/. The map g : [] f () is well-dened (since for all p p , f = f ), surjective because f is and injective because is transitive. Because acts properly discontinuously and xed-point freely, every point p in S has a neighborhood in which no two points are equivalent modulo . This means that : p [] is locally homeomorphic so a complex structure on S/ p can then be inherited from the complex structure on S. If S/ is equipped with this complex structure, then will be locally biholomorphic and since f is also 1 locally biholomorphic, then so is g, which is locally dened as g = f . So g is everywhere locally biholomorphic and bijective, hence a biholomorphism.

1.3.2

The uniformization theorem

In the previous section, we found out that every Riemann surface can be described as a quotient of its unique (theorem 1.3.3) associated universal covering surface. But the following theorem limits the number of possible universal covering surfaces to exactly three. 17

Theorem 1.3.9 (Uniformization theorem). All simply connected Riemann surfaces are biholomorphic to either the upper half plane H = {z C : Im z > 0}, the complex plane C or the extended complex plane P 1 = C {}. Here, P1 is a one-point compactication of C, homeomorphic to S 2 , see [20] for more information about e.g. its topology and complex structure. This famous theorem was proved by Klein, Poincar and Koebe. A proof can be e obtained using the existence of harmonic functions that can be used to generate biholomorphic mappings between simply connected Riemann surfaces [19]. In the remainder of this section, we will discuss some general and important implications of theorem 1.3.9. As one can deduce from the results in the previous section, a direct consequence is that any Riemann surface can be obtained as a quotient of one of these three surfaces modulo a group of biholomorphic self-mappings (recall the denition of a covering group 1.3.2). Lemma 1.3.10. The biholomorphic self-mappings or biholomorphic automorphisms (z) of the simply connected Riemann surfaces are of the form: For C: (z) = az + b For P1 : (z) = For H: (z) = az + b cz + d with a c b d P SL(2, R) az + b cz + d with a c b d P SL(2, C) (1.6) with a C = C\{0}, b C

This can be seen from the requirement of holomorphicity (so there exists a Laurent expansion) and invertibility of these mappings. The mappings (1.6) are the Mbius transformations. The above biholomorphic automorphisms form a o group for every one of the three simply connected Riemann surfaces, which are denoted by Aut(C), Aut(P1 ) and Aut(H), respectively. Finally, it is a simple exercise to show that Aut(H) P SL(2, R) and Aut(P1 ) P SL(2, C). = = Theorem 1.3.11. Every Riemann surface can be obtained as a suitable quotient of only one of the three simply connected Riemann surfaces C, P1 or H by a subgroup of its group of biholomorphic automorphisms Aut dened in the lemma above, with the subgroup acting properly discontinuously and xed point free. In particular: All Riemann surfaces with P1 as its universal covering are biholomorphic to P1 itself; The only Riemann surfaces with C as their universal covering are the Riemann surfaces that are biholomorphic to C, C or a torus; The universal covering of all other Riemann surfaces is H. 18

Proof. The statement is a direct consequence of the above theorems 1.3.3, 1.3.7, 1.3.8, 1.3.9 and lemma 1.3.10. The rst property of the list stems from the fact that every Mbius transformations has xed points on P1 , which is not admitted o for a covering group. So the covering group must be trivial if the universal covering is P1 . Second, let us look at a nontrivial uniformizing a Riemann surface in C. For any (z) = az + b not to have xed points, we require a = 1. We can then show [2] that is generated by either one or two elements yielding a cylinder (biholomorphically equivalent to the punctured plane C ) or a torus. The third item follows then from the uniqueness of a universal covering. Remark. It is not hard to show [25] that any torus cannot have a universal covering other than C. This follows, for example, from the fact that the fundamental group of the torus is free abelian but not cyclic and every abelian subgroup of P SL(2, R) such that H/ gives a Riemann surface is necessarily cyclic. The Riemann surfaces that do not admit H as their universal covering surface are said to be of exceptional type. By the previous theorem, the exceptional types are all Riemann surfaces biholomorphic to P1 , C, C and all tori.

1.3.3

Properties of universal coverings

In view of the results of the previous sections, notably theorem 1.3.11, it makes sense to discuss some general properties of universal coverings. Fundamental domains Suppose we have a Riemann surface S with universal covering D and covering group . Denition 1.3.12. A fundamental domain is a domain in D if it contains at most one point of every orbit of and its closure in D meets every orbit of . To x ideas, the fundamental domain with boundary points in D identied according to some specic elements is (biholomorphic to) S itself. A fundamental domain can be used to tile D. Many examples of fundamental domains will be given below. Lifts of mappings Suppose we also have a dierent Riemann surface S , uniformized as D / . Denote the covering maps or projections : D D/ and : D D / . Theorem 1.3.13. Given any continuous mapping f : S S , there exists a mapping f : D D such that f = f, which is unique if we choose a such that f () = q . Furthermore, if p p D, q D with ( ) = f (()) and f q p . f is dierentiable or holomorphic, then so is f f >D f >S

D S

19

Proof. If q = f (p), choose a p = [Cp , p] D and q = [Cq , q ] D so () = p p and ( ) = q . Note that since f is continuous, it maps curves to curves. q We note the image of a curve C under f by f (C). Now we dene f([C, r]) = 1 [Cq f (Cp ) f (C), f (r)] for all [C, r] D. It is clear that f = f (r) = f and f ([Cp , p]) = [Cq , q ] and that f is unique. Furthermore, if f is dierentiable is, because and are locally biholomorphic. or holomorphic f We call f a lift of f . The lift of a map induces a homomorphism between the covering groups and , as we will now proceed to show. Consider any invertible continuous map f : S S and a lift f : D D . From the above it is invertible as well. Since f = f and f is invertible, one has is clear f = f f 1 . An element of should satisfy = and substitution yields (f 1 f) = which means: = f f 1 for some . We say that is the image of under f. Transitivity of the covering groups now implies that the homomorphism is actually an isomorphism which we write = f f 1 and one has S = D / = f(D)/f f 1 . In the next chapter, this result will turn out to be crucial when we dene f to be a quasiconformal mapping. Lifts of tensors We saw that mappings could be lifted, but tensors can as well. They are called automorphic forms on D. Consider D and a Fuchsian group uniformizing a Riemann surface S. Denition 1.3.14. A (p, q) tensor g on D is an automorphic (p,q)-form for if q (g )(z )p (z ) = g on D and for all elements and for some xed integers p and q. Consider a tensor (z) of type (p, q) on S. Its lift (w) to a local coordinate w on D is given locally by the inverse of the projection , i.e. if is locally given by a biholomorphic function : w z then: (w) = (z(w)) dz dw
p

dz dw

is a denition of a (p, q) tensor (w) on D if we require this to hold for any chart w = w(w) on D. In particular, if w = (w) for and since = , meaning z((w)) = z(w), we nd (z(w)) = (z(w)) leading to ((w))( )p ( ) = (w) and we conclude that the lift of a (p,q) dierential on S is an automorphic (p,q) form for on D. That the converse also holds is not very dicult to show. 20
q

Metrics To conclude this section, we will dene a Riemannian metric on every Riemann surface. To do so, we introduce the metrics i.e. positive denite (1,1) tensors: dzd z dzd z (1 + |z|2 )2 dzd z (Im z)2 on on on C P1 H (1.7) (1.8) (1.9)

It is easy to verify that the metrics on H and C are automorphic (1,1) forms for all elements of Aut(H) and Aut(C), respectively. This means that the above metrics can be projected on the quotient Riemann surfaces as well. The uniqueness of the universal covering group, property (4) of theorem 1.3.7, makes this projection well-dened so that it makes sense to talk about for example the hyperbolic metric (1.9) on a non-exceptional Riemann surface. Also, lengths and angles dened in these metrics are complex analytic invariants for the Riemann surface. This nishes the general discussion of uniformization. We will now discuss the torus as the most interesting case of uniformization in C, followed by a discussion of the uniformization of almost all other Riemann surfaces in the upper half plane H.

1.4

The torus

The torus is somewhat peculiar, since it is not uniformized in H but in C. Any torus can be obtained by taking the covering group to be generated by z z + b0 and z z + b1 with b0 , b1 C and (arg b0 ) = (arg b1 ) mod . Then, by conjugation with an element of Aut(C), one can always obtain a torus where the rst generator is z z +1 and the second is z z + with Im > 0. A fundamental domain of such a torus and an idea of the tiling is given in the gure below.

PSfrag replacements

Figure 1.1: A fundamental domain (shaded) of a torus in C with modular parameter . The torus is the parallelogram with opposite edges identied. An interesting point is that the transformations z z + 1 and z z + + 1 also generate the covering group. The covering group remains exactly the same, and thus the quotient Riemann surface (the torus) does not change either. 21

0 1

However, we chose dierent generators of the covering group and thus of the fundamental group by the isomorphism from lemma 1.3.7. Another possibility is z z + 1 and z z . Again, the group is the same and all we did was to invert one generator. By using Aut(C) again to write this in the standard form with some such that Im > 0, one can see that = 1/ so this group is equivalent to z z 1 and z z + 1. Now, as will be explained in the next chapter, a set of generators for the fundamental group of a Riemann surface S is called a marking for S and choosing dierent generators is a change of marking. We have just seen that the transformations +1 1 generate the mapping class group whose elements are associated with dierent markings on the same Riemann surface. For the torus, the mapping class group is isomorphic to P SL(2, Z), because every element of the mapping class group can be written as a + b c + d with a b c d P SL(2, Z).

These are all the changes of generators of the fundamental group. We will explain more about the mapping class groups of Riemann surfaces in the next section, when we discuss Teichmller and moduli spaces. u

1.5

Uniformization in H

The upper half plane is the most important simply connected Riemann surface because most Riemann surfaces are uniformized in H. This section discusses a number of properties associated with uniformization in the upper half plane H which will be useful for later use. Types of Riemann surfaces Any compact orientable 2-manifold is classied up to homeomorphism by its genus g. We say that a Riemann surface S is of conformal type (g,n,m) if S is biholomorphic to a compact genus g Riemann surface S from which n points and m closed disks are removed. In other words, S has n punctures and m ideal boundaries. The exceptional types discussed in theorem 1.3.11 in terms of (g, n, m) are then P1 (0,0,0), C (0,1,0), C (0,2,0) and all tori of type (1,0,0). In this section, we consider all Riemann surfaces of non-exceptional type. Closure of H The map (Mbius transformation) o z zi z+i (1.10)

maps H biholomorphically to the unit disk = {z C : |z| < 1}. First of all, this means that we can equally well uniformize any non-exceptional Riemann 22

surface in . Secondly, the closure of in C is the unit circle {z C : |z| = 1}. Similarly, the closure of H in P1 is the real line R plus the point at innity R = R {}, topologically a circle. The hyperbolic metric We stress some important characteristics of the hyperbolic metric (1.9), because it can be projected on every Riemann surface. The following holds: The distance function it induces is complete, so H and every quotient Riemann surface with the hyperbolic metric is Cauchy complete; The boundary R is innitely far away from any interior point; Geodesics are either straight vertical lines (passing through innity) or semi-circles, intersecting R orthogonally on both ends; There is a unique geodesic connecting any two distinct points in H or R; The metric has constant negative Gaussian curvature 1. A more thorough introduction to hyperbolic geometry can be found in many books, e.g. [43]. Fuchsian groups The Mbius transformations, i.e. the elements of Aut(P1 ) are classied: o Denition 1.5.1. An element 0 of Aut(P1 ) which is not the identity is said to be: elliptic if it is conjugate to (z) = ei z for some (0, 2); parabolic if it is conjugate to (z) = z + b for some b C ; hyperbolic if it is conjugate to (z) = z for some R+ \{1}; loxodromic if it is not elliptic or parabolic. The loxodromic transformations include the hyperbolic ones; this is a historical convention. We subsequently only consider elements of Aut(H) P SL(2, R), = which is a subgroup of Aut(P1 ). This means that for an element Aut(H), written as az + b (z) = cz + d with ad bc = 1, we then necessarily have a, b, c, d R. We then dene Tr() = a + d. It can be shown [25] that (z) is elliptic if and only if |Tr()| < 2; parabolic if and only if |Tr()| = 2 and = Id; hyperbolic if and only if |Tr()| > 2. So by looking at the (absolute value of the) trace of the matrix, one can immediately deduce the type of the transformation. A direct calculation gives that for an element of Aut(H) 23

if it is elliptic, it has one xed point in H and one outside H; if it is parabolic, it has one xed point on R; if it is hyperbolic, it has two xed points on R. Note that the xed points of a parabolic and hyperbolic transformation do not lie in H. The possible covering groups acting on H are called Fuchsian groups: Denition 1.5.2. A Fuchsian group is a subgroup of Aut(H), acting properly discontinuously and xed-point free on H. By theorem 1.3.11, the covering group of every Riemann surface of nonexceptional type is then a Fuchsian group. Furthermore, all elements of a Fuchsian group are either parabolic or hyperbolic, because the elliptic transformations have xed points inside H. The xed point set of a Fuchsian group is either a subset of R or R entirely. Examples To give an idea of what the uniformization in H looks like, we consider a Fuchsian group generated by a single parabolic or hyperbolic element. Example 1.5.3 (The punctured disk). The punctured disk is the Riemann surface that is obtained by removing the origin from the unit disk, \{0}. Consider the Fuchsian group generated by z z + 1. (1.11)

This means that all nontrivial elements of are parabolic since every element is of the form z z + n with n Z. A fundamental domain for its action on H is {z H : 0 < Re z < 1} and is sketched in gure 1.2. As for the torus, we glue together the two edges of the fundamental domain; this identication is indicated by the arrow. The covering map w : H \{0} given by w = exp(2iz) maps the sketched fundamental domain (with boundary points identied) biholomorphically to the punctured unit disk. Note that the xed point of (1.11) lies at innity and the map w can be extended to map the point at innity to the origin. The puncture thus corresponds to the xed point of the parabolic transformation. It is common to draw the region around a puncture distorted as on the right gure, because it is innitely far away in the hyperbolic metric. Example 1.5.4 (The annulus). Similarly, we can look at the group generated by z z with > 1. A possible fundamental domain is {z H : 1 < |z| < }. To give an idea of the way H is tiled with fundamental domains, an adjacent fundamental domain {z H : 1 < |z| < 1} is also sketched. The rst fundamental domain is mapped to the annulus {w C : exp(2 2 / ln ) < |w| < 1} by the covering map w = exp 2i 24 ln z . ln

PSfrag replacements

Figure 1.2: The punctured disk is biholomorphic to the shaded fundamental domain with opposite edges identied. The puncture corresponds to the point at innity at the boundary of H. It is clear that the only points xed by a nontrivial element of this group are zero and innity, because every element of the covering group can be written as z m z with m Z. Also, every nontrivial element of the group is hyperbolic.

PSfrag replacements

Figure 1.3: The annulus, uniformized in H with the covering group generated by the transformation z z. Remark. The fundamental domains we have sketched and will sketch all have geodesic boundaries in the hyperbolic metric on H. It is both common and useful to look at such fundamental domains, which are called Dirichlet fundamental domains. Remark. Although the punctured disk and the annulus are topologically equivalent, they have dierent complex structures and no quasiconformal mapping exists between them. The punctured disk is of type (0,1,1) and the annulus of type (0,0,2). The cylinder of type (0,2,0), complex-analytically equivalent to C , is a dierent Riemann surface again. Closed loops and geodesics Remember that since the covering group is isomorphic to the fundamental group, every homotopy class of non-contractible closed curves on a Riemann surface S corresponds to an element of the covering group. In general, as in the above examples, the following holds [1]: A parabolic transformation always corresponds to a closed loop around a puncture; Any other nontrivial loop on S corresponds to a hyperbolic transformation. 25

0 1 1

In particular, simple loops around ideal boundaries (removed disks) also correspond to hyperbolic transformations. Theorem 1.5.5. For a hyperbolic transformation in , there exists a unique closed periodic geodesic C on S in every homotopy class of closed curves [C] 1 (S). Furthermore, if the hyperbolic length of C is L(C), the following holds: Tr2 () = 4 cosh2 (L(C)/2). (1.12)

Proof. We nd a closed geodesic C as follows. As was noted before, there are exactly two xed points on R corresponding to any hyperbolic element . In the hyperbolic metric (1.9) on H there exists a unique geodesic CH connecting these two xed points. By a conjugation of with an element of Aut(H), we can map the two xed points of to 0 and and CH is then the imaginary axis in H. We see that , which can now be written as z z with R+ \{1}, leaves CH invariant. If < > is the cyclic subgroup of generated by , then < > acts properly discontinuously on this geodesic. It thus makes sense to talk about CH / < >, which is (biholomorphic to) a closed periodic geodesic C on the Riemann surface S = H/. This is the geodesic we were looking for. It can be shown [1] that this is the unique periodic geodesic in its homotopy class [C] on S and hence we can associate a unique closed geodesic on S with every hyperbolic element of . The length L(C) can be easily calculated when (z) = z (conjugation with Aut(H) leaves the absolute value of the trace invariant) by integrating ds on the imaginary axis between i and i, for example. Note that Tr()= 1/2 + 1/2 . Further examples We will now discuss several examples, most of which will be used when we discuss either 2+1 dimensional wormholes or Teichmller spaces. u Example 1.5.6 (Trinions). A nice and, as we will see later, also fundamental example is a so-called pair of pants, pair of trousers, pantalon or a trinion. It is a Riemann surface of conformal type (0, 0, 3) and homeomorphic to a domain like {z C : |z| < 4, |z + 2| > 1, |z 2| > 1}. The fundamental group of this surface is generated by two hyperbolic elements 1 and 2 , for example each of them going around a dierent boundary component. The generator of the third simple loop (which should of course also be hyperbolic) is then a composition 1 of these two, for example 1 2 or 1 2 , depending on their orientation. In gure 1.4, we sketch a possible fundamental domain, the completed pair of pants and also the unique periodic geodesics in three homotopy classes. Remark. Intuitively, the two generators are two matrices in P SL(2, R) and hence give us six real degrees of freedom. But we can x three of them by conjugating with an arbitrary element of Aut(H), for example by setting 1 (z) = z and 2 (1) = 1. Hence we have three degrees of freedom left that determine the group completely. It is in fact straightforward [1] to show that, if the boundary components are ordered (so excluding mirror images), the Riemann surface H/ is determined uniquely by the lengths of the three geodesics in the homotopy class of simple closed curves around each of the three boundary 26

Figure 1.4: A fundamental domain for a trinion in H. Opposite line segments that are connected with the dashed lines are identied. The dashed lines are the unique geodesics in their respective homotopy classes and connect the xed points on R of the corresponding hyperbolic element of . components. Furthermore, every triple of real positive numbers corresponds to such a pair of pants. This is of course of great help in dening the set of all pairs of pants, or more formally the moduli space of Riemann surfaces of type (0,0,3). Example 1.5.7 (Tight pants). One (or more) of the loops of a pair of pants may be shrunk to a puncture. In the words of the previous example, we could have chosen 1 to be parabolic, for example. We then get a pair of pants where one of the cus or the waist is tight (the nomenclature is not my own). Such a pants is pictured in gure 1.5.

Figure 1.5: A closed trinion. There is no periodic geodesic in the homotopy class of closed curves around the puncture, since the corresponding element of the covering group is parabolic. Example 1.5.8 (Toroidal wormhole). A slightly more dicult Riemann surface that can be obtained as a quotient of H is a torus with a removed disk. (A punctured torus is also uniformized in H, by the way.) A possible fundamental domain is pictured in gure 1.6. The toroidal wormhole has a fundamental group that is generated by two elements and contains only hyperbolic elements. In the case of the wormhole, the commutator of the two generators corresponds to the loop encircling the removed disk in gure 1.6. Note that both a (non-tight) pair of pants and the toroidal wormhole have a two-generator covering group with all hyperbolic elements. We can distinguish them (only) by the so-called minimal intersection number of two closed loops 27

Figure 1.6: A fundamental domain for a toroidal wormhole in H with three closed geodesics. This structure will be investigated as a particular type of space (spatial slice) in 2+1-dimensional gravity in chapter 5. As we will see, the closed geodesic separating the torus from its ideal boundary component will play the role of a horizon. corresponding to the generators: in the case of the wormhole, the two loops necessarily intersect, whereas for the pair of pants we can always homotopically deform the loops to make them non-intersecting. Example 1.5.9 (Compact genus 2 surface). Our nal example is a compact Riemann surface of genus 2. Note that in gure 1.7, the fundamental domain does not reach the boundary of H, reecting the fact that the surface is compact. Still, in every homotopy class of simple closed curves, there exists a unique periodic geodesic. We indicated ve of such geodesics.

Figure 1.7: A genus 2 surface and a fundamental domain for its uniformization in H.

Schottky double, Nielsen kernel Finally, a bit of bricolage with Riemann surfaces. We will dene and explain the Schottky double, the Nielsen kernel and the Nielsen extension of a Riemann surface. For a noncompact surface of type (g, n, m) with m > 0, we explained before that it has m ideal boundaries corresponding to hyperbolic elements of the covering group. The ideal boundaries themselves can be seen as circles and correspond to the intersection of the closure of a fundamental domain with R, i.e. open domains in R, with endpoints identied. 28

We will now extend the action of a Fuchsian group to all of P1 instead of just on H, but should keep in mind its xed points. The limit set (G) of a subgroup G of Mb is the set of points that are o xed by an element of G\{Id}. As we said before, for a Fuchsian group, () is either R entirely or a (nowhere dense) subset of R. The complement of the limit set in P1 is the region of discontinuity (). In this region, Fuchsian groups act properly discontinuously [34]. Denition 1.5.10. The Schottky double of a Riemann surface, uniformized in H with covering group and m > 0 ideal boundaries, is the surface corresponding to ()/. Note that a mirror image of a surface that is uniformized in H can be obtained as the surface L/ where L = {z : Im z < 0}. So in fact, the Schottky double consists of the surface itself H/ and its mirror image L/, glued together along their common ideal boundaries (() R)/. The Schottky double of a Riemann surface of type (g, n, m) is of type (2g + m 1, 2n, 0). Example 1.5.11. The Schottky double of the annulus is the torus. This is because the two semicircles in the fundamental domain in 1.3 become complete circles when extended to L and R. These circles are identied under and this yields a torus. Similarly, the Schottky double of the pair of pants and of the torus with boundary is a genus two surface. The Schottky double of the genus two surface is not dened, since it has no ideal boundaries. Finally, the Schottky double of H itself is P1 . Note that the hyperbolic metric (extended to L) on a Schottky double becomes singular on the ideal boundary circles. We can simply nd a new metric on the surface by uniformizing the Schottky double in C or H (or P1 for the trivial Fuchsian group). Denition 1.5.12. The Nielsen kernel of a Riemann surface S with m > 0 ideal boundaries is obtained by removing the annular regions around every ideal boundary up to and including the unique periodic geodesic surrounding the ideal boundary. Example 1.5.13. The Nielsen kernel of the pair of pants in example 1.5.6 is the fundamental domain with the cus and the waist cut o until the dashed periodic geodesics in gure 1.4 and the remainder of the edges of the original fundamental domain identied. The Nielsen kernel is a dierent Riemann surface S in its own right, which can be uniformized in H again so that it inherits the complete hyperbolic metric of H. But this means that on S , we obtain a new unique closed geodesic surrounding the new ideal boundary. So we can repeat the procedure to obtain S , S until S , the innite Nielsen kernel of S. We also say that S is the Nielsen extension of S . It can be shown that the Nielsen extension is unique and can be repeated innitely many times as well, to obtain the innite Nielsen extension of S. See [1] for more details about this construction.

29

Chapter 2

Moduli and Teichmller u spaces


When we look at the set of dierent complex structures on a Riemann surface, it is useful to distinguish between the dierent types (g, n, m) of surfaces. This is because one can easily make these sets into topological and metrically complete spaces by considering quasiconformal homeomorphisms and no quasiconformal homeomorphism exists between surfaces of dierent type. Furthermore, the spaces we obtain for dierent type have in general dierent dimensions.

2.1

Introduction

There are several denitions one can give for Teichmller spaces and we start u with the most intuitive denition that is given in terms of dierent complex structures on a surface. One would normally consider two Riemann surfaces of the same type to be not the same if there is no biholomorphic mapping between them. This is the space we dene below. Denition 2.1.1. The moduli space of Riemann surfaces of type (g, n, m) R(g,n,m) is the space of all Riemann surfaces of type (g, n, m) modulo biholomorphisms. However, this space turned out quite dicult to handle and therefore a covering of the moduli space is dened, which is called the Teichmller space. u This covering is obtained by dening the equivalence relation to be stronger than the one given in denition 2.1.1. Before giving the actual denition, we should discuss a few technicalities. Denition 2.1.2. Consider a Riemann surface S of type (g, n, m). Then a marking (Ai , Bi , Cj , Dk ) with i 1...g, j 1...n, k 1...m on S is a choice of an ordered set of generators of the fundamental group 1 (S) that satisfy the sole relation:
g n m

[Ai , Bi ]
i=1 j=1

Cj
k=1

Dk = 1

31

1 where the commutator [Ai , Bi ] = Ai Bi A1 Bi . A marked Riemann surface is i a Riemann surface and a marking.

We can think of the Ai and Bi as going around cycles created by handles of the surface and the Cj resp. Dk as simple loops around punctures resp. ideal boundaries. A marking with this relation can always be chosen [1], but it is in general not unique. A continuous biholomorphic map : S R from a marked Riemann surface S induces a marking on R, by the image of the loops corresponding to the generators of the fundamental group under . This means that we could equally well say that maps a marked Riemann surface to a marked Riemann surface. Denition 2.1.3. The Teichmller space of Riemann surfaces of type (g, n, m) u T(g,n,m) is the space of all equivalently marked Riemann surfaces of type (g, n, m) modulo continuous biholomorphisms, where two markings are equivalent if they are related by a conjugation with an arbitrary curve C, so that Gi = C 1 Gi C for all generators Gi and Gi . To formalize how induces a map between markings, we lift to a biholomorphic map between the universal coverings of S and R. Let D be the universal covering of S and R, S and R their respective covering groups and S and R their covering maps. Then from the previous chapter we know that a lift : D D satises S = R and the image of a homotopy class of curves corresponding to an element S S is R = S 1 . From the perspective of universal covering surfaces, the equivalence relation between dierent markings is then captured in the freedom of conjugating with an arbitrary element of Aut(H). In the next section, we will discuss the Teichmller spaces of the exceptional u types, so the Riemann surfaces that are not uniformized in H. We will proceed with the other Riemann surfaces and the real analytic theory of Teichmller u spaces, following Abiko [1]. This is closely related to the uniformization in H as discussed in the previous chapter. Finally, the complex analytic theory of Teichmller spaces of the non-exceptional types will be briey discussed. The u material of that section can be found in Nag [34].

2.2

The exceptional types

First, let us consider the Riemann sphere, P1 . By the uniformization theorem, we know that there is only one simply connected Riemann surface of type (0, 0, 0). Hence, the Teichmller and moduli space of P1 consist of a single point. u This is also the case with C and H. As we saw before, a cylinder admits C as its universal covering. Since the covering group is generated by a single element, we can always conjugate it with an element of Aut(C) such that it is generated by z z + 1. We see that all marked cylinders are equivalent and hence the Teichmller space of cylinders is u also a single point. The torus, as usual, is a special case, but most of the work was done in the previous chapter. Using the fact that any biholomorphism between tori lifts to a biholomorphism between their universal coverings, we look at tori uniformized in C. In the previous chapter, we obtained the important result that any torus 32

can be obtained as C modulo a lattice that is generated by A : z z + 1 B : z z + (2.1) (2.2)

with a certain H. We can also always lift a marked torus S in such a way that the rst map is the A cycle and the second the B cycle. Any H then denes a marked torus so the Teichmller space of tori is at most H. u To show that the Teichmller space of tori is indeed H, we need to show the u converse: that two dierent values of such that the corresponding tori are biholomorphically equivalent indeed correspond to dierent markings on the same torus. Let us try to nd such a dierent marking by considering for example: A : z z + c + d B : z z + a + b (2.3) (2.4)

with a, b, c, d integers. These are generators if one can reproduce the original generators with a linear combination of these: k(a + b) + l(c + d) = m(a + b) + n(c + d) = 1 which has the following solution: k l m n = a c b d
1

We see that, in order for the solution to be an integer matrix, one needs ad bc = 1. Having found a dierent marking for the same torus, one can now uniformize it in the canonical way, namely with the rst generator corresponding to z z + 1 and the second z z + . This is done by composing the covering map with a suitable holomorphic mapping. Such a biholomorphism is given by: w(z) = z c + d

because under w, A and B are mapped to: w A w1 : z z + 1

w B w1 : z z + with = a + b c + d

Note that ad bc > 0 if we require Im( ) > 0, so ad bc = 1. Now we can draw the following conclusions: Any marked torus corresponds to a H and two tori are biholomorphically equivalent if and are related by a P SL(2, Z) transformation. This was also claimed in the previous chapter. 33

Two marked tori are Teichmller equivalent if and only if = . The if u is trivial, the only if follows from the fact that any biholomorphism w 1 from S to S induces a new marking on S , which is a linear combination of the original generators A and B on S . If we want the two markings to be equal, then we require A = A so that w necessarily reduces to a translation and = . The above shows that the Teichmller space of tori is H and the moduli u space of tori is H modulo the action of P SL(2, Z). This concludes our discussion on the Teichmller space of Riemann surfaces u of exceptional type. In the remainder of this chapter, we will be considering non-exceptional Riemann surfaces only.

2.3

Fricke coordinates

We now turn to the Riemann surfaces with H as their universal covering surface. Remember that any such a Riemann surface can be obtained as H/ with a Fuchsian subgroup of Aut(H) P SL(2, R). The elements of are either hyperbolic, parabolic or the identity. We can always choose the generators of the group in such a way that they satisfy the relation given in denition 2.1.2. This denes a marking on S = H/ via the isomorphism between the covering group and the fundamental group 1 (S) of S. We now x the ambiguity of conjugating with an arbitrary element of Aut(H). One normalizes in H by conjugating with an element of Aut(H) such that Bg is represented by z z for some real and Ag has a xed point at 1. Note again that an element of Aut(H) is a real two by two matrix with unit determinant, so there are indeed three degrees of freedom. We used them to move three xed points to 0, and 1. Remark. This particular normalization is not always possible, for example if contains only parabolic elements then no element can be written as z z. However, it is only important that the freedom of conjugation of is xed in a similar way, which can always be done. An example of a normalizing procedure for arbitrary non-exceptional Riemann surfaces can be found in [1]. Our complete set of generators consists of two generators per handle and one per puncture and also one per ideal boundary. In total we have 2g + n + m generators, but the relation from denition 2.1.2 they have to satisfy means that one generator (for example A1 ) can be obtained as a function of the others. Better still, it is not hard to check that if is normalized, Bg and Ag can both be obtained from this relation. We then need 2g 2 + n + m elements of Aut(H) to describe a marked Riemann surface with a normalized covering group. The 2g 2 + m hyperbolic elements have three degrees of freedom, but the n parabolic elements have only two degrees of freedom because of the trace condition. This leaves us with 3(2g 2 + m) + 2n = 6g 6 + 2n + 3m real degrees of freedom. Indeed, the following theorem holds. Theorem 2.3.1. Suppose S = H/ is a Riemann surface of type (g, n, m) with 2g 2 + n + m > 0 and is normalized. If the generators Gi {A1 , B1 , ...Ag , Bg , C1 , ..., Cn , D1 , ..., Dm } 34

of are written as follows: Gi = a Gi cGi bGi d Gi P SL(2, R)

then the the set of 6g 6 + 2n + 3m real numbers (aA1 , bA1 , cA1 , aB1 , bB1 , cB1 , ..., aAg1 , bAg1 , cAg1 , aBg1 , bBg1 , cBg1 , aC1 , cC1 , ..., aCn , cCn , aD1 , bD1 , cD1 , ..., aDm , bDm , cDm ) xes S and a marking on S up to biholomorphism. Thus the Fricke map: F : T(g,n,m) R6g6+2n+3m is a bijective map from the Teichmller space T(g,n,m) to a domain Im(F ) u R6g6+2n+3m . The above map denes Fricke coordinates on T(g,n,m) . We will not try to specify the image Im(F ) here. An important remark is that this is a map from Teichmller space to R6g6+2n+3m u and not from moduli space. As with the torus, there are many points in a Teichmller space corresponding (biholomorphically) to the same Riemann surface u but with dierent markings. Again, there exists a mapping class group, whose action on the Teichmller space identies biholomorphically equivalent Riemann u surfaces. The moduli space is then the quotient of the Teichmller space with u the mapping class group.

2.4

Quasiconformal mappings

A dierent denition of Teichmller and moduli spaces is given below, which u enables us to put things in the right perspective for the Bers embedding of Teichmller space. u We start with a basis Riemann surface of certain type which we call S. A marked Riemann surface modeled on S is a triple (S, f, S1 ) where S and S1 are marked Riemann surfaces of the same type and f : S S1 a quasiconformal homeomorphism. The set of Riemann surfaces modeled on S is denoted M (S). If S has a boundary, the extension of a quasiconformal homeomorphism f to a possible ideal boundary S is well-dened by a theorem of Mori [34]. Let (S, f, S1 ) and (S, g, S2 ) be marked Riemann surfaces modeled on S and a map from S1 to S2 , as indicated in the diagram below. S g S2
<

> S1

We now note several equivalence relations and dene quotient spaces of M (S) under the equivalence relations. 1. If we just demand the existence of a biholomorphic map , then S1 is said R to be Riemann equivalent to S2 , noted as S1 S2 . The Riemann space R or moduli space of S is R(S) = M (S)/ . 35

2. If there exists a that is biholomorphic such that g 1 f : S S is homotopic to the identity, S1 and S2 are called weakly Teichmller u # # equivalent, noted S1 S2 . The reduced Teichmller space of S is T (S) = u # M(S)/ . 3. If there exists a that is biholomorphic such that g 1 f : S S S S is homotopic to the identity and keeps every point of S xed, we note (S, f, S1 ) (S, g, S2 ) and S1 and S2 are Teichmller equivalent. u The Teichmller space of S is T (S) = M(S)/ . The Teichmller space u u and the reduced Teichmller space of S are equal if and only if the ideal u boundary of S is empty. 4. We say that (S, f, S1 )(S, g, S2 ) if g f 1 is biholomorphic. Let M (S) = M(S)/. By what was said in the previous chapter about the composition of two quasiconformal mappings, the Beltrami dierentials for f and g on S coincide and indeed M (S) is the space of Beltrami dierentials on S. One readily sees that the equivalence relations are ordered from weak to strong so one relation implies the previous one in the list above. It follows that one has the natural projections M (S) M (S) T (S) T # (S) R(S) and T (S), T # (S) and R(S) can be given the quotient topology from the space of Beltrami dierentials on S. Although we used a base point or reference surface S in our denitions, by constructing a few homotopies and compositions of quasiconformal mappings it can be shown that the actual spaces are all isomorphic for any other reference surface of the same type (g, n, m). Remark. The earlier denition we gave of the Teichmller space is actually u equivalent to the new denition of the reduced Teichmller space, for a detailed u proof see [25]. This means that ocially we should have added the word reduced everywhere in the previous sections, for example the Fricke coordinates are really dened on the reduced Teichmller spaces. We chose not to because u Abiko [1] and many others also omit the word reduced in this context. How ever, in what follows, we will always mean the space T (S) = M(S)/ if we write Teichmller space. In particular, the Bers embedding that will be disu cussed concerns these spaces. Last, the Teichmller spaces for surfaces with u non-empty ideal boundary (m > 0) turn out to be innite-dimensional.

2.5

Uniformization and Teichm ller theory u

In this section, we will establish the connection between Beltrami dierentials and uniformization in H in the light of Teichmller spaces. We synthesize much u of the material previously discussed. Proofs are largely omitted, see Nag [34] for more details. As stated before, any mapping between Riemann surfaces lifts to a mapping between their universal covering surfaces. Similarly, we saw that any (p, q) tensor lifts to an automorphic form on the universal covering surface and therefore 36

a Beltrami dierential on S lifts to a Beltrami dierential on H satisfying for any : z = ( ) z so an automorphic (1, 1) form. Now it should not be surprising that the following holds. Theorem 2.5.1. A lift of a quasiconformal mapping f : S S with complex dilatation is a quasiconformal mapping f : H H with complex dilatation f 1 which is an isomorphism between . One then has S = H / with = f and . Remark. It is custom to normalize f , or rather its well-dened extension to the boundary which we also note f. This is done by composing it with a Mbius o transformation such that f xes the points 0, 1 and and we assume this holds in what follows. Extensions of to P1 Consider a lift to H of a Beltrami dierential . We can extend to P1 entirely in any way we like and we like two dierent extensions for dierent reasons. The rst one we write as and is given by: (z) z H (z) = () z L z 0 zR

Recall that we dened L as the lower half plane, L = {z : Im z < 0}, in the previous chapter. The theorems of the previous chapter guarantee the existence of a unique -quasiconformal mapping f : P1 P1 that xes 0, 1 and . z From the Beltrami equation it follows that f () = f (z) so that f (R) = R and thus f (H) = H. This is a nice result, since a lift f of a -quasiconformal mapping f can be chosen to be f on P1 for which H = H in theorem 2.5.1. Then is a Fuchsian group again and S = H/ . Also, the type of the elements of is preserved, namely if is hyperbolic, parabolic or elliptic, then so is f (f )1 . This can be seen from the fact that the type of is completely determined by the properties of its xed points (1 or 2, lying in H or R) and the xed points of f f 1 have the same properties. The extension of a lift of a Beltrami dierential has the disadvantage of not depending holomorphically on . Consider therefore a second extension: (z) = (z) z H 0 z LR

This again denes a unique quasiconformal homeomorphism f : P1 P1 that xes 0, 1 and . The homeomorphism f has complex dilatation on P1 but does not leave H invariant. The new Riemann surface can be obtained as 37

f (H)/f (f )1 . It is easily seen to be biholomorphically the same surface as H/f (f )1 by considering the fact that both are modeled on S by the same Beltrami dierential . The group f (f )1 is not Fuchsian anymore but is still a subgroup of Mb and therefore said to be quasi-Fuchsian, since it o is a quasiconformal deformation of a Fuchsian group. Similarly, the image of R, is a quasicircle on P1 . f (R)

2.6

The Bers embedding

One of the most useful and famous results in the theory of Teichmller spaces u is the embedding by Bers of the Teichmller space of any Riemann surface S u in the space of holomorphic quadratic dierentials on S, that is the space of holomorphic (2,0) tensors on S. As in previous sections, we unfortunately need a rather extensive introduction before we can come to the heart of the matter. We start by introducing the Schwarzian of a (locally) holomorphic function, basically in order to test whether it is a Mbius transformation or not. o Denition 2.6.1. The Schwarzian of a holomorphic function w = w(z), dened on a domain D P1 and with nowhere vanishing derivative, is dened as: S(w) = 3 w w w 2 w
2

where a prime means derivation with respect to z. Remark. In the literature, the notation {w, z} instead of S(w) can also be found.

By dierentiating and integrating one obtains the following properties of the Schwarzian: S(w) = 0 on a domain in P1 if and only if w acts as a Mbius transforo mation on that domain; If g : D1 D2 and f : D2 C are two locally holomorphic maps then on D1 one has S(f g) = (S(f ) g)(gz )2 + S(g). Now remember that we dened f so that it is holomorphic on L. Also, consider the extension of the action of to P1 entirely, which is well-dened since is simply a subgroup of Mb. o Lemma 2.6.2. S(f ) is a holomorphic automorphic (2, 0) form for on L.

Proof. By the quasiconformality of f , we know that = f (f )1 is a Mbius transformation. Furthermore, from the properties of the Schwarzian we o discussed it is clear that S( g) = S(g) and S(g ) = (S(g) )(z )2 for any Mb and holomorphic g. In order to show that S(f ) is an automorphic o (2, 0) form for on L, we need to prove that (S(f ) )(z )2 = S(f ), but this is now straightforward: (S(f ) )(z )2 = S(f ) = S( f ) = S(f ). The holomorphicity stems from the holomorphicity of f on L. 38

We now dene a new Riemann surface S as S = L/, which is the mirror image of S we discussed at the end of the last chapter. The space of holomorphic quadratic dierentials on it is written as B2 (S ) = B2 (L, ). By the above lemma, the Schwarzian of f |L is an element of B2 (S ). Denote [] the Teichmller equivalence class of the Beltrami dierential u M (S) and the corresponding lift of the quasiconformal mapping f. The image under f of the limit set () R of the Fuchsian group uniformizing S plays an important role, since a quasiconformal mapping homotopic to the identity moves these points in such a way that they do not change place. In fact, there is an even better property: it can be shown that f |() = f |() if and only if [] = []. It is this that result leads, see [34], to the following important theorem. Theorem 2.6.3. The following are equivalent: 1. [] = [], so and are Teichmller equivalent on S; u 2. S(f ) = S(f ) on L. Again, for the proofs of all these statements the reader is referred to [34]. Lemma 2.6.2 and theorem 2.6.3 were used by Bers to dene a mapping from Teichmller space to the space of holomorphic quadratic dierentials on S . u The actual map, the very important Bers embedding : M (S) B2 (L, ) is given by: : S(f |L )(z) = (z). By the above theorem, projects all Teichmller equivalent Beltrami dierenu tials to the same and hence can be directly used to dene a map : T (S) B2 (L, ). Since B2 (S ) is a complex vector space (of dimension 3g 3 + 2n if S is of type (g, n, 0) and innite-dimensional if m > 0), can be used to dene a complex structure on T (S) by making it locally biholomorphic. The image of T (S) in B2 (S ) is contained in the ball |||| 3/2, with the norm dened as |||| = sup(Im z)2 |(z)|.
zL

It is not very easy do determine the actual image, something we saw before when we dened Fricke coordinates on T (S). Remark. Note that we can go from the mirror image S = L/ to the actual surface S = H/ by noting that (z) B2 (S ) if and only if () B2 (S). So z S(f )() for z H is a holomorphic automorphic (2,0) form for in H and thus z projects down to a holomorphic quadratic dierential on S. This follows trivially from the denition of an automorphic form and the fact that () = (z) for z Aut(H). A local inverse of can be constructed surprisingly easily. Remember that the canonical metric on a Riemann surface is given by the (1,1) tensor (Im z)2 dzd on H. Consider a B2 (L, ). Then () is an anti-holomorphic z z (0,2) automorphic form for on H. We dene as follows: = ()(Im z)2 z 39

for z on H where the right-hand side is easily veried to be a (-1,1) automorphic form because it is the product of a (-1,-1) form and a (0,2) form. If is a Beltrami dierential, so if || < 1 a.e., then it is called Bers Beltrami dierential. Although we used the canonical metric on H, the inverse of any nowhere vanishing (1,1) tensor on S will do in constructing a Beltrami dierential. The nice property of Bers Beltrami dierentials is [25]: ( ) = (z) for z L, so Bers Beltrami dierentials are indeed a local inverse of .

Remark. From the previous remark we also see that, starting from a B2 (S), one can obtain a Beltrami dierential = (z)(Im z)2 for z H with ( ) = () for z L. z Compactication of Teichm ller space u The Bers boundary of Teichmller space is given by the natural boundary of its u embedding in B2 (S ). We know that any B2 (S ) is the Schwarzian of some deformation of and for at the boundary of the image of we know that the deformed group is isomorphic to . One distinguishes cusps, where one of the cycles of the Riemann surface is contracted to a point, and totally degenerate groups, where the Riemann surface disappears altogether. The cusps correspond to hyperbolic elements of becoming parabolic, pinching loops to a point. For totally degenerate groups the structure of the image of is not very well understood, but they densely populate the Bers boundary and it might be the kind of singularity we would encouter in our study of three-dimensional gravity.

40

Chapter 3

Three-dimensional spacetimes
As we will see in the next chapter, the wormholes that we eventually want to discuss emerge as quotients of parts of the three-dimensional spacetime that is called Anti de Sitter. This chapter serves as a nal introductory chapter and discusses three-dimensional gravity, the procedure of taking quotients and important aspects of the Anti de Sitter spacetime. We assume basic knowledge of general relativity, the lost reader is referred to the introductory book by Carroll [17] or the standard works from Wald [44] and Hawking and Ellis [23]. A general introduction to 2 + 1-dimensional gravity was written by Carlip [14]. Henceforth, in the spirit of [7], we use the units c = 8G = k = = 1.

3.1

Gravity in three dimensions

First, let us give a short introduction to general relativity in three dimensions. We discuss a toy model of gravity by suppressing one spatial dimension. This means that one considers a three-dimensional manifold with a metric of signature ( + +) that is a solution of the vacuum Einsteins equations 1 Rab Rgab + gab = 0 2 (3.1)

with perhaps a nonvanishing cosmological constant. The Lagrangian formalism is not discussed for the moment, because it is somewhat complicated by boundary terms which we do not want to dwell upon. The Riemann tensor is given by: Rabcd = b d a d + f d f d ac bc ac bf bc af where the c are the connection coecients for the Levi-Civita connection ab c = ab 1 cd g (a gbd + b gad d gab ). 2

By symmetry arguments, it is well-known [17] that the number of independent components of the Riemann tensor Rabcd in n spacetime dimensions is not n4 41

but rather n2 (n2 2)/12. Hence for n = 3, the Riemann tensor has only six independent components. But the Ricci tensor Rab = Racbc is a symmetric three by three matrix and has also six independent components. This means that all freedom is captured by the Ricci tensor. We dene the Weyl tensor Cabcd as: Cabcd = Rabcd 2ga[c Rd]b + 2gb[c Rd]a + Rga[c gd]b (3.2)

with the convention that T[bc] = (Tbc Tcb )/2. The denition (3.2) ensures that the Weyl tensor has all the symmetries of Rabcd so it has also six degrees of freedom. Now, by taking the trace of (3.2), one obtains Cacb c = 0. These six equations x Cabcd completely and one is led to the well-known result that Cabcd = 0 on any three-dimensional manifold. This is a very general result: we did not demand the metric to satisfy Einsteins equations (3.1). Let us do so now. By taking the trace of equation (3.1), we see that R = 6 and thus Rab = 2gab . With this and Cabcd = 0 in (3.2), the Riemann tensor becomes Rabcd = (gac gbd gad gbc ) (3.3)

We see that the Riemann tensor does not involve any derivatives of the metric and hence is completely locally dened. This result is a direct consequence of the fact that there are also exactly six independent Einstein equations, completely xing Rab and hence Rabcd at any point in spacetime. This means that there are no propagating degrees of freedom like gravitational waves in three dimensions. With the absence of local degrees of freedom, three-dimensional gravity looks like a rather uninteresting arena. However, we still have the freedom of varying the global properties of a spacetime, like its topology. The wormholes we will discuss in the next chapter are such examples of spacetimes of a dierent topology: locally they are completely indistinguishable from each other but globally there certainly are dierences. One way to obtain these wormholes is by taking quotients of simple three-dimensional manifolds; much in the same way as a torus and C are indistinguishable locally but globally one is the quotient of the other.

3.2

Identications and quotient spacetimes

This section discusses some very general properties of making identications in a spacetime. Just as Riemann surfaces are constructed by taking quotients of H or C, we will now take quotients of a three-dimensional Lorentzian manifold M . Formally, taking a quotient means the following: consider a group of dieomorphisms M M . Two points p, p M are equivalent if there is a such that (p) = p . The quotient M/ is then the set of equivalence classes of points of M . We cannot, in general, make any identication we like. We briey discuss three conditions below. 42

1. We want the action of to be properly discontinuous, see denition 1.3.5 on page 16. If the group acts properly discontinuously and xed-point freely, then the quotient space is a smooth manifold. Notably, we then also avoid conical singularities, although they are often used as representations of point particles in 2 + 1-dimensional gravity. 2. We require that the metric inherited from M on M/ is well-dened. The dieomorphisms generating the identications should then be isometries, meaning that for every dieomorphism , the pullback leaves the metric g invariant, so (g) = g. In components, this becomes gcd ((x)) c d a b dx dx = gab (x)dxa dxb . xa xb

We can consider one-parameter families of dieomorphisms that are generated by a vector eld (p), so dieomorphisms that move points along the integral curves of the vector eld. It is well-known [44] that these diffeomorphisms leave the metric invariant if and only if is a Killing vector eld, meaning that everywhere L gab = 0. Making the identications along the integral curves of Killing vector elds thus ensures the smoothness of the inherited metric on the quotient manifold. 3. Last, we demand that the quotient manifold be free of closed timelike curves, to avoid time travel opportunities which are seen as unphysical. If a dieomorphism is generated by a Killing vector eld , a necessary condition is that be spacelike, so that everywhere a a > 0. However, the condition is not sucient and it is easy to nd a counterexample, for example on a cylinder with a Lorentzian metric. Below, we briey discuss an example of a quotient spacetime that satises these three criteria. Example 3.2.1. Consider R2+1 , which is R3 with the Minkowski metric ds2 = dx2 + dy 2 + dz 2 and a solution of the Einstein equations with = 0. Now we make the identication (x, y, z) (x, y, z + 1). Recall from the previous chapter that this means the following: consider the cyclic group of dieomorphisms generated by : z z + 1 which we denote < >. We take the quotient R2+1 / < > and write this as above. The identication is properly discontinuous, because any subset of R2+1 intersecting with innitely many images of itself would have innite z range and hence not be compact. The identication is taken along the Killing vector eld z , so the metric descends to a metric on the quotient space as well. Finally, z is everywhere spacelike and this spacetime is obviously free of closed timelike loops. 43

Remark. Note that punctured manifolds might be considered singular as well. Punctured manifolds, however, can also arise from properly discontinuously and xed-point freely acting groups and can have an everywhere well-dened complete metric. An example is the punctured disk discussed in chapter 1. In this report, we do not exclude these manifolds beforehand and indeed a spatial slice of our zero-mass black hole will correspond to a punctured two-manifold. As we will see, the wormhole spacetimes of the next chapter all have constant negative curvature and so they are solutions of the vacuum Einstein equations (3.1) with < 0. These spacetimes are also quotients of parts of Anti de Sitter space, a maximally symmetric solution to the Einstein equations with < 0. In what follows, we only regard negatively curved spacetimes; for simplicity we use = 1 from now on.

3.3

Three-dimensional Anti de Sitter

The remainder of this chapter is focused on the three-dimensional Anti de Sitter spacetime. In this section, we will give its denition, discuss its isometries and show a Penrose diagram. The last section is dedicated to a (very) detailed discussion of two coordinate systems that are of particular importance to us. Anti de Sitter (AdS) spaces in any dimension are generally dened in terms of a hyperboloid embedded in at space. We consider the space R2,2 which is R4 with the metric ds2 = dU 2 dV 2 + dX 2 + dY 2 4 and embed a hyperboloid given by U 2 V 2 + X 2 + Y 2 = 1 where the 1 should be replaced with 1/ if = 1. The embedding induces a signature ( + +) metric on the hyperboloid with constant curvature R = 6. In the coordinate system (t , , ) given by U = cosh() cos(t ) V = cosh() sin(t ) X = sinh() sin() Y = sinh() cos() direct calculation shows that this induced metric becomes ds2 = cosh2 ()dt 2 + d2 + sinh2 ()d2 . (3.5)

(3.4)

There are closed timelike curves on the hyperboloid along the vector t because of the periodicity of t . One has, in fact, closed timelike circles of length 2 cosh() bered over the (, ) plane. The problem is resolved by going to a covering space where t ranges across R entirely. The entire spacetime has then the topology of R3 with < t < , 0 < and periodic, 0 < 2. This covering space is what we will call AdS3 in the sequel. 44

Isometries The Killing vectors of AdS3 are given by two rotations and four boosts: V U U V XY Y X XU + U X XV + V X Y U + U Y Y V + V Y (3.6)

Note that AdS3 is maximally symmetric with six linearly independent Killing vector elds. The Lie algebra they generate is so(2, 2) sl(2, R) sl(2, R). From the hyperboloid embedding in R4 , it is easy to see that SO(2, 2) is the full isometry group of AdS3 , denoted Isom(AdS3 ) henceforth. The connected component of the identity SO0 (2, 2) are the transformations generated by exponentiating Killing vectors and those are precisely the isometries that preserve the orientation of time and space. The full isometry group also follows from the so-called group model of AdS3 , where AdS3 is identied with the group SL(2, R) by the map: (X, Y, U, V ) V +Y X +U X U V Y . (3.7)

The action of SL(2, R) SL(2, R) on a point in or an element of AdS3 is then simply right and left multiplication; since these are isometries, there is a natural morphism SL(2, R) SL(2, R) Isom(AdS3 ) that is surjective but in fact a two-sheeted covering. In the sequel, when we make identications in AdS, we will only consider orientation-preserving isometries, so exponentiated Killing vectors. Penrose diagram The coordinate transformation cosh() = brings the metric (3.5) to the form ds2 = 1 dt 2 + d 2 + sin2 ()d2 . cos2 () (3.9) 1 cos() (3.8)

If we then introduce new coordinates 1 t = tan[ (t )] 2 the metric takes the form ds2 = 4 cos2 [ 1 (t + )] cos2 [ 1 (t )] 2 2 dt2 + d 2 + 2 d2 cos2 ()

which is manifestly conformally at. We can draw a Penrose diagram as in gure 3.1. We used the (t , , ) coordinates and suppressed the coordinate, meaning that every point in gure 3.1 represents a circle, with the exception of the points at = 0. The spacetime is an innite solid cylinder with = /2 on the conformal boundary (see below) and = 0 in the center. The endpoints of the cylinder correspond to future and past timelike innity, i+ and i. 45

i+ t =

PSfrag replacements t = /2

t =0 =0 i = /2

Figure 3.1: Penrose diagram for AdS3 . Every point represents a circle, except the line = 0. Future and past timelike innity are indicated by i+ and i, respectively. Spacelike and timelike geodesics are dotted and the solid diagonal lines are lightlike geodesics.

Conformal boundary We now introduce the concept of a conformally compact manifold, which we will use in subsequent chapters. Consider a manifold M which is the interior of a compact manifold M with boundary M . The metric g on M is conformally compact if there exists a smooth function on M , satisfying > 0 in M , = 0 and d = 0 at M so that g is a smooth metric on M. We then call M the conformal boundary of M. Now, such a conformal compactication then denes an induced metric on the boundary M , namely simply g|M . However, for a dierent choice of we obtain a dierent boundary metric. For example, if e , then this induces a rescaling of the boundary metric. This implies that we only have a conformal structure (dened in chapter 1) on M . A particular choice of then gives us a representative of this conformal structure. For AdS3 , we can choose M to be AdS3 plus the cylinder = /2 and = cos2 (). We then see from the metric (3.9) that on the cylinder = /2 we have an induced conformal structure dt 2 + d2 , so this cylinder is the conformal boundary of AdS3 . Of course, the representative we singled out with our particular choice of is just dt 2 + d2 .

3.4

Coordinate systems on AdS3

For later convenience, we introduce two new coordinate systems on AdS3 : the black hole coordinates and the Poincar coordinates. Since they cover only a e part of the AdS3 cylinder, we will end this section by discussing the regions of validity of these coordinates. 46

Poincar coordinates e The so-called Poincar coordinates are dened as: e x= Y U +X y= V U +X z= 1 U +X (3.10)

giving the conformally at metric ds2 = 1 (dx2 dy 2 + dz 2 ) z2 (3.11)

The metric becomes singular for z = 0. We call the region < x, y < and z > 0 with the above metric H 2+1 . In the sequel, we will often consider H 2+1 as a separate manifold for reasons outlined below. For the interested reader, the nonzero connection coecients of the metric (3.11) are x = y = z = xz yz xx z = z = z 1 and their symmetric counterparts. yy zz The Poincar coordinates are useful because of the following two-step reae soning. First, if we complexify the y = 0 surface via x + iz = w, we obtain the upper half plane H that is well-known from the previous chapters. The induced metric on it from (3.11) equals the canonical metric that was given in (1.9) on page 21. This surface has vanishing extrinsic curvature in H 2+1 . Second, in H 2+1 , the Killing vector elds (3.6) are given by: x y xx +yy + zz yx +xy (x2 + y 2 z 2 )x + 2xyy + 2xzz

2xyx + (x2 + y 2 + z 2 )y + 2yzz .

Let us now focus on the Killing vector elds with orbits in the y = 0 plane, which are the three vector elds x xx +yy + zz (x2 + y 2 z 2 )x + 2xyy + 2xzz . (x2 z 2 )x + 2xzz . (3.12)

and on y = 0 they reduce to x xx + zz (3.13)

We can integrate these and, if we again dene w = x + iz, the result is that they generate isometries of the form: w aw + b cw + d (3.14)

for some a, b, c, d R and ad bc > 0. This is exactly the group Aut(H) that was extensively discussed in the previous chapters. We conclude that the three Killing vector elds (3.12) generate a P SL(2, R) subgroup of Isom(AdS3 ) that reduce to elements of Aut(H) on y = 0. We have only calculated their form on H, but clearly (3.14) is the restriction to y = 0 of a more general isometry AdS3 AdS3 generated by a linear combination of the vector elds (3.12). We will nd these isometries in the next chapter; in Poincar coordinates, the result will be equation (4.2) on page 61. e We allow ourselves to briey announce what we will do in the sequel in order to clarify the link between the previous chapters and AdS3 . From chapter 1, we know that we can obtain Riemann surfaces by taking a Fuchsian subgroup Aut(H) and taking a quotient H/. We just found out that every element of 47

Aut(H) can be extended to an isometry of AdS3 . Since this extension is in fact a morphism, any Fuchsian group can be extended to a subgroup of Isom(AdS3 ) as well. We will then roughly proceed by taking the quotient H 2+1 / to obtain a spacetime which allows a spatial slice equal to the Riemann surface that was uniformized by . Of course, one should check that H 2+1 / is a well-dened spacetime but with certain modications indeed it is. In chapter 4, we will discuss this procedure and show how a spacetime can be associated to every Riemann surface that is uniformized in H. Black hole coordinates We introduce one nal set of new coordinates: U = r cosh() V = Y = X = r sinh() r2 1 sinh(t) r2 1 cosh(t) (3.15)

with the coordinate ranges < t, < and 1 < r < . The metric in these coordinates is ds2 = (r2 1)dt2 + dr2 + r2 d2 . r2 1 (3.16)

This metric looks very much like a black hole metric in negatively curved space [22], which is the reason for bringing up these coordinates in the rst place. We lack, however, a periodicity in the angular coordinate and so in section 4.4 we shall identify points in AdS3 such that becomes periodic in order to obtain a black hole spacetime. We can also extend the coordinates to 0 < r < 1, but then we need a dierent transformation given by: U = r cosh() V = 1 r2 cosh(t) X = r sinh() Y = 1 r2 sinh(t) (3.17)

Note that the transformation properties of t have changed as well. One readily veries that the metric is still (3.16). This implies that r is timelike and t is spacelike if 0 < r < 1.

3.4.1

Coordinate ranges

In the previous section we dened two new coordinate systems on AdS3 , namely the Poincar coordinates (x, y, z) and the black hole coordinates (t, r, ). Of e course, these are nothing more than dierent charts on AdS3 , but with a slight disadvantage: both coordinate systems are not valid on AdS3 entirely and cover the spacetime only partially. The aim of this section is to calculate and show on which subdomains of AdS3 the coordinate systems dened by (3.10), (3.15) and (3.17) are valid. 48

= =0

= 3 2 = 2

sin() = 0

sin() = 1

PSfrag replacements

= 2 t =0

Figure 3.2: Surfaces in and on the AdS3 cylinder that we will use in the sequel. We indicated the cylindrical coordinates (t , , ) in the upper left cylinder. The surface in the center is given by sin() = 0 and in the top right cylinder by sin() = 1. On the bottom row, the horizontal slice on the left is t = 0 and the conformal boundary is indicated on the right and given by = /2. The results of the calculations will be visualized in the many gures in this section. We will show a three-dimensional image of the coordinate patches lying in the AdS3 cylinder (gure 3.3 and the cylinder on the lower right in gure 3.9), as well as the intersection of the patches with three cross-sections and the conformal boundary (the gures 3.4-3.7 and the other diagrams in gure 3.9). The three cross-sections are given by sin() = 0, sin() = 1 and t = 0 and the conformal boundary by = /2, as we mentioned before. All four surfaces are shown in gure 3.2. One should keep in mind that the surface = /2 is not part of AdS3 itself so in fact we will show the intersection of the boundary of a patch with the conformal boundary of the AdS3 cylinder. Note that for example the surface given by sin() = 0 covers both = 0 and = , so we have doubled the region drawn in gure 3.1, which only showed = 0. Black hole coordinates We will rst calculate the location of the domain of AdS3 where the black hole coordinates (t, r, ) are dened. Because of the coordinate singularity at r = 1, the coordinate system consists of two patches, given by (3.15) for r > 1 and by (3.17) for 0 < r < 1. The procedure is as follows: the patches of 49

the black hole coordinates are dened by the coordinate ranges 0 < r < 1 or 1 < r < , < t < and < < . We will express these inequalities in the (t , , ) coordinates to obtain domains in the AdS3 cylinder corresponding to the black hole coordinate patches. These domains are shown in gure 3.3. Afterwards, to understand the coordinate system in more detail, we will calculate the intersection of the patches with the four slices sin() = 0, sin() = 1, t = 0 and = /2 of gure 3.2. This ultimately results in the gures 3.4-3.7. Let us now derive the region of validity of the black hole coordinates in (t , , ) coordinates. First, r > 0 means U > 0 and hence, from (3.4), cos(t ) > 0 necessarily. We choose the interval /2 < t < /2 (3.18)

but of course we could equally well have chosen t (3/2, 5/2) or similarly. The boundary r = 0 is given by U = X = 0 and hence by the line r = 0 : t = /2 sin() = 0. (3.19)

Also, since 1 < tanh() < 1 and U > 0, we have |X| < U which equals: | tanh() sin()| = | sin() sin()| < cos(t ) (3.20)

from both (3.4) and (3.8). The boundaries = correspond to X = U so = : sin() sin() = cos(t ). (3.21)

Finally, one has 1 < tanh(t) < 1, which gives dierent constraints for r > 1 and r < 1. First, if r > 1, we have Y > 0 and it implies |V | < Y , yielding sin(t ) < cos() sin() if r > 1. (3.22)

Second, for r < 1, we have V > 0 and thus |Y | < V , implying | sin() cos()| < sin(t ) if r < 1 (3.23)

The boundaries t = are given by V = Y , giving t = : sin() cos() = sin(t ) (3.24)

Note that it is the surface V = Y , so t = , that separates the regions r > 1 and r < 1. The coordinate singularity r = 1 is a line given by V = Y = 0 and this means r = 1 : t = 0 cos() = 0 (3.25) so t = 0 sin() = 1. The constraints (3.18), (3.20), (3.22) and (3.23) dene the two patches that are shown in gure 3.3. From these inequalities we also calculate the intersection of the black hole coordinate patches with the four cross-sections of gure 3.2. We will be somewhat pedantic for the rst slice, but for the other surfaces we will skip most details since the calculations are very similar. 50

t = /2 r<1

PSfrag replacements

t =0 r>1

t = /2

=0

Figure 3.3: The patches r > 1 and r < 1 in the cylinder from the same viewpoint as in gure 3.2. The reader can nd the cross-sections drawn in the gures 3.43.7 in this cylinder. sin() = 0 In the black hole coordinates, this slice is given by = 0. From (3.22), we see that the patch r > 1 has no solutions for = . Using then = 0, (3.22) implies that | sin(t )| < | sin()|. With (3.18) this means < t < if = 0 r > 1 and the boundaries t = are given by t = , as follows from (3.24). For r < 1, from (3.23), we have for cos() = 1: t /2 if sin = 0 r < 1.

and it is bounded by t = again, as well as t = /2, which is r = 0 from equation (3.19). The above two surfaces are the shaded regions in gure 3.4. Note that the line r = 1, equation (3.25), has a single intersection point with this cross-section, namely at = 0. sin() = 1 From (3.22) it is clear that the patch r > 1 does not intersect the surface sin() = 1. For r < 1, combining (3.18) and (3.23) yields 0 < t < /2 immediately. Adding the restriction (3.20), which is now | sin()| < cos(t ), gives that: 0 < t < /2 if sin() = 1 r < 1

This is shown in gure 3.5. Its boundaries, from the equations (3.21) and (3.25), are given by = and r = 1. The top of the triangle is given by t = /2 and = 0, so by r = 0 from (3.19). t = 0 This disk contains the surface t = 0. We have from (3.22): /2 < < /2 51 if t = 0 r > 1.

sin() = 0 r=0 t = /2 =0 r<1 t=

PSfrag replacements

t =0 r=1 t = t = /2 = /2 = =0

=0 r>1 r=

= /2 =0

Figure 3.4: The slice sin() = 0 corresponding to the center diagram in gure 3.2. The surfaces are in black hole coordinates given by = 0 and r > 1 or r < 1. The patch boundaries are indicated. Solid lines are lines of constant t, dotted lines are lines of constant r. Note how r becomes the timelike coordinate for r < 1. This is gure 3.6. The outer boundary of the half-disk is given by r = , and from (3.21) and (3.25), the inner one by r = 1 and the points = lie on the corners. = /2 This surface corresponds to r = . We have /2 + < t < /2 and 3/2 < t < 3/2 + if = /2 3/2 < < 2 r > 1. if = /2 0 < < /2 r > 1

by a similar reasoning as we did before. The bounding lines in gure 3.7 are found by substituting = /2 in the equations (3.21) and (3.24). The result is shown in gure 3.7, where one can see that the bounding lines are given by combinations of both = and t = . We conclude this discussion by mentioning that a second set of black hole coordinates can be used to dene a mirror patch, beyond t = , where the 52

PSfrag replacements r=0 t = /2

sin() = 1

= t=0 r<1 t =0 =0 r=1 = /2 = 3/2 =0

= /2 = /2

Figure 3.5: The slice sin() = 1. The slice is given by t = 0 and r < 1. Dashed lines are lines of constant , dotted lines are lines of constant r again. t =0 PSfrag replacements =

r=1

t=0 r>1

=0 =0

r=

Figure 3.6: The slice t = 0 or t = 0, r > 1 in black hole coordinates. The coordinate system covers half the disk. Lines of constant r are dotted, lines of constant are dashed.

location of t = is clear in for example gure 3.4. The total looks then like gure 3.8. We will not give details; One can obtain everything formally by rotating all previous pictures a half-turn around the line r = 1, so t = 0 and sin() = 1, and changing time orientation if desired. The lines r = 1 of both patches coincide and are given by (3.25). Even two sets of black hole coordinates do not cover the entire interval /2 < t < /2, as is obvious from for example the sin() = 1 diagram of gure 3.8. Two parts they do not cover are the regions beyond = , so 53

= /2 PSfrag replacements t = /2 t = t = = =

t= = t = = r= t=0

r=

t =0

t=0 t = = = = 3/2 =0

t = /2 =0

t = = = /2

= 2

Figure 3.7: The boundary surface = /2 or r = . Only the region r > 1 intersects with the boundary, but also the lines = do, which bound the patch r < 1. Note that if = /2, then both |t| and || are innite on the edges of the patch, but their sign is dierent on the various bounding lines. As usual, lines of constant t are solid and lines of constant are dashed. beyond (3.21), given by: sin() sin() cos(t ) = 0. (3.26)

Interestingly, from the metric (3.9), these are null surfaces. Furthermore, they have vertices at t = 0 given by (t , , ) = (0, /2, /2). This means that they are in fact forward lightcones emanating from these boundary points. (Actually, they are half-cones since the other half lies beyond = /2.) A symmetry argument gives that the backwards lightcones from the same points are the boundaries = of the second, rotated, set of black hole coordinates. We will need this lightcone character of the boundary surfaces in the sequel. Poincar coordinates e We now turn to the Poincar coordinates and, assuming the reader has become e familiar with the procedure, we will present the results in a single gure, namely gure 3.9. Recall how the Poincar coordinates were dened in (3.10) with e coordinate ranges < x, y < and 0 < z < and the part of AdS3 covered by these coordinates was named H 2+1 . The Poincar coordinates cover only the part of the hyperboloid where U +X e has denite sign, we chose z > 0 so U + X > 0. This part of AdS3 is in the (t , , ) coordinates given by the region cos(t ) + sin() sin() > 0. 54 (3.27)

sin() = 0 t = /2 =0 =0 PSfrag replacements t =0 =0

sin() = 1

t =0

t=0

t=0 t=0

t=0 =0 t = /2 = /2 t = /2 r <1 r= r= r=0 t = /2 =0 = /2 = = 3/2 = 2 r= r= AdS3

t =0

r>1 r>1 r <1

Figure 3.8: Region of validity of two sets of black hole coordinates in the same cross-sections as before, as well as in the AdS3 solid cylinder. Let us calculate the intersection with the slices indicated in gure 3.2. Since this time there is only one restriction, equation (3.27), we can and will be very brief here. For sin() = 0 , the inequality (3.27) becomes cos(t ) > 0 so in this slice we cover a region that is given by /2 < t < /2 if sin() = 0.

Note that sin() = 0 means X = 0 which is the surface x2 y 2 + z 2 = 1. The calculation for sin() = 1 or x = 0 is dierent for = /2 and = 3/2. For = /2 one has cos(t ) > sin() 55

so, when we choose < t < , one covers the interval /2 < t < /2 + and similarly for = 3/2 /2 + < t < /2 if = 3/2. if = /2

The upper half plane y = 0 maps to the disk t = 0 in the same way as the Mbius transformation (1.10) so that slice is completely covered and o corresponds to y = 0. Finally, also from (3.27), the boundary of our patch intersects with the boundary = /2 if cos(t ) > sin(), implying the pattern shown in the lower left gure of gure 3.9. In gure 3.9, we also indicated the four quadrants given by constant sign of x and y so, from (3.10), of constant sign of Y and V . Checking this is left to the reader. We now calculate the location of the boundaries. First, z 0 is U +X so corresponds to the conformal boundary = /2. The edges of this square are given by x and y , where the signs of x and y are determined by the signs of Y and V . The two diagonal boundary surfaces inside the AdS3 cylinder (see the bottom right gure of 3.9) are given by |x|, |y|, z = so U + X = 0. Last, the three vertices of the patch are given by U + X = 0 and V = 0 or Y = 0, so z = and either x = and y = 0 or x = 0 and y = . Their location in the (t , , ) coordinates can easily be checked and we indicated them in gure 3.9. Note that the points x, z = correspond to the same point on the boundary of the cylinder, just like the one-point compactication of the real line. The same holds for y, z = on the hyperboloid, but if the time coordinate t is not made periodic, these become two distinct points.

3.4.2

Combining the coordinate systems


y x 1 2 r = 2 (x2 y 2 + z 2 ) z exp(2) = x2 y 2 + z 2 tanh(t) =

We can write the (t, r, ) coordinates as a function of the (x, y, z) coordinates:

(3.28)

Note the symmetry under (x, y) (x, y). Since this is a half-turn around x = y = 0 or r = 1, this means that (3.28) can be used to dene two sets of black hole coordinates, which from the gures 3.8 and 3.9 are seen to be exactly the two sets shown in gure 3.8. From these gures, it is also clear that the Poincar coordinates cover more e than the two sets of black hole coordinates. Remember that in the region /2 < t < /2, the boundaries of the two sets of black hole coordinates were four half lightcones corresponding to the forward and backward lightcones from the points (t , , ) = (0, /2, /2). The regions beyond these surfaces are 56

sin() = 0 PSfrag replacements t = x, y, z = t = /2 x2 y 2 +z 2 = 1 y>0 y>0 x>0 x<0 y<0 y<0 x<0 x>0 x2 y 2 +z 2 = 1 x, y, z = t = x, y = = /2 x = z= z=0

sin() = 1 x, y, z =

y, z = z=0 t =0

y>0 x>0 x=0 y=0 x<0 y<0 z=0 x, y, z = x, z = y, z = AdS3 t = =0

t =0 z=0 t = /2

x, y = x, y = x>0 y>0 x<0 y>0 x>0 t = /2 y>0

z=0 x<0 y<0 x, y = x>0 y<0

x>0 y>0 z=0 x>0 y<0

t =0

t = /2 x, y = = 3/2 t = = 2 =0

x, y = =0 = /2 =

Figure 3.9: Region of validity of the Poincar coordinates, the four quadrants e of H 2+1 and its various boundaries. The orientation of the diagrams is equal to the previous gures. In the four slices, the boundaries of a quadrant are always solid lines. Again, the reader can check the location of the four slices in the solid cylinder.

57

not covered by the black hole coordinates. On the other hand, the Poincar e coordinates do cover two of these four regions, namely the domains beyond the half-cones from the point (t , , ) = (0, /2, +/2). These are namely the solid half lightcones x2 y 2 + z 2 0 with z > 0 in H 2+1 , for which (3.28) has no solutions. The two other lightcones have their vertex at (t , , ) = (0, /2, /2) and they form exactly the boundary of the H 2+1 patch. They are not covered by either of the two coordinate systems nor is the region beyond them. The reader is advised to check these statements in the gures 3.8 and 3.9.

58

Chapter 4

Three-dimensional wormholes
This chapter combines the results of the previous chapters in order to obtain the three-dimensional wormhole spacetimes mentioned so many times already. The spacetimes generally have the topology S R with S some non-compact Riemann surface and R the timelike direction. An extensive introduction on how we will construct these spacetimes as quotients of parts of AdS3 is given below.

4.1

Introduction

The previous chapters gave us an introduction to the theory of Riemann surfaces and to three-dimensional spacetimes. We remind the reader of some important conclusions. First, we saw in chapter 1 how many Riemann surfaces can be obtained as a quotient of the upper half plane H. They were quotients of the form H/ with a particular (Fuchsian) subgroup of Aut(H). Second, the upper half plane H embeds as a totally geodesic surface in AdS3 , namely as y = 0 in Poincar coordinates, complexied via w = x + iz. Third, if we e restricted the six Killing vector elds of AdS3 to the elds that leave this surface invariant, we found that the three remaining Killing vector elds generated exactly Aut(H). However, we did not yet derive the full form of these isometries as dieomorphisms AdS3 AdS3 . Section 4.2 will ll this gap. We will show that the extension of the isometries is a morphism from Aut(H) to Isom(AdS3 ), and that a subgroup Aut(H) becomes a new subgroup of Isom(AdS3 ). We will denote this subgroup also as . Afterwards, we are ready to take a quotient in section 4.3. The limit set of Aut(H) on the boundary of H maps trivially to a subset of the x-axis in Poincar coordinates (plus a point at innity). This set is, of course, pointwise e xed also by the new Isom(AdS3 ). Now, we will see that we can only take a quotient that satises all the criteria from section 3.2 if we rst remove all points from AdS3 that have a lightlike or timelike separation to any one of these xed points. The remainder of AdS3 of course contains y = 0 entirely and is denoted AdS3 below. The quotients AdS3 / are then the spacetimes we are 59

after. In section 4.4 we will see that, if H/ is an annulus, then AdS3 / is almost exactly a three-dimensional copy of an extended Schwarzschild black hole. This fundamental example will be extensively discussed and we show how its mass and even its angular momentum are determined by the group we choose. Finally, we will see that the quotient spacetimes generally classify as wormholes if S = H/ has ideal boundaries so if S is of type (g, n, m > 0). Section 4.5 shows how all ideal boundaries of S become asymptotic regions on AdS3 / with singularities at the beginning and endpoint of their existence as well as associated black and white hole horizons. We show that the region between the horizons and the boundary is indistinguishable from the same region in the annulus spacetime from section 4.4. Literature Historically, one rst discovered in 1992 the spacetime associated with the annulus. This is better known as the BTZ black hole, named after their discoverers Baados, Teitelboim and Zanelli [8]. It is the subject of section 4.4, which relies n heavily on the review [7]. The fact that also topologically more complicated Riemann surfaces have a spacetime counterpart was, to the best of my knowledge, rst shown in 1997 by Aminneborg, Bengtsson, Brill, Holst and Peldan in [4]. One year later, a review of these spacetimes was written by Brill [13]. From the beginning, it was clear that the BTZ black hole could also have angular momentum. That there are also spinning wormholes was explored in detail in [5] and in [12]. We do not discuss, except for a brief digression for the BTZ black hole, wormholes with angular momentum in this chapter. Finally, the papers [9] and [10] were written by Barbot at about the same time as this chapter. They provide a mathematical approach to any wormhole spacetime and also include angular momentum. They are the basis for section 4.3 and probably also prove the result of section 4.5 that we show on our own. As far as the classical description of these spacetimes goes, I think there are not many more references to be given. However, since the BTZ black hole is the three-dimensional equivalent of an extended Schwarzschild or Kerr black hole, it is (for one) a very important playground for any theory of quantum gravity. In this light, it is understandable that the original BTZ paper [8] is, to date, cited over 800 times. References to the other papers mentioned above are a lot less in number; they are mainly discussed in the context of holography in general and AdS/CFT in particular. We will touch upon this subject at the end of this chapter as well as in later chapters. Remark. Hereafter, we will always ensure that S = H/ has at least one ideal boundary. For spacetimes of the topology S R with S a closed Riemann surface, the reader is referred to the articles by Witten [45] or by Mess [33]. In particular, it is interesting to note the theorem proved in [33] that such a solution to the vacuum Einsteins equations is completely characterized by a point in T (S) T (S) with T (S) the 3g 3 dimensional Teichmller space of u the closed Riemann surface S. Remark. Unfortunately, one does not have a uniformization theorem at hand that any three-dimensional spacetime with < 0 can be obtained as a quotient 60

of a part of AdS3 , as is the case for certain three-manifolds with a metric of Euclidean signature and constant negative curvature [43]. Henceforth, we will always identify H with the surface y = 0 in AdS3 , complexied via w = x + iz. The complex coordinate will always be denoted w whereas z is always the coordinate dened by (3.10).

4.2

Extension of Aut(H) to AdS3

Already in chapter 3, we noted how the isometries generated by the Killing vector elds: x xx + yy + zz (x2 + y 2 z 2 )x + 2xyy + 2xzz

leave the plane y = 0 invariant and, furthermore, reduce to Aut(H) on this plane. In this section we calculate the form of the isometry for general value of y, using the Poincar coordinate system. We start with a coordinate transfore mation: z 2 = z 2 y2 y = y/z because then the three Killing vectors (3.12) become: x xx + z z (x2 z 2 )x + 2xz z

and their form is the same as the restriction of the three Killing vectors to y = 0, equation (3.13), with the substitution z z . But since we already know that the Killing vectors (3.13) generate exactly Aut(H), the extension of these isometries to H 2+1 is now clear: one should dene w = x + iz and then the transformations have the familiar form w aw + b cw + d (4.1)

of an element of Aut(H) and they indeed reduce to (3.14) on H. The transformation (4.1) implies: (ax+b)(cx+d)+ac(z )2 x (cx+d)2 +c2 (z )2 y y . z (adbc) z 2 2 2
(cx+d) +c (z )

meaning: (ax + b)(cx + d) + ac(z 2 y 2 ) x 1 y y(ad bc) (cx + d)2 + c2 (z 2 y 2 ) z(ad bc) z

(4.2)

is the extension of (3.14). The coordinates y , z are not everywhere valid and the fact that 4.2 is valid for |y| z is not trivial. However, the isometry (4.2) is a solution to the dierential equation implied by a certain Killing vector eld and by analytic continuation remains a solutions for |y| z as well. The reader can also check that the map from Aut(H) to Isom(AdS3 ) that maps (3.14) to (4.2) is indeed a morphism. 61

Remark. Note that the prefactor in (4.2) c2 1 = (cx + d)2 + c2 (z 2 y 2 ) (x + d/c)2 y 2 + z 2 diverges on the past and future lightcone of the point that is mapped to innity, x = d/c and y = z = 0. However, since the Poincar coordinates simply e do not cover the parts of AdS3 at or beyond the lightcones from the single point (x, y, z) = (, 0, ), so (t , , ) = (0, /2, /2), this is merely an indication of a coordinate singularity. Strictly speaking, then, (4.2) is not a dieomorphism H 2+1 H 2+1 but rather a well-dened isometry of AdS3 in ill-suited coordinates. In the next sections, we will see why we favored the H 2+1 coordinate system anyway.

4.2.1

Fixed points and lightcones

We will now discuss where identications dened from the isometries of the form (4.2) generate closed null or timelike curves, which are undesirable as discussed in section 3.2. In agreement with what was said in the introduction, we shall see that we run into trouble at points that are timelike or lightlike separated from the xed points of the isometry that lie at the x-axis. Let us consider an element of Aut(H) of the form (3.14) with ad bc = 1 and |a + d| 2, so corresponding to a parabolic or an hyperbolic element of Aut(H). Suppose c = 0. Its extension (4.2) has xed points on the x-axis that are given by: a d (a + d)2 4 x = 2c which is, in fact, easier calculated from (4.1). In any case, the Killing vector eld generating this isometry should vanish at the xed points. So if the Killing vector eld generating (4.2) is given by (x2 + y 2 z 2 )x + 2xyy + 2xzz + (xx + yy + zz ) + x = (x2 + y 2 z 2 + x + )x + (2xy + y)y + (2xz + z)z for some , R, then it should vanish at x . It is then easy to verify that = x x+ so the isometry (4.2) is generated by: (x2 + y 2 z 2 )x + 2xyy + 2xzz (x + x+ )(xx + yy + zz ) + x x+ x There are some interesting things to be said about this Killing vector eld. Taking its norm in the metric (3.11), we see that it is null at 0 = (x x )(x x+ ) + y 2 z 2
2

= x + x

(y 2 z 2 ) 2x x x+

(4.3)

= (x x )2 y 2 + z 2 (x x+ )2 y 2 + z 2 These are four surfaces in H 2+1 that are exactly the forward and backward lightcones of the xed points. The vector eld is spacelike above these cones (in 62

the part including H), timelike on the other side of them and spacelike again in the domain that lies under two of these cones. Furthermore, we see that this vector eld completely vanishes on the lines given by x= 1 (x+ + x ) 2 1 z 2 y 2 = (x+ x )2 4 (4.4)

so these lines are pointwise xed as well. These are easily veried to be exactly the intersection lines of the lightcones (4.3), both for y > 0 and for y < 0.

4.3

Quotients

Consider a nitely generated Fuchsian group Aut(H) such that H/ is a Riemann surface of type (g, n, m) with m > 0. From the theory of Riemann surfaces [34], we know that m > 0 is equivalent to demanding that the limit set of the Fuchsian group should not be R entirely but rather a subset of it. If we denote the limit set () R, then we suppose () = R. The group is isomorphic to a subgroup Isom(AdS3 ) via the extension (4.2) and () now becomes a subset of the x-axis in Poincar coordinates plus a point at innity; e so a subset of the circle t = 0 = /2. We dene AdS3 {p AdS3 : p () : d(p, p ) > 0} with d(p1 , p2 ) the geodesic distance function between two points in AdS3 implied by the metric (3.5). In fact, this distance is either 0 or for points on the surface = /2 but its sign is always well-dened. Alternatively, one might calculate the distances using the metric g of the conformal compactication of AdS3 , introduced at the end of section 3.3; the resulting region AdS3 is the same. Now for such a and the corresponding AdS3 , the following holds. Theorem 4.3.1. The quotient M 2+1 AdS3 / is a manifold, inherits a metric from AdS3 , and it is free of closed timelike curves. This theorem is proved in a more general setting by Barbot in [9]; there, the domain AdS3 is suitably called the invisible domain for () since it is the set of all points with a spacelike separation to all of (). A few comments on theorem 4.3.1 are given below. Note that the Poincar coordinates can be very useful in visualizing the e spacetime, because its conformal atness makes it easier to nd the lightcones that bound AdS3 . As is clear from gure 3.9 from the previous chapter, the Poincar coordinates cover exactly one period of the hye perboloid with < t < minus the future and past lightcone from (t , , ) = (0, /2, /2). This means that H 2+1 covers any possible AdS3 completely as long as one of the limit points of lies at innity. In the 63

sequel, we will always conjugate with an element of Aut(H) such that one xed point lies at innity and then write H 2+1 for AdS3 in Poincar e coordinates. Second, if is non-trivial, we know that it has at least one xed point. So for any nontrivial we exclude at least one future and one past lightcone from this xed point. If we move this xed point to innity, then the remaining region is always exactly covered by a Poincar patch and from e for example gure 3.9 it is then clear that any possible AdS3 can never extend to a full period of t , since the surfaces t = and t = are always completely removed from it. By construction, if we follow the procedure outlined above of extending a Fuchsian group from H to H 2+1 , the spacetime is uniquely determined by the Fuchsian group acting on H. This means that all the information on the spacetime is captured by the point in moduli space of the Riemann surface S = H/ that represents the y = 0 slice. We now proceed by discussing the BTZ black hole as the spacetime associated with the annulus. We end the chapter by discussing the boundaries of more general wormholes and demonstrate the equivalence of the outer regions to the BTZ outer regions.

4.4

The BTZ black hole

The most famous three-dimensional spacetime that can be obtained as a quotient of some AdS3 is the BTZ black hole. It is the three-dimensional analogue of the extended Schwarzschild or Kerr black holes but it is asymptotically Anti de Sitter rather than asymptotically at. In this section, we will discuss the BTZ black hole in detail and also show that it is the spacetime obtained from the annulus of example 1.5.4 in chapter 1. This means that its covering group is cyclic and its limit set on R consists of only two points, which simplies the discussion considerably and makes it a suitable example to start with. From the black hole coordinates (t, r, ) dened in (3.15) and (3.17), we see that we probably need an identication along the Killing vector like (4.5) below. We will show that this is indeed the way to go and that the quotient spacetime is indeed a black hole with a horizon and a singularity. In section 4.4.1, we rst consider the quotient (t, r, ) (t, r, + 2). (4.5) which we will see gives a black hole spacetime of unit mass. In the parts 4.4.2 and 4.4.3, we will see that mass and angular momentum can be given to the black hole by slightly changing (4.5).

4.4.1

Unit mass BTZ black hole

This section considers the identication (4.5). We rst discuss the excluded regions and the quotient procedure in detail, using the various coordinate systems dened on AdS3 in chapter 3. To gain more insight in the new spacetime, we give a Penrose diagram and demonstrate the existence of a singularity hidden behind a horizon. 64

AdS3 and the black hole coordinates The identication is generated by , which is equal to XU + U X from (3.15) and (3.17). This vector eld is spacelike only when ds2 ( , ) > 0 which means X 2 + U 2 > 0. and we dene AdS3 as the connected component of AdS3 where |U | > |X| that includes the disk t = 0. At the boundaries of AdS3 we have |U | = |X| so becomes null. Equation (3.21) shows that this boundary consists of the four surfaces = of both sets of black hole coordinates. The domain AdS3 is thus exactly the region covered by two black hole coordinate patches, so the shaded region in gure 3.8. In gure 4.1, we repeated it; the solid lines indicate the boundary of AdS3 . This also shows that the above denition of AdS3 agrees with the above theorem: from the discussion around equation (3.26), we know that the four surfaces = or U = X are lightcones and we will see below that their vertices indeed lie at the xed points of the isometry at t = 0. The lines X = U = 0 are special since the Killing vector vanishes completely on them. From (3.19) we know that the lines U = X = 0 are exactly r = 0 in black hole coordinates. This implies that X = U = 0 are the pointwise xed lines of (4.5) similar to the lines (4.4). As the lines (4.4), they are also given by the intersection of the excluded lightcones, that are now given by U = X. In AdS3 , we have to identify the points at (t, r, ) and (t, r, +2). As before, we can also pick a fundamental domain whose edges are to be glued together. In gure 4.1, we visualized such a fundamental domain, given by < < . Below, we will see that indeed the slice t = 0 or y = 0 is the annulus and also demonstrate the existence of the horizons already shown in gure 4.1. H 2+1 and the annulus We take a brief look at the BTZ black hole in Poincar coordinates. Since, from e (3.28), = xx + yy + zz , the identication (4.5) becomes: (x, y, z) e2 (x, y, z). (4.6)

The isometry clearly leaves the plane y = 0 invariant. With w = x + iz, its restriction to y = 0 is given by w e2 w (4.7)

which means that the quotient space which is the restriction to y = 0 of the quotient spacetime is indeed the annulus from example 1.5.4 in chapter 1 and so the BTZ black hole is indeed the three-dimensional spacetime obtained from the annulus. We can also consider the excluded regions in the Poincar coordinates and e verify that they are indeed the regions beyond the four (half) lightcones with vertices at the xed points of (4.6) at the x-axis. These points are are the 65

sin() = 0 t = /2

sin() = 1

= t =0 =

=0 PSfrag replacements t =0 t = /2 AdS3 t = /2

= = =0

= /2 t =0 t=0 =0

Figure 4.1: A fundamental domain for the BTZ black hole and three of the four familiar slices. In the lower right gure, there is a mirror image corresponding to /2 < t < 0 below the t = 0 disk at the bottom of the cylinder, which for clarity is not shown. The dotted lines are removed from AdS3 , the solid lines bound AdS3 and the dashed lines bound the shaded fundamental domain. Opposite points, corresponding to = , are glued together at the quotient spacetime and the resulting surface at t = 0 is an annulus. The darker shaded surfaces are the horizons. point at innity (x, y, z) = (, 0, ) and the origin of H 2+1 . We mentioned before that the two lightcones from the point at innity are not covered by the Poincar coordinates anyway, but the two lightcones from the origin are and e they are given by x2 y 2 + z 2 = 0. This is indeed where xx + yy + zz , so , becomes null, and also where = . We see again that AdS3 or H 2+1 , which is now given by the domain of H 2+1 where x2 y 2 + z 2 > 0, is precisely the region covered by the two black hole coordinate patches. Of course, the domain H 2+1 could also be directly obtained from equation (3.10) and the requirement |U | > |X|. 66

Note that the pointwise xed line U = X = 0 or r = 0 does not lie in H 2+1 , which is a consequence of the fact that c = 0 for the identication (4.7) so (4.4) does not hold. Comparing the sin() = 0 diagrams from the gures 3.8 and 3.9 yields that r = 0 lies on the boundary of the Poincar patch, with e x, y, z while keeping x2 y 2 + z 2 = 1. Penrose diagram The Penrose diagram of this spacetime is easily obtained from the sin() = 0 slice in gure 4.1, where = 0. Before the identication, each point of the shaded region this gure represented a line corresponding to < < . Now, because of the identication (4.5), one has instead a circle bered over every point in the diagram, but the diagram remains unaltered. It is thus a Penrose diagram of the unit mass BTZ black hole. We repeated the diagram in gure 4.2. We remark that this spacetime is not globally hyperbolic, as is clear from the Penrose diagram, so in particular the annulus at y = 0 is not a Cauchy surface. =0 r=0

PSfrag replacements r= sin() = 1 AdS r = 1


3

r=

r=0 Figure 4.2: Penrose diagram of the BTZ black hole. Every point in the diagram corresponds to a circle. We can cover the spacetime with two black hole coordinate patches and there are two boundaries corresponding to r = and two singularities corresponding to r = 0.

Singularities and horizons Are there any singularities in this spacetime? Good candidates for singularities are the lines r = 0. However, since the Einstein equations (3.1) completely x the curvature at any point in spacetime, the curvature does not blow up at r = 0, as for example with the singularity of a Schwarzschild black hole. Rather, the spacetime simply ends at r = 0: beyond these lines, there is no spacetime. This, in turn, implies that there are inextensible geodesics that have an endpoint at r = 0. In the literature, the endpoints of inextensible geodesics are commonly called singular points as well, see for example Wald [44]. It is on this ground, then, that we call the lines r = 0 indeed singular. 67

The endpoints of the spacetime also yield horizons: since the spacetime ends at t = /2 and r = 0, the horizons are by denition [44] given by backward lightcones from the intersection points of this line with spatial innity. These backward lightcones lie with their tip at the points (t , , ) = (/2, /2, /2). One easily checks from the similarity in the gures 4.2 and 3.4 that they are precisely given by the null surface t = + (and the line r = 1 at t = 0) and similarly for the mirrored region. By the same argument, the surfaces t = are the white hole horizons. In the Poincar coordinates, the black and white hole horizons are given by e x = y, which follows directly from (3.28). In this coordinate system it is clear that these surfaces are invariant under (4.5), so they indeed descend to horizons on the quotient spacetime. All the horizons intersect the annulus y = 0 at the single periodic geodesic on it, the line x = y = 0, as was already shown in gure 4.1.

We have now covered the unit mass black hole and showed some of its geometrical properties. Next, we discuss how we can change the properties of the black hole a bit. We will see that there exist similar black holes with masses M not equal to 1 and that the spacetime with mass M = 1 surprisingly is AdS3 again. Furthermore, one can introduce angular momentum to construct the three-dimensional equivalent of a rotating black hole. We restrict ourselves to a pure gravity context so charged black holes will not be discussed.

4.4.2

Mass

In (4.5), the periodicity in the angular coordinate was chosen to equal 2, but we can make another identication: (t, r, ) (t, r, + 2). with some strictly positive real number. Since the same Killing vector eld generates the covering group, our discussion of excluded regions, singularities and horizons will not change. The resulting spacetime is, however, dierent since for example the horizon will have a dierent length. The dierent properties of the spacetime can be made explicit in the metric by performing a coordinate transformation: t = t = r = r / such that has the usual periodicity 2. In the primed coordinates, the metric becomes dr 2 ds2 = (r 2 M )dt 2 + 2 + r 2 d 2 (4.8) r M 2 with M = a dimensionless mass parameter. The horizons at r = 1 are now at r = = M so the length of the horizon is 2 M. Zero mass and the punctured disk Let us look at what happens for M = 0. From (4.8), one has the metric ds2 = r 2 dt 2 + 68 dr 2 + r 2 d 2 r2

which is also a space of constant negative curvature for r = 0. The coordinate transformation: 1 r = r brings the metric to the upper-half space form ds2 = 1 r
2

(dt 2 + dr

+ d 2 )

with x = , y = t and z = r . The identication is now given by (x, y, z) (x + 2, y, z), which is a parabolic Mbius transformation in the y = 0 slice. In o fact, it is exactly the one giving rise to the punctured disk that was example 1.5.3 in chapter 1. The puncture lies at r = or r = 0. So, surprisingly, the zero mass black hole is not empty AdS3 , but rather still a quotient of a part of it. Note that the identication is generated by the everywhere spacelike Killing vector x , so we do not have to exclude any parts of H 2+1 . Of course, we do have to exclude everything at and beyond the lightcones from the xed point at innity, so in this case AdS3 = H 2+1 . Empty AdS3 as M = 1 We saw that a Fuchsian group with a single hyperbolic generator that was extended to AdS3 entirely gave rise to a black hole spacetime. Furthermore, the unique periodic geodesic on the t = 0 surface happened to coincide with the intersection of the horizons with this surface and its length was parametrized by a mass M > 0. In chapter 1, we noted a correspondence between the trace of an element Aut(H) and the length L(C) of this geodesic C, which is given in terms of M by: |Tr()| = 2 cosh(L(C)/2) = 2 cosh( M ) In the limit M 0, the generator became parabolic with a trace equal to 2 so we could say this formula still holds for M = 0, although formally there is no notion of a periodic geodesic anymore. If we yield to the temptation of extending the above formula to negative masses, one expects the transformation to become elliptic. Indeed, this is the case: for 1 < M < 0 the metric (4.8) has a conical singularity. For M = 1 one has |Tr()| = 2 again. This means that =< > is either generated by a parabolic element or it is trivial. That is in fact trivial follows from (4.8) with M = 1 and the coordinate transformation r = sinh(): one then obtains the empty AdS3 metric (3.5) again. We thus conclude that M parametrizes a set of elements of Aut(H), which are hyperbolic for M > 0, parabolic for M = 0, elliptic for 1 < M < 0 and is the identity for M = 1. For more negative mass, there are only conical singularities, this time with excess angle.

4.4.3

Angular momentum

For completeness, we briey discuss a spinning black hole that can be obtained as a quotient of some AdS3 . Such a spacetime is obtained by boosting the identication. The new Killing vector is given by r+ + r t = r+ (xx + yy + zz ) + r (yx + xy ) 69 (4.9)

with r+ , r nonzero constants and r+ |r |. It is clear that the isometry it generates does not leave y = 0 invariant if r = 0 so then we are outside the scope of theorem 4.3.1. Nevertheless, a good quotient spacetime still exists. We want a new angular direction that runs along this Killing vector, so we need dierent coordinates (, , ) given by = r+ + r r2 = We subsequently identify (, , ) (, , + 2) which is easily veried to be generated by (4.9). Since this identication does not leave the y = 0 plane invariant, our intuitive pictures of the annulus or the punctured disk are not so easy to nd this time. Furthermore, the Killing vector (4.9) becomes timelike whenever 1 r r+
2 2 2 r 2 r2 . r+

t = r+ + r

(4.10)

(x2 y 2 ) + z 2 = 0.

We again exclude the part of spacetime where the left-hand side is zero or negative. Starting from (3.16), the metric in these coordinates is given by:
2 2 ds2 = (2 r+ r )d 2 + 2r+ r d d + 2 d2 +

2 2 (2 r )(2 r+ ) 2 2 d2 r+ r 2 d + 2 2 )(2 r 2 ) + d + 2 d 2 ( r +

2 d2 2 2 (2 r )(2 r+ )

Now we dene:
2 2 M = r + + r

J = 2r+ r

2 r =

J2 M 1 1 2 2 M

1/2

(4.11)

and the metric becomes ds2 = N 2 d 2 + N 2 d2 + 2 d + with N 2 = 2 M + J2 . 42 J d 22


2

(4.12)

This is the standard form of the metric in the outer region of a spinning black hole in Anti de Sitter spacetimes and this three-dimensional version was found in the original paper [8]. We do not, for now, discuss any details of the spinning BTZ black hole but will instead discuss more complicated topologies. 70

Figure 4.3: A spacetime associated with a Riemann surface of type (2, 0, 2) has two cylindrical boundary components.

4.5

Outer regions

In section 4.3 we have associated a spacetime with every non-compact Riemann surface S that is uniformized in H. If S = H/ for some Fuchsian , then a corresponding spacetime M 2+1 is obtained as H 2+1 /. In general this spacetime has multiple boundary components. More precisely, if S is of type (g, n, m) and hence has n punctures and m ideal boundaries, then the boundary of M 2+1 will have n one-dimensional and m cylindrical components. An example of this is shown in gure 4.3 for n = 0 and m = 2. In this section, we take a closer look at the regions of M 2+1 close to the m cylindrical boundaries. Below, we will show how each of the m cylindrical boundaries has a natural starting point and an endpoint. Furthermore, lightcones emanating from these points will form white and black hole horizons for this boundary. An outer region of M 2+1 is then a part of M 2+1 between a cylindrical boundary and its corresponding horizons. The topology of such an outer region is always an annulus times R, with R the timelike coordinate and the annulus a part of S like on the right side of gure 4.4. Now, the key result of this section will be that the outer regions of M 2+1 are exactly the BTZ black hole outer regions. This means that every outer region can be covered with a single chart with the metric (4.8) for some mass parameter M and with coordinate ranges t R, r > M and periodic, 0 < 2. It also implies that the Penrose diagram of any outer region equals one of the two r > M parts of the Penrose diagram of BTZ (gure 4.2), as is shown on the left in gure 4.4. In order to show the equality of the outer regions (up to the mass parameter M ), we rst briey return to the theory of Riemann surfaces.

4.5.1

Riemann surfaces revisited

Let us consider the Riemann surface S of type (g, n, m > 0), obtained as H/ or M 2+1 |y=0 . The intersection of the outer regions with y = 0 are annular parts of S close to its ideal boundaries like the lightly shaded domain on the right in gure 4.4. We will rst split S in these m annular parts and its remainder. 71

r=

M ?

PSfrag replacementsr =

M ? r=

Figure 4.4: The outer regions of any wormhole spacetime M 2+1 have the Penrose diagram indicated as the lightly shaded region on the left. Every point in it corresponds to a circle, but we do not know in general what lies beyond the horizons in the darker parts. The thick horizontal line is part of the slice y = 0 and the intersection of the outer region with this slice corresponds to an annular end of S. Afterwards, we can lift this splitting to H to see how each annular part is uniquely determined by a single real parameter. Splitting S A result from section 4.5.2 below will be that the white and black hole horizons intersect S exactly at the unique periodic geodesics that are homotopic to the corresponding ideal boundary components. The dashed line on the right of gure 4.4 is thus not only bounding the outer region on S; it also is precisely a periodic geodesic on S. The intersection of an outer region with S then lies between an ideal boundary component and the periodic geodesic homotopic to it. We label these annular ends of S as Sj with j = 1 . . . m. But as dened in chapter 1 on page 29, if we cut o the Sj from S we are left with exactly the Nielsen kernel SK . We have: S = S 1 S m SK For the trinion we have indicated this splitting into SK and the three Sj components on the left in gure 4.5. The aim of this section is to show the uniqueness up to a single parameter of such Sj ; extending this result to a wormhole spacetime is done in section 4.5.2. Splitting H From the theory of Riemann surfaces, it is well-known [25] that SK lifts to the part of H that is the convex hull of the limit set of , which we call HK . The Sj lift thus to H\HK and if we write Hj for the subset of H that descends to 72

Sj , we have H = H K H1 H m . Of course, there are innitely many lifts of every Sj and of SK in H, but it is essential to us that every point in HK descends to a point in SK and similarly for the Hj and Sj . The separation of H is shown on the right hand side of gure 4.5 and discussed more extensively below. Note that each component of H is invariant under , so for example: (HK ) = HK since the image under any element of of any point in HK still should descend to SK and so is by denition still part of HK . It is then not hard to see that HK / = SK . These two results of course also hold for the Hj .

S3 PSfrag replacements SK S1 S2

Figure 4.5: On the left, the splitting of a trinion in its Nielsen kernel SK and its three Sj . On the right, an idea of the splitting of H.

The structure of Hj What do the Hj look like as a subset of H? Since the Sj are bounded by the ideal boundary of S and a periodic geodesic, the Hj must be bounded on one side by the boundary R of H and on the other side by a lift of the periodic geodesic. Again from chapter 1, we know that such a periodic geodesic on S lifts precisely to the geodesic connecting the xed points of an element of j generating the corresponding loop. The domain between this geodesic and the boundary R is a semidisk in H and descends to Sj : it is thus a part of Hj . However, in general, there are innitely many other elements in , all conjugate to j , that also generate the simple loop homotopic to the boundary. This means that any Hj consists in fact of innitely many semidisks, all bounded by a part of R and the geodesic between the xed points of an element j 1 for some . Some of these semidisks are shown on the right in gure 4.5. Remark. If S is a non-compact Riemann surface with ideal boundaries, then the limit set of the Fuchsian group uniformizing it is a nowhere dense subset of R [34]. The structure of the limit set and hence of H\HK is that of a fractal, so it would be pointless to try nding it entirely. We now focus on only one of the semidisks, namely the one that is bounded by the geodesic associated to j , and denote it Dj . As we said before, it descends 73

completely to Sj . In fact, since it is simply connected, it plays the role of a universal covering of Sj (note that all points on Sj lift at least once to Dj since Sj is connected). Because the fundamental group of Sj is cyclic, we see that there can be only one cyclic subgroup of that identies points within Dj . This is precisely the subgroup generated by j , denoted < j >. The elements of \ < j > then map Dj to another place on H. More formally, we have for all : (Dj ) = Dj < j > (4.13) and (Dj ) Dj = Now recall that taking a quotient like Sj = Hj / denes equivalence classes of points. But since every connected component (semidisk) of Hj is mapped to Dj by some element of (recall its transitivity properties), we can equally well take the equivalence classes of points within D j only. But from (4.13) and (4.14) it is clear that we then only have to consider the subgroup < j >, which means that we can also obtain Sj as the quotient: Sj = Hj / = Dj / < j > . This quotient procedure is visualized in gure 4.6. Finally, note that we can always conjugate such that a new j is of the form j : z z Dj = H|Re z0 . After conjugation, Sj is the surface given by: S j = Dj / < j > . (4.15) < j > . / (4.14)

with R+ \{1}, and such that

PSfrag replacements

p1

p2

Figure 4.6: One of the semidisks of the lift of an Sj . It is divided in many dierent fundamental regions bounded by the solid lines. The edges of a fundamental region should be glued together to obtain Sj from it.

74

This makes it clear that the disconnected parts Sj on S are equal to half of the familiar annuli of example 1.5.4, cut o along their periodic geodesic. They are uniquely determined by ; in particular, they are independent of the details of the rest of S.

4.5.2

Extension to a spacetime

Having shown the uniqueness up to a single parameter of the Sj , we will now show what all of the above has to do with the outer regions on M 2+1 . We start from the same construction with a semidisk like in gure 4.6 that we label Dj . Denote the xed points of j on the x-axis as p1 and p2 . The four lightcones with their tip at these points are part of the boundary of H 2+1 and are the four dark cones in gure 4.7. Their lines of intersection are given by equation (4.4). We will be particularly interested in the intersection of these lines with the plane z = 0, which will correspond to a starting point and endpoint of spacelike innity on M 2+1 . We call the endpoint p3 and the starting point p4 . In Poincar coordinates, they are given by ((x1 + x2 )/2, |x1 x2 |/2, 0) if x1 , e x2 are the x-coordinates of p1 , p2 . The backwards lightcone from p3 is easily calculated to be given by 1 x (x1 + x2 ) 2
2

1 (y |x1 x2 |)2 + z 2 = 0. 2

(4.16)

This surface intersects the plane y = 0 at 1 x (x1 + x2 ) 2


2

+ z2 =

1 (x1 x2 )2 4

which clearly is the geodesic connecting the xed points of j . A similar reasoning holds for the forwards lightcone from p4 . These lightcones are the two transparant cones in gure 4.7. 2+1 We now dene the region Dj as the connected component of H 2+1 that is bounded by the lightcones (4.16) and the plane z = 0 that includes the semidisk Dj . Note that all the cones in gure 4.7 are uniquely determined by p1 and p2 . Now, since the covering group maps lightcones to lightcones, the image of these cones under an arbitrary element of are determined uniquely by the image of the points p1 and p2 on R. But then one readily sees that our discussion about 2+1 Dj holds for Dj as well. In particular, the equations (4.13) and (4.14) hold 2+1 identically for Dj . Hence, we can take a similar quotient and obtain a part 2+1 of Mj dened as 2+1 2+1 Mj = Dj / < j > .
2+1 We now show is that Mj is the outer region we are looking for. The 2+1 is drawn in gure 4.8. As in gure 4.6, it can be boundary z = 0 of Dj divided in innitely many fundamental regions. In particular, comparing the gures 4.6 and 4.8 illustrates how a circular boundary component of S becomes a cylindrical boundary component on M 2+1 . We also see that such a boundary component has an endpoint at p3 and a beginning at p4 . On M 2+1 , the lightcones emanating from these points are (by denition) the horizons associated with these endpoints. These lightcones lift to lightcones with their tip at p3 and p4 , which are exactly the cones (4.16). Now, since the lifts of the horizons

75

z 6 Q ykQ 0 p3 p2 x

p4 p1

Figure 4.7: The covering of an outer region. The bottom plane is z = 0 and the straight line is the x-axis. The dark cones are excluded from spacetime and the lightcones emanating from their intersection at z = 0 are a lift of the horizons of M 2+1 . These lightcones intersect y = 0 at the periodic geodesic connecting the xed points of the corresponding isometry. The gures 4.6 and 4.8 are the y = 0 and z = 0 cross-sections of the region below the horizons, respectively.
2+1 2+1 on to H 2+1 bound Dj , it follows that the horizons themselves bound Mj 2+1 lies between the j-th cylindrical boundary comM 2+1 . This means that Mj 2+1 ponent of M 2+1 and the associated horizon, so Mj is indeed the outer region for this boundary component.

p3

y PSfrag replacements x p1 p2

p4 Figure 4.8: The boundary at z = 0 of the region between the four lightcones of the previous gure. Dierent fundamental regions are bounded by the solid lines. The points p1 -p4 are xed points of the isometry j . This gure is orthogonal in H 2+1 to the surface in gure 4.6, with the x-axis as their line of intersection.

76

Uniqueness of the outer regions If we now apply the same isometry such that Sj is given by (4.15), then the xed points of j lie at zero and at innity. Two of the excluded lightcones emanate from the origin of H 2+1 and two lie at innity so are not covered by the Poincar coordinates. The hyperbola (4.4) is, like in the BTZ case, also e mapped beyond the reach of these coordinates and the lightcones (4.16) now become the surfaces x = y. On one side of them, namely in the wedge 2+1 Dj : {(x, y, z) H 2+1 : |y| < x} we have exactly the same setup as for a BTZ black hole. The region 2+1 2+1 Mj = D j / < j > is an outer region on M 2+1 . However, via (3.28) followed by the coordinate transformation in section 4.4.2, we can put the metric in the form (4.8) with r > M and the usual periodicity in . This shows what we claimed in the introduction to this section: the outer regions of M 2+1 are always equal to a BTZ outer region.

4.6

Non-fuchsian isometries

We briey discuss the isometries that do not leave the plane y = 0 invariant, like the spinning BTZ black hole. Allowing for a group of isometries whose generators are of the most general type, namely generated by an arbitrary linear combination of the six Killing vectors of H 2+1 one has to be more careful about the concept of closed timelike curves. There can only be closed timelike curves in H 2+1 / if a point is identied with a point that lies within its future lightcone in H 2+1 , since this is the only way a timelike path can close in H 2+1 . Since the covering group is transitive, there is an element of for every identication of points in H 2+1 to obtain the space H 2+1 /, so if closed timelike curves appear then there is also an element of that identies points within each others future and past lightcones. Note that the associated Killing vector may be spacelike, as mentioned in section 3.2. A necessary and sucient condition for the absence of closed timelike curves is thus that no element of maps a point to its own future lightcone. So, consider an arbitrary nontrivial element . Now the demand ((x) x)2 ((y) y)2 + ((z) z)2 > 0 (x, y, z) H 2+1

is sucient for not to generate closed timelike curves. If this holds for all then H 2+1 /, if it is a manifold, is free of closed timelike curves. In [9] it is proved that the above holds in the domain invisible to the spatially separated xed points of . The quotient of this domain with respect to is a well-dened manifold that is free of closed timelike curves. This, however, does not seem to be a maximal manifold in for example the rotating BTZ case. There, from [7], we know that there are new asymptotic regions beyond the horizon (similar to a Kerr black hole), which are now excluded from spacetime since they have a timelike separation to a xed point. 77

Chapter 5

Toroidal wormhole
In this chapter we will discuss in detail a specic example of a 2 + 1-dimensional wormhole with one asymptotic region. A spatial slice will be a Riemann surface of type (1, 0, 1), the toroidal wormhole that we pictured already in gure 1.6 on page 28. In the description of the spacetime as a quotient of a part of H 2+1 , the metric does not give us much information about the spacetime and all the global information is captured by the transition functions (the covering transformations). Our main objective in this chapter is to nd new coordinates in which the metric explicitly shows some physical properties of the spacetime and to nd a description of the spacetime purely in terms of these new coordinates, so completely independent of the description in H 2+1 . We will start in section 5.1 with the denition of the spacetime as a quotient of a subdomain of H 2+1 . Then, in the sections 5.2 and 5.4, we proceed by covering rst the space and then the spacetime with two patches that have both the topology of an annulus (plus time). Section 5.3 deals with some issues that arise when we try to cut the spacetime in two this way. Finally, in the last two sections of this chapter, we will be able to dene coordinate systems on these patches. We do this in two steps: in section 5.5 we rst introduce an easy version of the nal coordinate systems, which is helpful in understanding the nal coordinate system of section 5.6. In this latter coordinate system the metric indeed explicitly shows (at least some of) the physics involved.

5.1

Denition of the spacetime

As usual, we restrict ourselves to the isometries of H 2+1 that leave the y = 0 plane invariant and form the subgroup Aut(H) of Isom(AdS3 ). Below, we rst nd a Fuchsian group such that S = H/ is a Riemann surface of type (1, 0, 1). Afterwards, a spacetime is obtained by extending the action of the isometries of Aut(H) to the corresponding H 2+1 and taking a quotient, as was shown to be possible in the previous chapter. 79

5.1.1

Initial data

Let us rst consider H and see what can be said about the generators of if we require S to be of type (1, 0, 1). From chapter 1, we know that we need two generators that are both hyperbolic elements of Aut(H). Using conjugation freedom to map its xed points to 0 and innity, the rst generator is without loss of generality given by: 1 : w w with > 1. A second generator will be of the form: 2 : w aw + b cw + d

We now use the fact that we can always set ad bc = 1 to eliminate b. Also, with the single degree of conjugation freedom left we make sure 2 maps innity to one, so a = c. This can always be done if we also use the freedom to invert 2 . One then has: 1 aw + d a 2 : w aw + d Since 2 should be hyperbolic, we need (a + d)2 > 4. With these two generators, we have the right number of degrees of free dom (three) but the geodesics connecting their xed points on R must intersect somewhere in H, because if not the quotient surface would be a pair of pants and not a toroidal wormhole. This means, since the xed points of 1 are at 0 and , that the xed points of 2 must be nite and of opposite sign. A short calculation of 2 (w) = w gives that this means that ad > 1. Now multiplying numerator and denominator with a and setting a2 = and ad = , one obtains: 2 : w 1 1 w + (5.1)

and with > 0 and > 1 we have (a + d)2 > 4 so 2 is also hyperbolic for this range of parameters. Note that no generality is lost by choosing the generators this way. Our requirement that is mapped to 1 means that the positive xed point of 2 lies between 0 and 1 and is attractive. The commutator [1 , 2 ] generates a loop around the ideal boundary. One should remember that in fact every element in the covering group that is con1 1 jugate to [1 , 2 ] generates this loop and that we dened [1 , 2 ] = 1 2 1 2 previously. One has: [1 , 2 ] : w [ ( 1)]w + ( 1)( 1) [ ]w + 1 (5.2)

and since we require [1 , 2 ] to be hyperbolic, then some calculation shows that we must have: +1 2 > (5.3) 1 Note that the right-hand side is always bigger than one for > 1, so then it covers the previous demand > 1. The three-dimensional space spanned by , and , restricted to the values for which the above inequalities hold, is a parametrization on the Teichmller space of Riemann surfaces of type (1, 0, 1). u 80

Of course, it is somewhat reminiscent of the Fricke coordinates discussed in chapter 2. However, in our case we do not cover the entire Teichmller space u since we no longer have the freedom to invert 2 . On the other hand, the moduli space remains completely covered so no essential generality is lost. PSfrag replacements

1 2 A B D A 0 C D B C

Figure 5.1: Fundamental region for a toroidal wormhole. The reader can verify that there is indeed one asymptotic region because of the identications; the path ABCDA is closed. In gure 5.1 we drew a fundamental region that is dierent from the one shown in chapter 1, but it will be more useful in the sequel. The unique periodic geodesic homotopic to the ideal boundary is split in four parts. It is important to note and not very dicult to verify that these four circle segments, so the segments AB, DA, CD and BC, are parts of the geodesics connecting the xed 1 1 1 1 points of [1 , 2 ], [1 , 2 ], [1 , 2 ] and [1 , 2 ], respectively. Clearly, these four elements of or their inverses are all conjugate in , consider for example
1 1 1 1 1 1 [1 , 2 ] = 2 2 [1 , 2 ]2 2 = 2 [2 , 1 ]2 = 2 [1 , 2 ]1 2 .

From this we also conclude that the geodesic connecting the xed points of 1 1 [1 , 2 ] (which contains the AB segment) is the image under 2 of the geodesic connecting the xed points of [1 , 2 ] (which contains the BC segment).

5.1.2

Spacetime

We are now ready to formally and completely dene the toroidal wormhole spacetime. Again, we will not distinguish between the elements of Aut(H) and their extension to isometries of AdS3 ; both are denoted for example . Consider the group of isometries of AdS3 generated by: 1 : (x, y, z) (x, y, z) and (x + 1)(x + ) + 2 (z 2 y 2 ) 1 y 2 : (x, y, z) (x + )2 + 2 (z 2 y 2 ) z with > 0, > 1 and >
+1 2 1 .

It is a group of isometries for

H 2+1 {p H 2+1 : d(p, p) > 0 ()}. p 81

with () the xed point set of on the x-axis. Furthermore, using theorem 4.3.1, acts properly discontinuously and xed point freely on H 2+1 and does not induce closed timelike curves. We thus are able to dene the spacetime: M 2+1 H 2+1 / (5.4)

which has at least one spatial slice (S H/) equal to a toroidal wormhole. This is a complete description of the spacetime. Indeed, every suitably small open domain in H 2+1 denes a chart on M 2+1 and M 2+1 is completely covered with these charts, which is required for a manifold. Furthermore, the canonical metric on H 2+1 is a well-dened metric on M 2+1 since we used isometries to take the quotient (5.4). In the next section, rst we will see that the surface S can be covered with two charts only. Afterwards, we extend this idea to cover (almost) the entire spacetime M 2+1 with two charts as well. The calculation of the location of these patches and dening suitable coordinates on them is the main subject of this chapter.

5.2

Patching the surface

We will see that the surface S can be almost completely covered by two annuli, glued together on an overlapping region and whose boundaries contain special points that are identied with each other as well. The procedure is roughly indicated in gure 5.2 but we will go through it step by step below. The calculations will be done after we explained how the patches will be dened. PSfrag replacements l2 III l1 I 1 Figure 5.2: Dening patches on the wormhole and on the fundamental region. With a slight abuse of language, we did not dierentiate between the points, lines and regions on S and their lifts to H in the right gure. In particular, the indicated lifts of the splitting line l1 are ow lines of the commutator [1 , 2 ] 1 1 1 (the rightmost circle) and, from right to left, of its images under 1 , 2 1 1 and 2 . A lift of the splitting lines l2 and l3 is also indicated; they are both ow lines of 1 . II p1 1 l3 III l1 l1 III II I l1 l2 III

l3

p1 II

II

5.2.1

Denition

From gure 5.2, we see that the boundaries of the patches will be closed curves. If we want these closed curves to be smooth, it is a good idea to take them 82

l1 III

l1 III

p1

II

to be ow lines of elements of , since these are smooth closed curves on S. Unfortunately we cannot use a single ow line that ensures that each of the two annuli has smooth boundary curves. This is why we introduced the overlapping region that is part of either annulus, indicated as region III in gure 5.2. The outer annulus then consists of the regions II and III and the inner annulus of the regions I and III. We stress that region III is just the overlap of these two patches and its denition is implicit in the denitions of only the two annuli, which are given below. Since we dene everything on S rather than on H, mostly the left-hand side of gure 5.2 is used for now. Outer annulus To nd the splitting lines we look for, we start at the ideal boundary that is connected to region II in gure 5.2. Here, we construct a foliation by taking the leaves to be the ow lines of the commutator [1 , 2 ] that generates a simple loop around the ideal boundary. One particular leaf of this foliation is of course the periodic geodesic associated with it, but in fact the foliation can be extended a little bit further away from the ideal boundary. At a certain point, however, the foliation becomes invalid: a circle that is part of the foliation touches itself on S. It thus splits into two closed curves, one around each leg of the torus, that have one point in common. We then stop the foliation and call this ow line the critical ow line l1 and the point where the critical ow line touches itself p1 , as indicated in gure 5.2. By denition, the regions II and III in the gure lie between this ow line and the ideal boundary. If we cut S along this ow line, these regions together form an annulus which we denote Sa . The annulus has two connected boundary components: at one side, there is the ideal boundary, which is innitely far away from any interior point. At the other side, at nite distance from the interior, there is l1 and for a complete description of this region we need to identify two points on l1 . Inner annulus For the inner annulus we follow a similar procedure but we now use ow lines of 1 . We then start the foliation in region I, for example at the closed periodic geodesic associated to 1 , supposing it generates the loop we indicated in gure 5.2. This time, we foliate in two directions and stop the foliation at a point where the ow lines meet again and touch each other. We will see below that the point where they touch can be chosen to be p1 as well. We call the two touching leaves of this foliation l2 and l3 . Note that whereas l1 is a smooth circle that touches itself, l2 and l3 are smooth circles that touch each other. The regions I and III also form an annulus, denoted Si , with both boundaries at nite distance this time and cut o along two ow lines of 1 . This is also indicated in gure 5.2. In the compactication of Si in S, we also need to identify two points on the boundary of Si but now there is one point at each boundary circle. Issues Is gure 5.2 always correct? In fact it is not. Namely, if we want to dene p1 and l1 as sketched above, there is quite a subtle point which we will need to 83

verify. Note that we drew gure 5.2 in such a way that 1 generates the loop we indicated, hence assuming that the rst foliation we constructed breaks down because it wraps around the loop generated by 1 and not any other element of . In other words, we assumed but did not prove that the two loops that l1 splits into are homotopic to the periodic geodesic associated with 1 and not to any other periodic geodesic. This ambiguity will be at least partially solved in section 5.3. First, however, we continue with the above assumption as though it is completely true and calculate the lift of the patches to H below. This we did, of course, because section 5.3 ensures that the calculation we are about to start remains generically valid (at least for a certain range of the parameters and ) so we are not undertaking pointless calculations. Likewise, a similar issue exists for l2 and l3 , for which we have to verify that each does not intersect itself by wrapping around the asymptotic region. This is resolved in a similar way in section 5.3.

5.2.2

The patches in H

The only description of S we have is as a quotient of H. Hence, if we want to do calculations on Sa and Si , it is necessary to lift them to H. In this section, we will rst lift the critical ow lines, from which the domains Ha and Hi (to be dened below) automatically follow. So what do critical ow lines as l1 , l2 and l3 look like on the universal covering surface? We drew them as part of a fundamental domain in gure 5.2, but there are innitely many other lifts of them as well. All lifts are, by denition, ow lines of a group element conjugate to [1 , 2 ] or to 1 . Since the ow lines of hyperbolic Mbius transformations are circles on P1 that go through the xed o points of the transformation, the lifts will look like circle segments in H. Lifts of the outer annulus Let us rst consider the lifts of l1 . Of the innitely many lifts of l1 , we single out the particular lift that is a ow line of [1 , 2 ]. We denote this line also l1 , see gure 5.3. Following a ow circle of [1 , 2 ] on H, [1 , 2 ] maps any point on the circle to a point that lies a bit further on the same circle. The circle segment between such a pair of points descends as exactly one complete closed loop on the quotient surface. For the critical ow line the same holds, but there is also another element of the covering group, which we assumed is 1 , that identies two points on this line that were not images under [1 , 2 ]. This is where, on the quotient surface, the ow line touches itself. The two points on this line that are identied under 1 are both lifts of p1 . So if the rst lift of p1 is at, say, w1 H, then another lift is at 1 (w1 ) = w1 and it lies on the same ow line of [1 , 2 ]. A third lift is obviously [1 , 2 ](w1 ). The circle segment between w1 and [1 , 2 ](w1 ), with the point w1 between them, is then a lift of the entire critical ow line, see gure 5.3. Of course, the points [1 , 2 ]m (w1 ) or [1 , 2 ]n (w1 ) or m, n Z are also lifts of p1 that lie on the same ow line of [1 , 2 ] in H, but they are not important for the calculation below. Equivalent to saying that two points on the ow line get identied by 1 is the statement that the image of the entire ow line under 1 intersects the original. In the case of the critical ow line, the line and its image just touch 84

each other. This is why lifts of the critical ow lines in H will form a series of circles that touch heir neighbours, all being the image under 1 of the circle to the left of it, as is indicated in gure 5.3. PSfrag replacements

[1 , 2 ]

w1

l1

wc

p /

p+ / p

p+

Figure 5.3: The calculation of the exact location of the critical ow lines in H. The points w1 , w1 and [1 , 2 ]w1 are all lifts of p1 . The xed points of [1 , 2 ] are indicated p and p+ . This observation allows us to actually calculate where the critical ow line of [1 , 2 ] lies in H, which we will do immediately. Denote the xed points of [1 , 2 ] by p , p+ R such that 0 < p < p+ . Then [1 , 2 ] leaves the circles parametrized by h in the following way: |w (p+ p )2 p+ + p ih|2 = h2 + 2 4 (5.5)

invariant for all h R. These are the ow lines we look for. The image of the ow lines given by (5.5) under 1 is simply: | (p+ p )2 w p+ + p ih|2 = h2 + . 2 4 (5.6)

These are circles through p and p+ . One calculates that the circles given by (5.5) and (5.6) have exactly one intersection point if: h2 = h 2 c 1 ( + 1)( 1) 2 (p+ p )(p p+ ) . 4 (5.7)

Now from (5.2), one calculates that the xed points of [1 , 2 ] are given by: p = ( 1)2 ( + 1)2 4( 1) . (5.8)

The factor under the square root is always positive, because this is exactly the requirement (5.3) that [1 , 2 ] be hyperbolic. Note also that with the given range of parameters p+ > p > 1. Substituting (5.8) in (5.7), we obtain: h2 = c ( 1)( + 1)2 . 4 2 85 (5.9)

wc /

w1 [1 , 2 ]w1 p

We are interested in the positive square root, hc > 0. If we substitute (5.9) and (5.8) in (5.5), we see that the lift of the critical ow line to a ow line of [1 , 2 ] is given by: l1 : |w i 1 ( + 1)( 1) ( + 1) 2 2 1|2 = ( 1)( 1)2 . 4 (5.10)

Now for the lift w1 of p1 . The critical circles touch at a point that lies on a ow line of 1 , which is a straight line through the origin. Because all transformations are angle preserving, a little thinking gives that this ow line of 1 goes through the center of each circle, so through the center points wc = (p+ + p )/2 + ihc and through wc , as indicated in gure 5.3. Now from this gure with a little further calculation, one deduces that the point w1 where two circles touch each other (and hence is a lift of p1 ) is given by: 2wc 1 w1 = (5.11) = ( 1 + i 1). +1 The point w1 lies on the critical ow line l1 of [1 , 2 ] and also on its image 1 1 1 (l1 ), which is a ow line of [1 , 2 ]. This is also shown in gure 5.3. Now we dene Ha to be the region of H bounded by l1 and the interval [p , p+ ] R, which after a little rewriting of (5.10) equals: Ha {w H : | 2 1 w ( 1) i 1|2 < ( 1) +1 +1
2

} (5.12)

and then formally dene the outer annulus by taking the quotient Sa = Ha / < [1 , 2 ] > . (5.13)

In the natural compactication, we should also identify the boundary points on Sa that lift to w1 and w1 . With these lifts of Sa , we can now proceed by nding lifts of Si in a similar way. Lifts of the inner annulus The inner annulus is in a similar manner lifted innitely many times to H. To nd it, one needs lifts of the ow lines l2 and l3 . We denote the particular lifts of l2 and l3 that are ow lines of 1 also by l2 and l3 . We rst dene l2 to be the ow line of 1 through the lift w1 of p1 . Since the ow lines of 1 are straight lines through the origin, it is easily seen to be given by arg w = arg w1 or, with (5.11): l2 : z = x . 1 (5.14)

This is l2 in gure 5.2 and it is also shown in gure 5.4. Now we search a ow line of 1 that has one common point on S with l2 and we furthermore want both points to be related in H by a 2 transformation. Equivalently, exactly as 1 we did with l1 , we need a ow line of 2 1 2 that touches l2 . 1 The ow lines of 2 1 2 are given by circles through 2 (0) = 1 1/ and 1 2 () = 1. There is only one ow line of 2 1 2 that touches l2 and this is 86

PSfrag replacements [1 , 2 ] l2

zc /

w1

2 (l3 ) p+ 2 (0) 0 p / p+ / 2 () p

Figure 5.4: Zooming in on gure 5.3. The calculation of the ow lines l2 and l3 in H.
1 the one we dene as 2 (l3 ), so that its image under 2 is by denition l3 which in turn is a ow line of 1 . In a similar way as with l1 and w1 , we calculate:

2 (l3 ) : |w

2 1 i 1 2 1 | = 2 2 4

(5.15)

and substituting (5.14) shows that l2 and 2 (l3 ) touch exactly at w1 . 1 Since 2 (l3 ) is the unique ow line of 2 1 2 through w1 , the line l3 is the 1 unique ow line of 1 through 2 (w1 ) and is thus given by:
1 l3 : arg w = arg 2 (w1 )

and since
1 2 (w1 ) =

1 (1 + i x 1

1)

(5.16)

one has l3 : z =

(5.17)

We did not expect in advance that l2 and 2 (l3 ) touch each other at w1 and invite the reader to come up with an intuitive explanation. Another indication that w1 , or rather p1 , is indeed a special point follows from the fact that it also lies on the periodic geodesic associated with 2 . If we now take the region of H between l2 and l3 , which by (5.14) and (5.17) is given by: (5.18) Hi {(x + iz) H : |x| < z 1} then the inner annulus can formally be dened as: Si = H i / < 1 > . (5.19)

A complete description identies the boundary points on Si that lift to w1 on 1 l2 and 2 (w1 ) on l3 . 87

Lifts of region III Having dened the annuli Sa and Si , we should also search for their overlap on S. This is, by denition, region III on S. As can be seen from gure 5.2, this region consists in fact of two pinched annuli that share the point where they are PSfrag replacements pinched, which is p1 again.

1 l2 III b l3
1 2 (l1 )

l1

III a [1 , 2 ]

1 1 (III b ) 1 0 1 (III a )

[1 , 2 ]2 (III b ) [1 , 2 ](III a )

2 (III b )

Figure 5.5: Regions in H. The part lying between l2 and l3 is Hi and descends to the inner annulus (regions I and III). The shaded domain bounded by l1 is Ha and descends to the outer annulus. The shading indicates whether points in the corresponding domain in H descend to region I, II or III of S, corresponding to gure 5.2. Note that both in Hi and Ha there are innitely many lifts of the regions IIIa and IIIb ; all are identically shaded. We are interested in the lifts of region III that lie within either one of the domains of Ha and Hi that we dened above. This is because we will dene new coordinates on the spacetime via the extension of Ha and Hi to H 2+1 : the various lifts of region III that lie within Ha and Hi will thus form exactly the domain of validity of the transition functions between the two coordinate systems. The intersection of Hi and Ha covers exactly a lift of one of the pinched annuli, the part of region III which we denote IIIa . This particular lift of IIIa is then III a : III a : Hi Ha . (5.20) Because l2 and l1 always intersect, namely at w1 and at w1 , III a is always nonempty. There is no need to take a quotient now, but the two points w1 and w1 on its boundary are to be identied. We chose to label the other part of region III as IIIb , perhaps somewhat unsurprisingly. It lifts, for example, to the domain that is bounded by l3 and 1 2 (l1 ). We write this particular lift as III b , see gure 5.5. We have: 1 2 (p ) = p
1 and combining this with (5.16), we see that the image of l1 under 2 is: 1 2 (l1 ) : |

2 1 w + 1 i 1|2 = ( 1) +1 +1 88

(5.21)

1 1 which is the only circle through the three points 2 (p ) and 2 (w1 ). Note the similarity with (5.10). We see that III b is the domain in Hi :

III b : {w Hi : |

1 2 w + 1 i 1|2 < ( 1) +1 +1

(5.22)

We leave it to the reader to show which boundary points should be identied. This domain is mapped by 2 to a part of Ha . Keeping (5.15) in mind, it is given by: 2 1 i 1 1 2 (III b ) : {w Ha : |w |< }. (5.23) 2 2 4 We now have constructed one lift of IIIa and IIIb in both Hi and Ha . As is sketched in gure 5.5, there are innitely many other lifts of the regions III in Hi and Ha . In Ha , they are the images of III a and 2 (III b ) under < [1 , 2 ] >; these touch each other at the images of w1 and w1 . In Hi , they are the images under < 1 > of either III a or III b . We leave this calculation to the reader.

5.2.3

A nal isometry

Although we are quite happy with the regions Ha , Hi , III a and III b we dened above (given by the equations (5.12), (5.18), (5.20) and (5.22), respectively), we need to go one step further to be fully prepared for the new coordinate systems we introduce in sections 5.5 and 5.6. This is because Ha is not very nice in the sense that it proves to be somewhat dicult to nd a good coordinate system on Sa directly via Ha . To dene new coordinates, it would be better if [1 , 2 ] were generated by some element w w. This can of course be done by applying an isometry that maps p and p+ to zero and innity. An example of such an isometry is: w + p (5.24) :w w p+ which is readily veried to have positive determinant and hence is an element of Aut(H). As expected, one has: [1 , 2 ] 1 : w w with = 1 ( 1)2 2 1 + ( 1) ( 1)2 ( + 1)2 4( 1) 42 2[( + 1) + i ( 1)( + 1)2 4]
2

Note that with the given range of parameters, > 1. Now under , we have: (w1 ) = ( 1)[( 1) ( + 1) + ( 1)2 ( + 1)2 4( 1)]

The crucial ow line l1 that we used to split the surface now is dened by the straight line: arg w = arg (w1 ) since this is the unique ow line of [1 , 2 ] 1 through (w1 ). Substitution gives: ( 1)( + 1)2 4 z . (5.25) (l1 ) : = x +1 89

As one might have guessed, the points (w1 ) and (w1 ) on (l1 ) are also to be identied, with the latter point given by: (w1 ) = 2[( + 1) + i ( 1)[( 1) + + 1 ( 1)( + 1)2 4] ( 1)2 ( + 1)2 4( 1)]

which from (5.25) clearly lies on (l1 ). From direct calculation one can obtain the following interesting result: (w1 ) = (w1 ) (5.26) so, in a sense, w1 lies halfway between w1 and [1 , 2 ](w1 ) on l1 . Of course, all images under [1 , 2 ] 1 , so all their images under multiplication by n , n Z, should also be taken into account. We conclude that maps Ha to the wedge (Ha ) = {w H : 0 < arg w < arg (w1 )} which notably contains the entire x > 0 quadrant. The outer annulus is now dened as: (Ha )/ < [1 , 2 ] 1 > (5.27) again with boundary points at the line arg w = arg (w1 ) appropriately identied. This ends our discussion on and its image of points, lines and domains. In the sequel, we will not need the image under of the lifts of the inner annulus nor of the regions III, so that calculation is left to the reader.

5.2.4

Summary

We briey pause to summarize what we have done and what needs to be done. The discussion of dening suitable patches on S and their lifts to H has been completed: we dened the patches Si and Sa that cover all of S but p1 and found explicitly a lift to H as well as the regions in H where they overlap. Finally, we dened the isometry such that the group element that generates the fundamental group of the outer annulus is a simple scaling. What clearly needs to be done now is to give a thorough discussion of the aforementioned issues, which the next section deals with. With these (at least partially) out of the way, we can extend all the patches to the entire spacetime M 2+1 , which is the subject of section 5.4. In our nal sections 5.5 and 5.6, we will nd the promised new coordinate systems and thus new forms of the hyperbolic metric on M 2+1 .

5.3

Lifting the issues

The annuli Sa and Si we dened above do not always correspond to parts of S. Basically, what we need is that there exists only one cyclic nontrivial subgroup of whose elements identify points within the regions Ha and Hi . These should be the groups generated by [1 , 2 ] for Ha and by 1 for Hi . However, this is only true if all the other nontrivial elements of map the entire region to a new, disjoint, domain in H, because then the quotient spaces like Ha / < [1 , 2 ] > are in one-to-one correspondence with a part of H/. We will now show when this is certainly true, but in doing so we might exclude certain parts of the moduli space. 90

5.3.1

Objective

We start with Ha . Recall that we dened a foliation on S that breaks down at a line l1 which splits into two loops that share a single common point. Translated to H, this meant that ow lines of elements conjugate to [1 , 2 ] start touching their images under some elements of . What we are after in this section are the possible a whose image of the particular ow line l1 , by denition a ow line of [1 , 2 ], just touches the original. In the most general case, a is some word of 1 s, 2 s and their inverses. However, not all elements of are eligible for a : for example, since [1 , 2 ](l1 ) = l1 , we can rule out all the 2 words in < [1 , 2 ] > and similarly if 1 (l1 ) touches l1 , it is clear that 1 (l1 ) will n intersect l1 so 1 for n = 1 can also be ruled out. In short, some words can and some words cannot be the critical element a . Note also that a is dened only up to inversion and left or right multiplication with [1 , 2 ]n for some n Z. This denes an obvious equivalence relation between a s which is denoted . The possible equivalence classes of a is what we are really interested in and these we will now deduce.

5.3.2

Procedure

With the problem claried, we can now describe our method for obtaining the a .

The surface S is topologically (C\L)/L We need to juggle a bit with homotopy classes of lines for our analysis. For this, it is convenient to picture S homeomorphically as C minus a lattice L modulo that same lattice, so (C\L)/L. This description is reminiscent of that of an ordinary torus but removing the lattice itself obviously yields a punctured torus, which is topologically (but not biholomorphically) equivalent to S. In fact, if we drop the holomorphicity requirement from denition 1.3.1, C\L is a covering surface of S, but since it is not simply connected, the covering is not universal. The covering group and the fundamental group are thus not isomorphic. Yet there still is a nice map of the fundamental group to this covering space. Note that every element of the covering group is a word of 1 s and 2 s and that we can map each word to a set of lattice translations from a particular point that ends somewhere at an image under L of the point or perhaps at the original point. Formally, then, given a base point p on C\L, there is a one-toone map of the fundamental group of S to the set of all homotopy classes of all paths between the base point p and its images L(p ), with the homotopy of course dened on C\L. An example is shown on the right hand side of gure 3 2 5.6, where the dotted lines connect p with 1 2 (p ) and the particular path 1 2 they follow is 1 2 1 2 (read from right to left). Conversely, from this example it is also clear that a word in the fundamental group of S denes a lattice translation by abelianizing it, so by setting [1 , 2 ] = Id. 91

PSfrag replacements 1 2
3 2 1 2 (p )

Figure 5.6: The foliation represented on C\L. We start with a small circle around every lattice point and blow it up until two circles touch. The foliation on (C\L)/L We recall that, in section 5.2.1, we dened a foliation starting at the ideal boundary of S that ended at a critical ow line which we called l1 . Now, on the covering surface C\L, our foliation boils down to drawing the same (topological) circle around every point in the lattice and blowing them up until a circle just touches one of its images on the lattice. This procedure is schematically outlined in gure 5.6. The touching circles are the lifts of l1 to C\L and their interior (shaded in gure 5.6) represents Sa . Consider now two such touching circles. First, they single out two points on n m the lattice that are related by a lattice translation dened by 1 2 with n and m necessarily relatively prime. Second, a new (topological) circle that envelops both circles but no other lattice points than the two inside the circles must be retractable (on C\L) into the straight line joining these two lattice points. If any of these conditions does not hold, it is easy to see that at least one of the two the circles must intersect some of their images on the lattice, which contradicts with the denition of the circles. Below, we will show that these two conditions determine all possible words that a can be. Relation to a To see how this all relates to a and l1 in H, we can restrict ourselves to a fundamental domain on C\L, because this can be mapped to a fundamental domain on H. If we choose the particular fundamental domain whose edges are identied by 1 and 2 , we can map it to the fundamental domain that was shown earlier in gure 5.1. The result is shown schematically in gure 5.7; we will later explain why the ow lines are so distorted. This distortion, on the other hand, is irrelevant for the discussion. We do need a continuous map, however; this is the case if the opposite edges of both fundamental domains are identied using the same group elements, so using 1 and 2 in the indicated case. We will only consider these fundamental domains below. Now from the fundamental domain in H it is easy to deduce a as follows. First, note how dierent images of l1 enter the fundamental domain 2 1 1 and how eventually 1 l1 touches 2 1 2 l1 . But this means that l1 touches 92

2 A 1 PSfrag replacements

1 1 l1

1 2

1 1 2 l1

1 2 l1 1 1 1 1 l1

l1 A

1 1 2 1 2 l1

Figure 5.7: A map between a fundamental region on C\L and one on H. We are not faithfully representing the ow line l1 or its images, since they are by denition circles in H. However, for the discussion only the continuity of the map is relevant. The point A in the upper right corner of the fundamental domain on C\L is mapped to the corner of the fundamental domain that is enveloped by [1 , 2 ], compare gure 5.1.

3 2 Note that a denes the lattice translation 1 2 which could already be read o from gure 5.6.

1 1 1 1 1 1 2 1 2 1 1 l1 and we are done: in this case, a 2 1 2 1 1 .

Generalization We have now found one possible a but from the above construction, we directly see how we can nd the others as well, which we do now. Consider the continuous invertible map between the fundamental domains in H and in C\L of gure 5.7. We immediately see that l1 in H (so the ow line of [1 , 2 ]) is always mapped to the circle part that is closest to the upper right corner of the fundamental domain in C\L, so the corner marked with an A in gure 5.7. We can then read o the equivalence classes of all a directly from the construction of S as (C\L)/L via the following recipe: rst we pick an arbitrary point p at this circle segment in the upper right corner of the fundamental domain we sketched. We then follow a path along two touching circles like the bold n m path in gure 5.6 until we reach 1 2 (p ) on C\L. Finally, the word dened by this path is exactly a . The reader can easily check that this procedure holds for 1 1 1 any possible a and not only for the sketched case where a 2 1 2 1 1 .

Now, we already stated above that the words we will obtain are always the n m approximation of a straight line between p and a 1 2 (p ) and that n and m are always relatively prime. In general, the approximation of the straight line 93

with [a, b] the number of integers lying strictly between a and b. The verication of this is left to the reader. We have now met the objective stated at the beginning of this section: the right hand side of this formula contains all the possible equivalence classes of a if we let n and m range over all relatively prime integers. Remark. We should note that if n and m are of equal sign, there is something nontrivial in the above construction. To see this, rotate the fundamental square in gure 5.7 90 degrees and dene l1 to be the circle segment that envelops the upper right corner: one readily veries that l1 consists of two parts. However, a more careful analysis shows that picking p on either component is ne; the dierence in the resulting a will be an extra [1 , 2 ] which does not change its equivalence class.

n m between p and 1 2 (p ) yields the words: n n [0, m ] [ m , 2n ] . . . [ (m1)n ,n] m m 1 2 1 2 1 2 2 1 (n1)m m m 2m [0, n ] [ n , n ] . . . [ n ,m] 2 1 2 1 2 1 1 2 a [0, |n| ] [ |n| , 2|n| ] [ (m1)|n| ,|n|] 1 m 2 1 m 2 1 m m 2 . . . 2 1 1 m m [ |n1|m ,m] 1 [0, |n| ] 1 [ |n| , 2m ] 1 |n| . . . 1 |n| 2 2 1 2 1 1 2 1

if n > m > 0 if m > n > 0 if n < 0 m > 0 |n| > |m| if n < 0 m > 0 |n| < |m|

5.3.3

Conclusion and a new condition

Recall that we parametrized the ow lines of [1 , 2 ] by h in equation (5.5). Now, since the equivalence classes of all possible a are parametrized by some relatively prime n and m, we in principle could nd the distance in H between a ow line lh of [1 , 2 ] and its image under all a = a (n, m). Our quest for the critical ow line then formally reduces to nding the smallest h > 0 for which the distance function d(h, n, m) = 0. Instead of working this out, we nd a way out by noting that if l1 and the periodic geodesic associated to 1 do not intersect on S, then a 1 necessarily. The problem then reduces to the non-intersection of l1 and this periodic geodesic, which is a lot easier. In H, we should demand that neither l1 nor all its images (l1 ) intersect with the z-axis. Luckily, as we show with the aid of gure 5.8, we only need to 1 check l1 and 2 (l1 ) instead of all (l1 ). This is because, on C\L, the periodic geodesic associated to 1 lifts to a vertical line as the one we indicated on the left of gure 5.8. We now note that we can always deform the fundamental domain we showed, if we keep the corners xed. In particular, we can deform the top and bottom horizontal edges (together), until either one of the following holds: The line starting in the upper right corner, so the one that corresponds to 1 l1 on H, intersects the vertical line, or its image 2 (l1 ) does so, or both do (perhaps multiple times, if we make really wiggly lines); Neither of the lines intersect the periodic geodesic and then a 1 . The reader can again verify this more general statement starting from the special case we show in the gure. Unfortunately, the converse of the second possibility, 94

A l1

A l1

PSfrag replacements

Figure 5.8: Deforming the fundamental domain on C\L. The vertical dashed line represents a lift to C\L of the periodic geodesic associated to 1 , which is mapped to the z-axis as in gure 5.7. By deforming it we see that in this case l1 intersects the geodesic at the indicated point; this statement is also valid in H. so the claim that if a 1 then the lines do not intersect, does not hold; we will have more to say about that later. This deformed fundamental domain can also be mapped to H again as a deformation of the fundamental domain in gure 5.7. And since everything that happens in H also happens in the fundamental domain, this means that
1 {l1 2 (l1 )} {x = 0} = (l1 ) {x = 0} = . 1 So we need to check only for l1 itself and 2 (l1 ) instead of all its images (l1 ). But from the equations (5.10) and (5.21), we see that if l1 does not intersect 1 the z-axis, then neither does 2 (l1 ). So in the end we only need to check that l1 does not intersect the z-axis. So when is l1 small enough that it does not intersect the z-axis? Using (5.10), this holds whenever:

1 ( + 1)( 1) > 2 which gives

( 1)( 1)2 4
2

This condition is not in conict with previous demands, but it means that we need at least > 2 for both (5.3) and (5.28) to be true. Remark. With > 2 and (5.3), we also deduce: > (1 + 2)2 5.8
2 and this is why gure 5.7 was drawn so distorted; if not, the circle 1 (l1 ) would be about 34 times smaller than l1 so the drawing would have been a lot less clear.

+1 < 1 1

(5.28)

95

So if (5.28) holds, we are sure that l1 is the right point to stop the foliation and that the Sa we dened descends to a part of S. On the other hand, if (5.28) does not hold, then two things can happen. First, a might not be equivalent to 1 . We then should nd a and we can continue as usual if we conjugate the entire group with an element of Aut(H) such that a is of the form z z and we call this 1 . We can then continue with new values and for which (5.28) does hold. Unfortunately, the particular a might be hard to nd; for one, since there are innitely many possibilities labeled by n and m. But the worst is yet to come: probably equation (5.28) might sometimes not hold but still a 1 , so we think it is a sucient but not a necessary condition (we have proven only the sucient part). This would be problematic, because (5.28) excludes the surfaces for which this happens, so it excludes a part of the moduli space. We yet have to nd a simple way around this issue, but for now we continue our analysis anyway and suppose everywhere that (5.28) is true.

5.3.4

The denition of Si

With a 1 , the construction of Si in the description (C\L)/L is a lot simpler. The only nontriviality here is to check whether the lines l2 and l3 on S do not wrap around the asymptotic region and intersect with themselves. Translated to a fundamental domain and then H, similar to what we did before, this means 1 1 that the images [1 , 2 ](l2 ) and [1 , 2 ](l3 ) do not intersect with l2 and with l3 , respectively. A straightforward calculation that we will not repeat here shows that this is always true with the range of parameters required above, including (5.28). It means, for example, that we were right in drawing the point [1 , 2 ](w1 ) below w1 in gure 5.3.

5.4

Patching the spacetime

With the issues out of the way, we now continue with the splitting of the spacetime by extending the splitting of S to M 2+1 . In the spacetime, the ow lines l1 , l2 and l3 we dened will become ow 2+1 surfaces that separate M 2+1 in two parts, which we call Ma and Mi2+1 . To obtain these patches, we will nd the extension of the ow lines l1 , l2 and l3 , pick two domains in H 2+1 as the extension of Hi and Ha and then take a quotient. In his short section, we will restrict ourselves to nding the correct denitions 2+1 of Ma and Mi2+1 ; the extension of the regions III a and III b and of is postponed until section 5.5. Outer annulus We rst look for an extension of the critical ow line l1 to a surface in H 2+1 . The circle segment l1 is a ow line of [1 , 2 ] that just touches its image under 1 . We search for a surface in H 2+1 with similar properties: it is invariant under [1 , 2 ] and the intersection with its image under 1 is a single line. This extension is also obtained via the primed coordinates y , z and w of section 4.2. Since l1 is given by (5.10), the extension of Ha is given by: | 2 w ( 1) i +1 1|2 < ( 1) 96 1 +1
2

meaning 1 : l 2x ( 1) +1
2

z 2 y2 +1

= ( 1)

1 +1

(5.29)

where we used to indicate the surface associated to l1 . The image under 1 l1 of this surface touches the original at the extension of the point w1 , which from (5.11) is the line: w1 : x = 1 1 z 2 y2 = . 2 (5.30)

Remark. In fact, following a similar reasoning as in section 4.5, we obtain that the entire region between the lightcones emanating from p and p+ and beyond |y| = z certainly never intersects with its image under any element of \ < [1 , 2 ] >. This again follows from the fact that the these lightcones are completely determined by p+ and p and the rigidity of the elements of that map lightcones to lightcones. This means that the issue about whether 1 is the right element of that stops the foliation on S, leading to the constraint (5.28), becomes irrelevant for |y| > z. 2+1 Anyway, the extension of Ha , which we call Ha , is thus the domain bounded by (5.29) 2x ( 1) +1
2

On the planes y = z, this point is pushed away to innity and the surface (5.29) is split into two straight lines given by x = p . However, these lines are excluded from the spacetime since they lie on lightcones emanating from these xed points. The part of spacetime beyond |y| = z that lies between these lightcones, so with p < x < p+ and |y| > z, never intersects with its image under 1 , since p+ < p . We thus completely include this part in the outer annulus.

z 2 y2 +1

< ( 1)

1 +1

(5.31)

for |y| z, together with the part of H 2+1 that it is connected with for |y| > z, so the domain between the lightcones from p and p+ , both for y > 0 and for y < 0. Then the outer part of H 2+1 is given by:
2+1 2+1 Ma = Ha / < [1 , 2 ] > .

Inner annulus Similarly, one can extend the ow lines l2 and l3 . In general, the extension via the new coordinate w of the ow lines of 1 in H are the lines: arg w = c with c a constant, 0 < c < , depending on which ow line we choose. In other words: z2 y2 = tan c x but this means that x 0 when y z, independent of c. The line x = 0, y = z is excluded from spacetime because 1 becomes lightlike here. We conclude 97

that the inner annulus lying between l2 and l3 disappears for |y| z and, from (5.18), that the extension of the inner annulus is thus the part of spacetime x Hi2+1 : z 2 y 2 > 1
2

(5.32)

modulo the action of a cyclic group that is now generated by the extension of 1 : Mi2+1 = Hi2+1 / < 1 > . With this time development of the two annuli, we are ready to dene some new coordinates on them.

5.5

Toy coordinates

With the denitions completed and the issues settled, we now turn our attention to new coordinate systems and new forms of the metric on M 2+1 . Of course, merely specifying a coordinate system is only part of the story. As we said in the introduction, our aim is to be able to drop the description of the spacetime as a quotient of H 2+1 entirely so we can use the new coordinates to completely describe the spacetime. We then not only need the form of the metric in the new coordinates, but also coordinate ranges and periodicities, as well as the transition functions between them on the regions III. Since this turns out to be a somewhat extensive story, this section discusses a simple new toy coordinate system on either patch and discusses the metric, coordinate ranges and the transition functions for them. The coordinates we will introduce below are simplied versions of the nal physical coordinate systems, which are the subject of section 5.6. We expect this two-step approach to clarify our motivation for choosing these coordinates. Before we start, we should note that the line w1 , given in (5.30), is not covered by any chart we dene below. However, we will ignore this issue since the line can be covered with a small extension of either chart to a suitably small open domain around it.

5.5.1

Inner annulus
exp(2) = x2 y 2 + z 2 y tanh(t) = z x tan(r) = 2 y2 z

Let us start by introducing new coordinates on the inner annulus. We dene:

with inverse transformation x = e sin(r) y = e cos(r) sinh(t) z = e cos(r) cosh(t). 98

(5.33)

The metric of H 2+1 in these coordinates becomes ds2 = dr2 + d2 dt2 + . cosh2 (t) cos2 (r) cosh2 (t)

The coordinate ranges can be found by rewriting (5.32), from which we see that the region Hi2+1 is given by: | tan(r)| < 1
1 2 (w1 ),

(5.34) given by the extension of

since on these surfaces one identies w1 and (5.11) and (5.16). These are the lines (t, r = arctan (t, r = arctan

1, =

1 ( 1) ln ) 2 1 ( 1) ) 1, = ln 2 2

so here the patch (even when restricted to a single period of ) starts overlapping itself and hence is no longer a valid chart. As for the periodicities, taking the quotient with respect to the group generated by 1 gives: (x, y, z) (x, y, z) which means (t, r, ) (t, r, + ln ).

This nishes our discussion the toy coordinates on the inner annulus; note how it is completely dened by the above formulas. We will now turn to the outer annulus, which is a bit more complicated.

5.5.2

Outer annulus

For the outside, we change coordinates after applying the isometry , so when [1 , 2 ] is of the form w w. We express the new coordinates in a new system of primed coordinates (x , y , z ) given by x x (x p )(x p+ ) + y 2 z 2 1 y = y = . y(p+ p ) (x p+ )2 y 2 + z 2 z z z(p+ p ) (5.35) Our coordinates for the outer annulus are given by: exp(2 ) = (x )2 (y )2 + (z )2 tanh(t ) = tan(r ) = The inverse transformation is: x = e cosh(t ) sin(r ) y = e sinh(t ) z = e cosh(t ) cos(r ) 99 x z y (x )2 + (z )2 . (5.36)

and the metric becomes in these coordinates: ds2 = (dr )2 (dt )2 (d )2 + . + 2 2 (r ) 2 (r ) cos cosh (t ) cos cosh (t ) cos2 (r )
2

The identication [1 , 2 ] 1 : (x, y, z) (x, y, z) becomes: (t , r , ) (t , r , + ln )


2+1 and the patch is valid in (Ha ), which is bounded by the extension of (5.25), so it is the domain where:

(z )2 (y )2 >

( 1)( + 1)2 4 +1

as one readily veries. On the bounding surface, once again, one identies pointwise the extension of (w1 ) and (w1 ). From our discussion in 5.2.3, notably equation (5.26), we see that they are given by the lines 1 + 1 cosh(t ) sin(r ) = = 0 1 1 + 1 1 cosh(t ) sin(r ) = = 0 + ln 2 1 with 0 ln ( 1) ( + 1) + 2 1 ( 1)( + 1)2 4 = ln[ (p+ 1)].

plus the part of H 2+1 with |y | > z , x > 0 and (x )2 (y )2 + (z )2 > 0. We recall that this was the part lying between the lightcones from p and p+ with 2+1 |y| > z as discussed below equation (5.31). In the new coordinates, (Ha ) is given by: 1 + 1 cosh(t ) sin(r ) > 1

(5.37) All images of these lines under [1 , 2 ] 1 should also be pointwise identied. As a sidenote, one may get an intuitive idea of the coordinate system by noting that the surfaces of constant t wrap around the excluded lightcones x2 y 2 + z 2 = 0 in H 2+1 for t . Also, t = 0 is simply the upper half plane.

5.5.3

Transition functions

The relation between the two coordinate systems is via the extension of the isometry , so via (5.24). In fact, the transition functions from the inner to the outer annulus is a series of coordinate transformations: t x x t r y y r z z

Where the transformations are given by (from left to right) the transformation (5.33), then or 2 on III a or III b , respectively, and then (5.36). We see 100

that the transition functions are well-dened coordinate transformations (on the designated regions) since each of the individual transformations is well-dened on the designated domain. We start with region III a , where the outer coordinates are written as a function of the inner coordinates via . Using the above relations, one has e cosh(t ) sin(r ) 1 = e sinh(t ) (e sin(r) p+ )2 + e2 cos2 (r) e cosh(t ) cos(r ) (5.38) (e sin(r) p )(e sin(r) p+ ) e2 cos2 (r) e cos(r) sinh(t)(p+ p ) e cos(r) cosh(t)(p+ p ) which gives us implicitly the transition functions on region III a . We give the explicit calculation only for the nal coordinates in the next section. Their domain is III a which, from (5.29), is now given by: 2e sin(r) ( 1) +1
2

2e cos(r) +1

< ( 1)

1 2 +1 . | tan(r)| < 1

We can rewrite it as ( + 1)e [( 1) sin(r) + 1 cos(r)] e2 > ( 1) | tan(r)| < 1 (5.39)

On region III b , on the other hand, the transition functions also involve an isometry 2 . Since, from (5.1) and (5.24), the composition 2 is given by: 2 : w it extends to H 2+1 as: x 1 2 : y [(1 p+ )x + (1 p+ ) 1]2 + 2 (1 p+ )2 (z 2 y 2 ) z [(p 1)x + (p 1) + 1][(1 p+ )x + (1 p+ ) 1] 2 (p 1)(p+ 1)(z 2 y 2 ) (p+ p )y (p+ p )z and the transition functions on III b can thus be dened as: e cosh(t ) sin(r ) 1 = e sinh(t ) 2 (1 p )2 e2 + [(1 p ) 1][2e sin(r) + (1 p ) 1] + + + e cosh(t ) cos(r ) (p 1)(1 p+ )[2 e2 + 2e sin(r) + 2 ] (p+ + p 2)[e sin(r) + ] 1 (p+ p )e cos(r) sinh(t) (p+ p )e cos(r) cosh(t) 101 (p 1)w + (p 1) + 1 (p+ 1)w (p+ 1) 1 (5.40)

This transformation is valid on the extension of region III b , given by the extension of equation (5.22), which in the new coordinates is: ( + 1)e [(1 ) sin(r) + 1 cos(r)] 2 e2 > ( 1) | tan(r)| < 1 . (5.41)

Note the dependence on and of the transition functions and their domains of validity. The transition functions on the other lifts of region III are found by rst mapping + m ln and + n ln , with m and n suitable integers such that either (5.39) or (5.41) holds.

5.6

New metric, old metric

We are now ready to come with nal forms of the metrics on the two patches. We will perform some more coordinate transformations in order to make the periodicities and identications somewhat more natural. We recall that our aim is to (be able to) forget about the description in H 2+1 entirely and use only the new coordinates we dene below: the spacetime is then completely described by the coordinate ranges, the metric and the transition functions. Below, we emphasized this by putting a box around these essential formulas. The remainder of the section is either motivation or calculation.

5.6.1

Inner annulus

First of all, the inner annulus is twisted in the sense that the periodic geodesic associated to 2 is not orthogonal to the one associated to 1 if = . In the metric, we can make this explicit by ensuring that both geodesics, at least at t = 0, are given by some coordinate xa = constant. This is the reason for introducing the following coordinates: e2(c+d) = (x aec+d )2 y 2 + z 2 meaning: x = ec+d 2 + b2 (1 a2 ) 2 + b2 + 2 + b2 (1 a2 ) y = ec+d sinh(t) 2 + b2 + 2 + b2 (1 a2 ) z = ec+d cosh(t) 2 + b2 + = bx z 2 y2 tanh t = y . z

(5.42)

with a, b, c, d R to be determined by some naturalness conditions for the bounding surface and the periodicities, see the next paragraphs. We obviously need |a| < 1 for these coordinates to be well-dened and then the Jacobian of the above coordinate transformation is given by c(z 2 y 2 )3/2 . This means that it is indeed a valid coordinate transformation for the inner annulus, which does not extend beyond |y| = z. Note also how the (t, , ) coordinates are almost equal to the toy coordinates (5.33) of the previous section if a = 0. 102

At t = 0, the z-axis is parametrized by = 0. Also in H, the periodic geodesic associated to 2 goes through its xed points at x= 1 2 ( + )2 4 .

which we, of course, obtained by setting 2 (x) = x in (5.1). Now we require this circle to correspond to = 0. This gives: a= ed = ( + )2 4 2

( + )2 4

Similarly, we require that the lines = bound the inner annulus. From (5.34) we see that we then should set: b= 1

Finally, the periodicity in should be 2, giving: c= ln . 2

We now introduce the physical constants: L1 = ln cosh( L2 + )= 2 2 cos = (5.43)

( + )2 4

and promise to explain their physical meaning below. We can then rewrite a, b, c, d as a function of L1 , L2 , : a = cos() b= L2 sinh( 2 ) sin() c= L1 2 ed = tanh( L2 ) 2 1 + cos() tanh( L2 ) 2

With these constants, the metric becomes: ds2 = L2 1 L2 d2 + 12 2 [sinh( )]2 2 + 1 d2 dt2 + 2 2 2 [sinh( L2 )]2 4 2 + cosh (t) 2 2 cos()L1 + dd 2 2 + 2 [sinh( L2 )]2 2

and coordinate ranges < t < , < < and 0 < 2, with the identications (t, , ) (t, , + 2) and on the bounding surface: (t, = , = 0) (t, = , = 0) (5.44)

The above three formulas capture all the information about the inner annulus. 103

Intermezzo Let us calculate the length of the periodic geodesics of the generators of in H. The one associated to 1 is easily veried to have length L1 = ln . The periodic geodesic associated to 2 is the straight line l2 , given by < < and t = = 0, whose length is:

ds =
l 2 =

d 2 + 2 [sinh( L2 )]2 2

= 2 sinh1 sinh(

L2 ) = L2 . 2

So we explained the motivation behind the introduction of L1 and L2 . It also is not hard to check where the third constant we introduced comes from: is precisely the angle between these two periodic geodesics. We can also use these lengths as a consistency check of the new metric. Since + L2 = 2 cosh1 ( ) 2 we indeed see that both lengths agree with (1.12) in chapter 1. Note that, for t > 0, the length of either loop decreases with increasing time, so an observer spending time in this particular region clearly is able to signal that his spacetime is collapsing in nite (proper) time. On the other hand, the advantage of being on this side is that there is no upper bound on the number of times he wants to go around his preferred loop. Another test would be that the metric on the inner annulus satises the Einstein equations (3.1). With the nonvanishing connection coecients: t = = = tanh(t) t tt t t = tanh(t) 2 + 2 [sinh( L2 )]2 2

t = t =

tanh(t)L2 2 L2 1 [sinh( )]2 2 + 1 2 4 2 tanh(t) cos()L1 2 + 2 [sinh( L2 )]2 2 2 +

= = = L2 1 2 4

2 [sinh( L2 )]2 2 L2 2 2 )] + 1 2 L1 cos()

2 [sinh(

= =

2 + 2 [sinh( L2 )]2 2

one indeed nds that Rab = 2gab .

5.6.2

Outer annulus

The new coordinates on the outer annulus are a bit easier. Starting from the toy coordinates (5.36), we dene: t t M r M tan(r ) 0 M (5.45)

104

To avoid confusion, we note that the above means that we for example dene a new coordinate t / M which we call t in the sequel. The mass is determined by: ln M= 2 which equals: L2 L1 cosh( M ) = 2 sinh2 ( ) sinh2 ( ) sin2 () 1. 2 2 The requirement that the the right hand side be bigger than one is, of course, exactly the condition (5.3). One has: ds2 = ((r )2 + M )(dt )2 (dr )2 ((r )2 + M )d 2 + + (r )2 + M cosh2 ( M t ) cosh2 ( Mt )

with standard periodicity in : (t , r , ) (t , r , + 2) and the patch is restricted to r cosh( M t ) M + (r )2 1 + 1 = > 1 cosh( L1 ) 2
1 2

cosh2 ( L1 ) + 2

cosh( M) 1

(5.46)

since on the bounding surface which we identify the lines given by cosh( L1 ) r cosh( M t ) 2 =0 = M + (r )2 cosh2 ( L1 ) + 1 cosh( M) 1 2 2 cosh( L1 ) r cosh( M t ) 2 = = . 2 L1 1 M + (r )2 cosh ( 2 ) + 2 cosh( M ) 1 Old metric From the transformation 2 = (r )2 + M cosh2 ( Mt ) tanh( M ) = (r )2 + M 2 tanh( M t ) r

one deduces that the metric takes the familiar form: ds2 = (2 M )d 2 + 2 d2 + 2 d 2 M

of a spinless BTZ black hole with mass M . The horizon at = given by r = M | sinh( Mt )| 105

M is also

2+1 and we emphasize that our coordinates and thus Ha always extend beyond it. In fact, the distance at t = 0 between the boundary of our patch and the horizon is the distance between r = 0 and (r )2 = (rc )2 which equals
rc r =0

2M cosh2 ( L1 ) 2 cosh( M ) 1

dr (r )2 +M

= sinh1 (rc / M) = sinh1

2 cosh( 2

L1

cosh( M )1

We note that the backwards lightcones of points on the bounding surface (5.46) always intersect the conformal boundary r = at t < 0. In other words, an observer entering the spacetime from r = at t > 0 will not be able to 2+1 leave the patch Ma and cannot actively probe the nontrivial topology of the inner annulus. In the (x , y , z ) coordinates, this is trivial: the bounding surface never intersects the lightsheet (actually lightcone in AdS3 ) y = z in H 2+1 . Remark. The new coordinates on the outer annulus can also be used to cover the entire BTZ spacetime with a single patch. Indeed, suppose the spacetime were BTZ, so the part of H 2+1 with (x )2 (y )2 +(z )2 > 0 modulo the identication (x , y , z ) (x , y , z ). Our coordinates clearly cover this region and also has the right periodicity, compatible with the identication. There are then two boundary components at r and their associated horizons are given by x = y, which equals: r = M| sinh( Mt )| In the toroidal wormhole, the boundary component with r is not present; instead it is replaced by the inner annulus. The horizon with r = M| sinh( Mt )| is thus also absent.

5.6.3

Transition functions

The transition functions follow directly from the denitions of the coordinates above. We will give them as functions of the inner (unprimed) coordinates to the outer (primed) coordinates. As we noted already when we discussed the toy coordinates, the transition functions are essentially given by the extension of on region III a and by the extension of 2 on III b . These transformations in turn are elements of P SL(2, R) extended to H 2+1 via the familiar (4.2). With this in mind, let us rst do some preliminary work. We dene: + 2 + b2 (1 a2 ) f (, ) ec 2 + b2 which equals
L1

f (, ) = e 2

2 + 2 [sinh( L2 )]2 2

2 + 2 [sinh( L2 ) sin()]2 2

We isolated ed for later convenience. 106

Second, we can rewrite a Aut(H) for which a, c, = 0 as: : w A w p 1 w p 2

with real A , p , p and A (p p ) > 0. Its extension to H 2+1 then becomes, 1 2 1 2 in the coordinates on the inner annulus: ed f (, ) h12 (, ) A sinh(t) : sinh(t)ed f (, ) h22 (, ) d cosh(t) cosh(t)e f (, ) with: h = (f (, ) ed p )(f (, ) ed p ) + f 2 (, ) ed (p p )f (, ) 1 2

and , {1, 2}. We will use h below. Note that h = h and h h 11 22 21 12 11 (h )2 = 1, so in fact we only have two independent functions. 12 Now we can go back to the transition functions. As we said above, the primed and unprimed coordinates are related by on III a and by 2 on III b . This means: cosh( Mt ) Mr+r 2 h12 (, ) A sinh(t) sinh( M t ) e M +0 = h22 (, ) M cosh(t) cosh( Mt ) M +r 2 on III a ; with the replacement 2 they are valid on III b . Solving for t , r and , we get: h11 (, ) h22 (, ) h12 (, ) r = cosh(t) M sinh(t) sinh( M t ) = h11 (, )h22 (, ) e2 = A2 e20 where the use of A and h on III a and A2 and h2 on III b is intended. We will now determine the parameters (A )2 e20 and ed p and ed p 1 2 and their counterparts (A2 )2 e20 and ed p2 and ed p2 as a function of 1 2 our physical parameters L1 , L2 and . From (5.24) and (5.40) we can read o: A = 1 p = p 1 p = p + 2 A2 = p2 1 p2 2 p 1 p+ 1 1 1 = + p 1 1 1 = + p+ 1
M

Using (p+ 1)(p 1) = 1 , a result following from direct calculation, we also nd: p2 = p p2 = p+ 2 1 107

Now we rewrite p and 0 in terms of L1 , L2 and , using (5.8) and (5.37). We obtain: p = e
L1 2

sinh( L2 ) sin() 2

1+

sinh2 ( L2 ) sin2 () 2

cosh(

L1 L2 ) sinh( ) sin() 2 2

sinh2 (

L1 L2 ) sinh2 ( ) sin2 () 1 . 2 2

We can also write p as a function of M and L1 only: L1 L1 e2 cosh M + cosh( ) p = 2 2 cosh(L1 ) + cosh( M) Also, e
0
L1

p+ 1 = p 1

M + e L1 e M + eL1

from which we obtain that (A )2 e20 = [(A2 )2 e20 ]1 . We will now put everything together and in the process clean up our notation a little bit. We introduce an upper index i {a, b} to indicate the region (III a or III b ) where the transition functions are valid. Also, we will redene our constants: for example, we dene pb ed p 2 and Aa e0 A . With this new notation, we see that the functions: hi = (f (, ) pi )(f (, ) pi ) + f 2 (, ) (pi pi )f (, ) 1 2

with f (, ) dened as above, become the building blocks of the transition functions on III i : hi (, ) 11 h22 (, ) hi (, ) r = 12 cosh(t) M sinh(t) sinh( M t ) = i (, )hi (, ) h11 22 e2
M

= (Ai )2

with the constants: (A ) = pa 1 pa 2


a 2 M + e L1 e M + eL1

[eL1 + e M ][e M + 1] = sinh2 ( L1 ) sinh(L2 )[1 cos() tanh( L2 )] 2 2 108

[eL1 + e M ][e M + 1] = sinh2 ( L1 ) sinh(L2 )[1 cos() tanh( L2 )] 2 2


and (A ) = pb 1 pb 2
b 2

M + e L1 e M + e L1

[eL1 + e M ][e M + 1] = sinh2 ( L1 ) sinh(L2 )[1 + cos() tanh( L2 )] 2 2 [eL1 + e M ][e M + 1] = sinh2 ( L1 ) sinh(L2 )[1 + cos() tanh( L2 )] 2 2

We hope to have claried how this form of the transition functions comes about; note also the similarities between the six constants above. Our last task is to dene the domain of the transition functions. The domain III a is given by Hi Ha , so we write (5.46), which is the same inequality as (5.31), in the new coordinates on the inner annulus: 2f (, ) ( 1) +1 which can be rewritten as: ha (, ) 12 > 2 cosh( L1 ) 2 cosh( M ) 1
2

2f (, ) +1

< ( 1)

1 +1

where we remark that we used (5.6) and (5.7) and the right hand side is also equal to 2hc /(p p+ ). Similarly III b , which is given by (5.22): 2f (, ) ( 1) +1 can be written as: hb (, ) 12 >
2

2f (, ) +1

< ( 1)

1 +1

2 cosh( L1 ) 2 cosh( M ) 1

Note how these inequalities indeed merely re-expres the inequality (5.46), as they should, which can be seen trivially for t = 0 from the transformation functions we just dened.

5.7

Discussion

In this chapter, we dened a spacetime M 2+1 , which has a spatial slice equal to a Riemann surface S of type (1, 0, 1). Afterwards, we covered S with two annuli and subsequently covered M 2+1 with two patches with both the topology of an annulus times R. We nished by nding an explicit, physical metric on both patches, their coordinate ranges and the relevant transition functions: all the information is captured in the boxed equations of the previous section. In this nal section, we draw some conclusions from our treatment of this threedimensional spacetime. 109

First, note again the equality with the BTZ black hole of the outer part 2+1 Ma that we already discussed in the previous chapter. Indeed, if we were to cut the extended BTZ black hole spacetime with mass M along the surface 2+1 given by (5.46), the resulting spacetime would be Ma . Again, we conclude that an observer moving outside of the horizons cannot actively probe whether her spacetime is a BTZ black hole or a toroidal wormhole spacetime. Of course, from the reasoning of section 4.5, it is clear that this holds for any surface that is glued to the outer annulus. As a sidenote, we would like to emphasize that there are no global Killing vectors elds in this spacetime; such a Killing vector eld would lift to a Killing vector eld on AdS3 and should commute with at least the Killing vector elds that generate 1 and 2 . However, such an element does not exist in sl(2, R). There are three obvious next steps to identify. First, we should completely lift the ambiguity in which a stops the foliation in Ha in order to be able to completely describe all the toroidal wormhole spacetimes in this way. Second, especially once this issue is solved, it would be interesting to apply the same method to other spacetimes. We can probably always split a wormhole spacetime in the same annulus-type components and therefore cover them with a couple of charts very similar to the ones we dened for the toroidal wormhole. We eventually may even prove that the transition functions can be completely determined from the metric in two overlapping patches. This would be a nice result, since it implies that one can specify the complete structure of the spacetime by specifying only which charts overlap and the form of the metric in all charts. Third, we can look beyond the scope of general relativity. As an example, I am quite interested in what a QFT-like computation of the radiation spectrum of this black hole would yield. Also, I am curious how the geometry behind the horizon can be encoded by a dual conformal eld theory. Because of the complete equivalence of the metric to a BTZ black hole on the outer region, this seems to require more than just gravity and probably string theory oers a way to encode this part of the spacetime in a dual CFT. On the other hand, it could also be sucient just to Euclideanize the spacetimes and nd a conformal eld theoretic description of the obtained Euclidean three-manifolds. This latter possibility is the subject of the next chapter.

110

Chapter 6

Hyperbolic three-manifolds
For certain spacetimes we can analytically continue a time coordinate t to imaginary values. This means that we apply the coordinate transformation t = i. The advantage of this transformation is that the metric becomes positive denite: ds2 = (. . .)dt2 + . . . = (. . .)d 2 + . . . so its signature becomes (+++). This procedure has led to interesting insights, notably the topology change that occurs for black holes and that indicates their thermodynamical properties. Also, solutions to the Einstein equations with Euclidean signature are probably interesting as instantons in a quantumgravitational theory. This is why this chapter is dedicated to nding such an analytic continuation for the wormhole spacetimes we discussed previously. We furthermore use the opportunity to also treat a few aspects of Euclidean signature gravity. For simple (i.e. symmetric) spacetimes, like for example a Schwarzschild black hole, the above procedure is well-known and textbook material. On the other hand, no general procedure is known for arbitrary, more complicated, spacetimes [44]. In particular, we do not generally know which time coordinate we should analytically continue. In our three-dimensional case, the BTZ black hole is such a symmetric spacetime: its analytic continuation is very similar to that of a Schwarzschild (or rather Kerr) black hole. We treat the BTZ black hole in section 6.2 as a rst example and will indeed encounter the familiar periodicity in Euclidean time. The other wormholes are more complicated. There exists, however, a procedure to obtain a Euclidean signature counterpart to these as well, which was proposed by Krasnov in [26] and extended in [27, 28]. We discuss this procedure in section 6.3 and will also extend Krasnovs results a bit. Section 6.4 then applies this procedure to our wormholes. The result will be that we obtain so-called handlebodies as their Euclidean counterparts. Roughly, a handlebody is a lled closed Riemann surface but we will see below what we mean exactly by it. Although the new procedure works for the static wormholes (by which we mean quotients of some AdS3 with respect to groups of isometries that leave 111

the plane y = 0 invariant), we are unsure whether it remains a valid procedure in the general case of rotating wormholes (quotients without such an invariant plane), discussed in for example [12, 9, 10]. In section 6.5, we discuss which problems arise and what the open questions are. We do not give a complete proof nor a falsication of the validity of the procedure in these cases. We start, however, with some mathematical preliminaries.

6.1

Denitions

This section introduces rst the Euclidean counterpart of AdS3 , which we call H 3 . As we did in chapter 3 for AdS3 , we give a denition, the metric and the isometries of the manifold. Afterwards, we introduce Schottky groups and Schottky spaces, that we will see describe our Euclidean wormholes. Schottky groups are similar to the Fuchsian groups of chapter 1 in that they also can be used to uniformize Riemann surfaces. We will also show how Schottky spaces are for Schottky groups what the Teichmller spaces of chapter 2 are for Fuchsian groups. As u one might expect, Schottky and Teichmller spaces are related and we end this u section with a brief explanation of this relation.

6.1.1

Hyperbolic three-space

As the Euclidean counterpart of AdS3 , we dene hyperbolic three-space to be one of the sheets of the two-sheeted hyperboloid in R3,1 dened as: U 2 + V 2 + X 2 + Y 2 = 1. The metric ds2 = dU 2 + dV 2 + dX 2 + dY 2 induces the metric ds2 = 1 (dx2 + dy 2 + dz 2 ) z2 (6.1)

on the hyperboloid if we apply the coordinate transformation: x= Y U +X y= V U +X z= 1 . U +X

which is in fact equal to the Lorentzian case, equation (3.10). We pick the sheet with U 1 if we require z > 0 so U + X > 0. The scalar curvature of the metric (6.1) is also everywhere constant and equal to 6. Unlike H 2+1 , H 3 is geodesically complete; its conformal boundary is the (x, y)-plane with a point at innity, so P1 . From the conformal compactication, we have the natural conformal structure dx2 + dy 2 on the boundary and hence an associated complex structure. Below, we use the complex coordinate w = x + iy. The Killing vectors of H 3 are: x y xx + yy + zz yx + xy (x2 y 2 z 2 )x + 2xyy + 2xzz 2xyx (x2 y 2 + z 2 )y + 2yzz . 112 (6.2)

and they generate Isom0 (H 3 ). This component of the isometry group is in fact isomorphic to Mb (so to P SL(2, C)), which can be shown as follows. If we set o z = 0 and complexify the Killing vectors, then they become: w ww w 2 w

plus their complex conjugates. These are easily integrated to obtain the Mbius o transformations (1.6). So the isometries of H 3 in Isom0 (H 3 ) extend to Mbius o transformations on the conformal boundary. Furthermore, the correspondence is one-to-one, since the Killing vector elds, once xed on the boundary, have a unique extension to H 3 . So Isom0 (H 3 ) and Mb are indeed isomorphic. o The full form of the isometries as dieomorphisms H 3 H 3 can be given in terms of the quaternions, see [34]. We will not bother ourselves with their explicit form but work almost always on the conformal boundary if we discuss isometries. We can again identify the manifold with a space of matrices, this time with the Hermitian unit determinant matrices: U X Y iV U X Y iV Y + iV U +X (6.3)

The action of w P SL(2, C) on it is given by: w : Y + iV U +X w U X Y iV Y + iV U +X


w .

(6.4)

where the right hand side is still a Hermitian unit determinant matrix. Quotients of H 3 Just like theorem 1.3.9 showed the uniqueness of simply connected Riemann surfaces, we know that H 3 is the unique connected and simply connected complete three-manifold with a maximally symmetric metric of everywhere constant negative curvature [43]. This implies that the universal covering of every complete three-manifold with a Riemannian metric satisfying the vacuum Einstein equations with = 1 is also exactly H 3 . We call these manifolds hyperbolic three-manifolds and they have been studied extensively by mathematicians, see again [43]. So all non-singular solutions to the Einstein equations in this setting are quotients of H 3 and hence are completely determined by their covering group, so by a subgroup of Isom0 (H 3 ). It is thus natural to take a closer look at the possible covering groups. First, the covering groups act properly discontinuously and xed-point freely on H 3 . This means that the nontrivial elements of the covering group are parabolic or loxodromic Mbius transformations, since elliptic Mbius transforo o mations have xed points inside H 3 . Second, all Mbius transformations have xed points on P1 . As for the o Fuchsian groups, we call the limit set () P1 of a covering group the set of points that are xed by a nontrivial element of . It can happen that () is P1 entirely, for example if H 3 / is compact. If this is not the case, we can take its complement in P1 () P1 \() 113

which is of course nonempty and then is called a Kleinian group. If does not contain elliptic elements then it acts properly discontinuously and xed-point freely on (), these are the torsion-free Kleinian groups. So for a torsion-free Kleinian group, we can take the quotient ()/ and this quotient is generally a set of disconnected Riemann surfaces (remember that we have a complex structure on P1 and the covering transformations are biholomorphic so the complex structure descends to the quotient). Now if () is connected, then its quotient is only a single Riemann surface, which is compact if does not contain parabolic elements, i.e. is purely loxodromic. These particular Kleinian groups are called Schottky groups. We discuss them in the next section while only considering their action on P1 , but our motivation is of course that they extend to H 3 uniquely and that H 3 / is a complete hyperbolic three-manifold with the Riemann surface ()/ as its sole boundary component. We henceforth call these manifolds handlebodies and will consider some examples in section 6.4.

6.1.2

Schottky uniformization

In chapter 1, we mentioned that all Riemann surfaces can be obtained as quotients of one of the three simply connected Riemann surfaces. We now briey discuss a dierent uniformization procedure of the compact Riemann surfaces, called Schottky uniformization. Consider a purely loxodromic subgroup of Mb, freely generated by g generators. Then is a Schottky group if () is o nonempty and connected. The quotient S ()/ is then a compact Riemann surface. It has genus g, which can be most easily seen from a fundamental region which consists of a sphere minus 2g disks; the elements of glue the bounding circles pairwise together. Note that (if g > 0) neither () nor the fundamental region are simply connected, so the covering of S by () is not universal and is not isomorphic to 1 (S). Example 6.1.1 (Schottky group of genus 1). The rst nontrivial example is the group which is generated by: w w this time for some C with || = 1. The limit set of this group is the origin and the point at innity so in this case () = P1 \{0, } = C . It is easily veried that acts properly discontinuously on C , so the quotient C / is a Riemann surface. The two-holed sphere: {w P1 : 1 < |w| < ||} is a fundamental region in P1 ; gluing the edges together gives us a torus. 114

The retrosection theorem (see for example [11] for an introduction) guarantees that every closed Riemann surface can be obtained as a quotient of a part of P1 with respect to a Schottky group and conversely that every marked closed Riemann surface determines in a unique way a Schottky group up to conjugation. So we could give plenty of other examples, but in fact we already mentioned a few, namely in example 1.5.11 where we discussed the Schottky doubles of the Riemann surfaces uniformized in H. In these case, is just a Fuchsian subgroup of P SL(2, R) which trivially maps to a subgroup of P SL(2, C), so of Aut(P1 ). This Fuchsian group is a Schottky group only if it uniformizes a Riemann surface with ideal boundaries (so that () is connected) but without punctures (so that it is purely loxodromic). We now introduce some denitions and, assuming familiarity with chapter 2, we keep the discussion very brief. A marking on a Schottky group is a choice of a sequence 1 , . . . , g that freely generate the group. A marked Schottky group is a Schottky group plus a marking and two marked Schottky groups and are equivalent if i = i 1 for i = 1 . . . g and some Mb. The o Schottky space of genus g is the set of all equivalence classes of marked Schottky groups of genus g. Note that the marking of a Schottky group denes only a partial marking of S, in contrast with the marking needed for the Teichmller u space. For example, for the torus, we have only specied one generator and not two. Indeed, with this weaker equivalence relation of equivalent markings, the Schottky space turns out to be a quotient of the Teichmller space [11]. It is, u however, not equal to the moduli space; the latter is even smaller (in the sense that multiple points in Schottky space correspond to the same point in moduli space). The Schottky space of genus 1 Let us again take the torus as an example. Note rst that we do not need to choose a marking for the Schottky groups that yield tori since there is only one equivalence class of markings possible, which is simply the sole generator of the group. Using conjugation freedom, we can always pick this generator to be A : w w with C and || < 1 (since the generator and its inverse are conjugate Mbius transformations). This, in fact, exhausts the conjugation freedom and o hence the Schottky space of genus 1 is the punctured disk. It is interesting to relate this to the Teichmller space that we found in u chapter 2 to be the upper half plane. There, we dened a marking on the torus by choosing two generators A and B . The Teichmller space of tori was then u obtained by dening A to be of the form w w + 1 and B to be anything like w w + with H. Afterwards, the action of P SL(2, Z) on the Teichmller u space was related to relabeling and interchanging the generators and we saw that the moduli space of the torus was the quotient H/P SL(2, Z). Now the marking of the Schottky group, so the sole generator, can be used to dene a partial marking on the torus. In particular, we can always make sure that the Schottky group generator equals A : w w + 1. But then we cannot perform an arbitrary relabeling (so an arbitrary P SL(2, Z) transformation) since we just xed A to be our Schottky generator. The only freedom we have left is 115

to interchange B with B + nA for some n Z. So, for genus 1, the Schottky space is the quotient of the Teichmller space with respect to + 1. Of u course, this agrees with what we just obtained: this quotient of H is exactly the way we uniformized the punctured disk in example 1.5.3.

For higher genus, there is a similar construction and the Schottky space is always the quotient of the Teichmller space with respect to a subgroup of the mapping u class group (and never with respect to the mapping class group entirely). Since we do not need it, we will not work this out here and refer the reader to [11].

6.2

The Euclidean BTZ black hole

This section introduces the canonical formalism for an analytic continuation of a black hole spacetime, applied to the case of a rotating BTZ black hole. The rotation is now incorporated because we will use this later and we want to remain as general as possible. The results of this section are well-known and can be found in for example [16], where the Euclidean path integral is also approximated and the expected thermodynamic behavior is found. A higherdimensional case with < 0 is discussed in [22]. First, consider the region outside the horizon with the metric (4.12): ds2 = r2 M + dr2 J2 dt2 + 4r2 r2 M +
J2 4r 2

+ r2 d +

J dt 2r2

where we relabeled the and coordinates, which is more in line with existing literature. Coordinate ranges are < t < , r+ < r < and 0 < 2 with periodicity (t, r, ) (t, r, + 2). Now we suppose t is complex, so t = i. To make the metric real, we then also need: J = iJ E . One obtains: ds2 = r2 M We introduce (J E )2 dr2 d 2 + 2 4r r2 M
E (r )2 = (J E )2 4r 2

+ r2 d

JE d 2r2

(6.5)

(J E )2 1/2 M 1+ 1+ 2 M2 which are the Euclidean counterparts of r , dened in terms of M and J in (4.11), which was where the horizons were in the Lorentzian spacetime. Note E E that the metric (6.5) becomes singular at r 2 = (r+ )2 and at r2 = (r )2 , so the latter point is never attained. We have:
E E M = (r+ )2 (r )2 E E J E = 2r r+

116

In the Lorentzian case, for any real value of r , we had |J| < M . This is not automatically satised in the analytically continued spacetime. If we impose it, we obtain: E E |J E | < M r+ > (1 + 2)|r |. Below, we drop the Es in the superscripts. Our conventions are a little dierent E from [16], where r has an extra factor of i and thus is imaginary. The metric (6.5) becomes singular at r = r+ . Since the horizons are smooth surfaces in the Lorentzian case and (locally) nothing happens on it, it is natural to demand smoothness of the corresponding Euclidean space as well. We thus seek a suitable coordinate transformation that avoids the singularity. We rst dene new coordinates similar to (4.10): = r+ + r The metric becomes: ds2 = r 2 d
2

= r+ r

r2 =

2 r 2 r+ 2 + r2 . r+

(6.6)

dr 2 + (1 + r 2 )d 2 1+r2

which is nonsingular at r = r+ , so at r = 0, if has periodicity 2, so we need ( , r , ) ( + 2, r , ). This implies: (, r, ) ( + 2r 2r+ 2 , r , + r 2 + r 2 ). 2 r+ + r + (6.7)

We can apply a second coordinate transformation: y x 1 r 2 = 2 (x2 + y 2 ) z exp(2 ) = x2 + y 2 + z 2 tan( ) =

(6.8)

which transforms the metric into (6.1), which is clearly everywhere regular, in particular at r = r+ , which is now the z-axis. In H 3 , the demand: (, r, ) (, r, + 2) becomes the identication (x, y, z) e2r+ x cos(2r ) + y sin(2r ), y cos(2r ) x sin(2r ), z which for J = 0, so for r = 0, reduces to the familiar: (x, y, z) e2
M

(6.9)

(x, y, z).

and which has an added rotation of angle 2r in the (x, y) plane if J = 0. The Euclidean BTZ black hole is thus the quotient of H 3 with respect to the identication (6.9). A fundamental region is bounded by two hemispheres 117

Figure 6.1: The analytically continued BTZ spacetime is a solid torus, which we indicated schematically as a fundamental region in H 3 (left) and as a whole on the right for J = 0. The shaded surface is the annulus given by = t = 0 and the dark line is the horizon, so the periodic geodesic on the annulus. of constant x2 + y 2 + z 2 , that are to glued together after a rotation of angle 2r . The resulting geometry is that of a solid torus, so D 2 S 1 (here D2 is the disk), with the horizon x = y = 0 lying at its center. This fundamental region and the resulting solid torus are sketched in gure 6.1. For J = 0, the slice t = 0 is invariant under the analytic continuation which is still the familiar annulus; its embedding in the torus is also shown in the gure. The boundary torus is Schottky uniformized on P1 : comparing (6.9) and example 6.1.1, we see that = exp(2[r+ + ir ]). This also implies that the fundamental groups of the Euclidean and the Lorentzian manifold are isomorphic: both have cyclic fundamental group. For the solid torus, this is because one homotopy class of nontrivial loops on the boundary is contractible inside the torus. We stop our analysis of the BTZ black hole here and refer again to [16] for details about the Euclidean path integral calculation. Their conclusions are the familiar properties that the black hole radiates at a temperature inversely proportional to the periodicity in Euclidean Killing time. This periodicity is seen to be, from (6.7), to be equal to: T 1 = 2r+ 2 2 . r+ + r 2r+ 4 G

With this result, the BTZ black hole has an entropy S = 4r+ =

equal to one quarter of the horizon area or rather the horizon length in Planck units (remember that we use G = 1/8). Ambiguities There unfortunately are two ambiguities considering the analytic continuation of the BTZ black hole. We only mention them as mathematical possibilities in the above construction and we do not know whether they are physically relevant in any way. 118

First of all, the reader might have noticed that the identication (6.9) is invariant under r r + k (6.10) for k Z. This means that we obtain the same Euclidean manifold for a number of dierent BTZ black holes [16]. Second, it is sometimes argued that (6.7) is the unique identication to make if one wants to avoid a conical singularity at the horizon in the Euclidean space. However, this is not true in three dimensions and in fact easy to see. Intuitively, we can mix the identications (6.7) and the identication + 2 to obtain a new identication which gives a smaller manifold. Formally, this means the following. We take an identication like: (, r, ) ( + 1 2r m 1 2r+ 2 2 , r , + n r 2 + r 2 + n 2) n r+ + r +

(6.11)

for some relatively prime and nonzero n, m Z. This new identication implies (6.7) if we apply it n times, followed by m times the identication + 2. The coordinate transformations (6.6) and (6.8) therefore still go through. However, once we are in the Poincar coordinates, we should now also take the e quotient of H 3 with respect to (6.11). This becomes the quotient: (x, y, z) exp(2 m r+ ) x cos(2n1 [mr + 1]) + y sin(2n1 [mr + 1]) , n y cos(2n1 [mr + 1]) x sin(2n1 [mr + 1]) , z (6.12)

Of course, we also have (6.9) and in general the group generated by two identications with the same xed points does not act properly discontinuously on H 3 . But this problem is circumvented because n and m are integers: since (6.12) applied n times plus (6.9) applied m times is trivial, the group they generate together is in fact cyclic and does act properly discontinuously on H 3 . It is namely readily veried to be exactly the group generated by the identication: 1 (x, y, z) exp(2 r+ ) x cos(2n1 [r + a]) + y sin(2n1 [r + a]) , n y cos(2n1 [r + a]) x sin(2n1 [r + a]) , z . (6.13) with a Z determined by am + bn = 1

for some b Z. This follows from the following two observations. First, we can apply (6.13) m times to obtain (6.12) and n times to obtain (6.9). Conversely, if we apply a times (6.12) and b times (6.9) then we obtain (6.13). So if n = 1 and m = 0, we have a quotient of a quotient. This ambiguity yields intuitively only smaller manifolds than the one we are used to (with n = 1) and imposing for example that the length of the horizon remains invariant is sucient to overcome it.

6.3

A dierent analytic continuation

We will now discuss a procedure to obtain Euclidean spaces corresponding to the wormhole spacetimes of the previous chapters, but we emphasize that we shall see in section (6.5) that this procedure also has its ambiguities. 119

First, note that there is a naive analytic continuation that also seems to work. Recall that in section 4.5 we discussed how the outer regions of every wormhole spacetime are indistinguishable from the BTZ outer regions. So in fact the Wick rotation procedure we just dened for BTZ works equally well for the wormholes discussed previously: for the m asymptotic regions, we would simply obtain m disconnected solid tori. Krasnov [26] however argues that this naive procedure is unreasonable, using two main arguments. First, one would like the dual conformal eld theories on the disconnected boundary components to be correlated in some sense, which would not happen in a natural way if we dene these theories on disconnected tori. Second, he argues that the absence of everywhere timelike Killing vector elds for the wormhole spacetimes makes the canonical analytic procedure harder to justify (although there is no globally timelike Killing vector eld for BTZ either). He therefore proposes a new prescription which we outline and extend below. As we will see, it also has the additional advantage that the Riemann surface at y = 0 is left intact, which we like since we do not lose any information regarding the nontrivial topology of the interior.

6.3.1

Formalism

Our wormhole spacetimes are constant negative curvature three-dimensional manifolds and are quotients of a simply connected part of H 2+1 . Their Euclidean counterparts should also have constant negative curvature and a Riemannian metric; this makes them the hyperbolic three-manifolds we discussed above. Since these manifolds are also quotient spaces (in this case of H 3 ), we are naturally led to the following idea. Suppose we have a spacetime dened as M 2+1 = H 2+1 /. Can we then try to nd a Euclidean counterpart E to the universal covering group and use it to dene M 3 H 3 /E as the Euclidean counterpart of M 2+1 ? Of course, in order for this to work, we at the very least need a map between the isometry groups: : Isom0 (AdS3 ) Isom0 (H 3 ) and dening this map is what we will do rst, before considering the other ingredients we need for such a procedure. We consider only the connected component of the identity of the isometry groups, since these act properly discontinuously on the spaces. Another advantage is that these elements are generated by Killing vector elds, for which we have the following result. Recall the Killing vectors of H 2+1 : x y xx + yy + zz yx + xy (x2 + y 2 z 2 )x + 2xyy + 2xzz 2xyx (x2 + y 2 + z 2 )y 2yzz (6.14)

where the Killing vectors that leave the plane y = 0 invariant are on the rst row. Now, if we map y iy these become: x iy xx + yy + zz i(yx + xy ) (x2 y 2 z 2 )x + 2xyy + 2xzz i(2xyx (x2 y 2 + z 2 )y + 2yzz ) (6.15)

and removing the extra factors of i yields exactly the Killing vector elds (6.2) of H 3 . We note that y iy is equivalent to V iV from (3.10). 120

This observation allows us to dene a map between the isometry groups. Let j be the basis of the vector elds (6.14), labeled following the rows (so E for example 5 = yx + xy ), and let j be their corresponding Euclidean counterparts, the vector elds (6.2). If we have a general Lorentzian Killing vector eld j aj j with aj R, its exponential denes an isometry and we dene:
6 6

: exp(
j=1

aj j ) exp(

E aj j ). j=1

(6.16)

From the above, we like to see as an analytic continuation of the isometries generated by (6.14) to those generated by (6.15) and we will keep calling it an analytic continuation below. We absorb the appearing factors of i in 4 , 5 and 6 in the constants a4 , a5 and a6 , exactly like we do with the Euclidean angular momentum J E = iJ of the BTZ black hole. Although is a straightforward map at the level of the Lie algebra, from (6.16) it is not yet clear how acts at the level of the groups. We work this out in the next subsection. It will be a three-stage enterprise where we rst explicitly show the P SL(2, R)P SL(2, R) decomposition of Isom0 (AdS3 ), then relate the elements of P SL(2, R) and P SL(2, C) to their Killing vectors, and nally nd an explicit form of as a map between isometries (groups) instead of a map between Killing vector elds (Lie algebras). We will work on the conformal boundaries of the spacetimes, because we already noted that there is a one-to-one correspondence between the isometries on the boundary and their extension to either H 2+1 or H 3 and calculations are easier by restricting ourselves to the conformal boundary.

6.3.2

General isometries of H 2+1

We will rst nd the general form of an isometry of AdS3 in suitable coordinates on the conformal boundary. (We do not have it yet, since in previous chapters we considered almost always isometries that left the plane y = 0 invariant.) To this end, we introduce lightcone coordinates on the conformal boundary of H 2+1 , given by u=x+y v =xy

In these coordinates, the Killing vectors of H 2+1 can be (re)written as: 2u 2v 2uu + zz 2vv + zz 2u2 u 2z 2 v + 2uzz 2v 2 v 2z 2 u + 2vzz

which are obtained by taking the sum and the dierence of two Killing vectors in the same column in equation (6.14). They have the well-known sl(2, R) form on z = 0, so on the conformal boundary the isometries they generate reduce to: (u, v) au u + b u av v + b v , cu u + d u cv v + d v (6.17)

with real coecients satisfying au du bu cu > 0 and av dv bv cv > 0; to simplify the calculations below we do not assume they have unit determinant (of course we can do so since any Mbius transformation is invariant under a simultaneous o 121

constant rescaling of a, b, c, d). Equation (6.17) shows explicitly the P SL(2, R) P SL(2, R) character of Isom0 (AdS3 ), as we desired. Below, we denote an arbitrary element of Isom0 (AdS3 ) by and its P SL(2, R) components by u and v so we have (u , v ) P SL(2, R) P SL(2, R).

6.3.3

Isometries and their Killing vectors

We will now relate the isometries of AdS3 and H 3 to their Killing vectors. The calculations below are done for Isom0 (H 3 ), so for P SL(2, C), but the story is completely analogous for P SL(2, R) so also directly applicable to Isom0 (AdS3 ). We consider the elements of Isom0 (H 3 ), written as Mbius transformations o of the conformal boundary: aw + b (6.18) w cw + d with a, b, c, d C. We should require invertibility, so ad bc = 0, but again do not assume the unit determinant condition. In the calculation, we shall also assume c = 0 and we leave it to the reader to show that the case c = 0 gives the same end result. Finally, because we will use to map covering groups to covering groups, we consider the case where (6.18) is either loxodromic or parabolic. The isometry (6.18), satisfying the aforementioned conditions, is generated by some Killing vector eld of the form w + ww + w2 w + c.c. (6.19)

and below we will nd the relation between the , , C and the a, b, c, d. To do so, let us rst look at the xed points. They are given by: w = from (6.18), or by w = 1 (a d 2c 1 ( 2 (a d)2 + 4bc) 2 4)

from the points where (6.19) vanishes. Branch cuts of the square root are along the negative real axis. Comparing w , we have, up to a scale parameter which we call : = b = (d a) b + (d a)w + cw 2 w + c.c. for some C . However, we also want to know . In general, since a nite version of a transformation induced by a Killing vector eld is of the form et for some t C, the value of is undetermined until we x t. Henceforth, we will suppose t = 1, so that we associate the transformation exp() to a Killing 122 = c.

In other words, (6.18) is generated by

vector eld . This xes the ambiguity in . We will explain our choice for t = 1 in the next section. Now let us introduce a new complex coordinate w , given by: w = kw + l mw + n

where we pick k, l, m, n C such that (6.18) is given by w e w with e = from which we also have a+d+ a+d (d a)2 + 4bc (d a)2 + 4bc . (6.20)

a+d 2 cosh( ) = 2 ad bc which is the familiar relation between trace and multiplier of a Mbius transo formation we already saw in chapter 1. We should make sure that the right hand side does not lie (strictly) in the interval (, 2) which for loxodromic or parabolic transformations can always be done by multiplying with minus the identity. A solution is given by: k l m n = 1 1 w . w+

But (6.20) is generated by the Killing vector eld w w + c.c. because its integral curves are of the form w (t) = w (0)et which clearly reduces to (6.20) for t = 1. On the other hand, the coordinate transformation gives: w w = (w w+ )(w w ) w w+ w (d a)2 + 4bc

and comparing the two Killing vector elds we obtain = so the isometry is generated by: (b + (d a)w + cw 2 )w + c.c. (d a)2 + 4bc We can also invert the relation between an isometry and a Killing vector eld. If one has an isometry generated by a Killing vector eld of the form: (Aw2 + Bw + C)w + c.c. 123

then, if B 2 4AC (, 0) and with = B 2 4AC, this becomes / (Aw2 + Bw + C)w + c.c. 2 4AC B from which one readily calculates that the corresponding Mbius transformation o is given by ( coth( 1 ) B)w 2C 2 . w 1 2Aw + coth( 2 ) + B

6.3.4

An explicit map to Isom0 (H 3 )

Let us consider an isometry of AdS3 which reduces at the conformal boundary to the form (6.17). From the previous section, we know that such an isometry is generated by the Killing vector eld: u bu + (du au )u + cu u2 u + u v with u , v dened as in (6.3.3). In x, y coordinates, this becomes: 1 u bu + v bv x u bu v bv y + 2 (du au )u + (dv av )v [xx + yy ] + (du au )u (dv av )v [yx + xy ]+ u cu + v cv [(x2 + y 2 )x + 2xyy ] + u cu v cv [2xyx + (x2 + y 2 )y ] .
3

(6.21)

Under , the isometry (6.17) maps to the element in Isom(H ) that is generated by: 1 u bu + v bv x u bu v bv y + 2 (du au )u +(dv av )v [xx +yy ]+ (du au )u (dv av )v [yx +xy ]+ u cu + v cv [(x2 y 2 )x + 2xyy ] + u cu v cv [2xyx + (x2 y 2 )y ] . Complexifying with w = x + iy gives: 1 u bu (1 + i) v bv (1 i) + [u (du au )(1 + i) + v (dv av )(1 i)]w+ 2 [u cu (1 + i) + v cv (1 i)]w2 w + c.c. which is the Killing vector eld that generates the isometry: w with: 1 [u cu (1 + i) + v cv (1 i)] 2 1 B = [u (du au )(1 + i) + v (dv av )(1 i)] 2 1 C = [u bu (1 + i) + v bv (1 i)] 2 i 2 (du au )(dv av ) + 4bu cv + 4bv cu 2 = (u 2 ) + u v w v 2 (du au )2 + 4bu cu (dv av )2 + 4bv cv A= 124 (w coth( 1 w ) B)z 2C 2 1 2Aw + w coth( 2 w ) + B (6.22)

and we are done: maps (6.17) to (6.22).

6.3.5

A dierent viewpoint

Krasnov in [27] only dened the above map back from Euclidean to Lorentzian signature and used a dierent notation. To see the agreement, we should make the translation from our notation to his work. This is what is done below, where in fact we merely re-derive the explicit form of using the (U, V, X, Y ) coordinates, grouped in the unit determinant matrices (3.7) for AdS3 and (6.3) for H 3 . This notation implies that we can use the fundamental representations of the isometry groups. To see that [27] indeed denes the inverse of , we should check three things. First, we verify that the action of the isometry groups in these coordinate systems agrees with the action dened above. Second, we check that the exponential map between Lie algebras and groups in this representation is the same as the one dened above using Killing vectors. Last, we show how the map dened in [27] is indeed the inverse of . Action of the groups on the manifolds First, we look at the map from P SL(2, R)P SL(2, R) to Isom0 (AdS3 ). If AdS3 is the space of unit determinant matrices: x= U X Y V Y +V U +X

then we dene the action of (u , v ) P SL(2, R) P SL(2, R) on it as:


T (u , v )(x) = u xv

(6.23)

This agrees with the expression of u and v as P SL(2, R) transformations of the u, v coordinates we dened in (6.17). This is because: u=x+y = and with u = Y +V U +X au cu bu du v =xy = v = av cv bv dv Y V U +X

the action on the boundary z = 0 or U + X is readily calculated to be equal to (6.17). Now for the map from P SL(2, C) to Isom0 (H 3 ). We recall the action (6.4) of P SL(2, C) on H 3 . If a w bw w = cw d w its action on the conformal boundary U +X in the coordinate w = x+iy = (Y + iV )(U + X)1 is given by: w : w aw w + b w . cw w + d w

We conclude that (6.23) and (6.4) dene the same action of the groups on the manifolds as before. 125

A representation for the Lie algebras We have seen that we can work in the two-dimensional representation of the isometry groups. We now introduce this representation for the Lie algebras as well, in order to make the exponential map a simple matrix exponential. The basis: g0 (1 + u2 )u [g0 , g1 ] = 2g2 0 1 1 0 g1 (u2 1)u [g0 , g2 ] = 2g1 g1 = 0 1 1 0 g2 2uu [g1 , g2 ] = 2g0 g2 = 1 0 0 1 of the Lie algebra sl(2, R) is easily veried to satisfy: so we can use the representation g0 =

and the exponential map now becomes a standard matrix exponential. We verify that it agrees with the one above, where we chose somewhat oddly t = 1 when we exponentiated a Killing vector eld. A general element u P SL(2, R) can be written as: u = exp[(Au u2 + Bu u + Cu )u ] 1 1 1 = exp[ (Au + Cu )g0 + (Au Cu )g1 Bu g2 ] 2 2 2 1 Bu 2Cu = exp Bu 2 2Au 1 1 1 u cosh( 2 u ) Bu sinh( 1 u ) 2Cu sinh( 2 u ) 2 = 1 1 cosh( 2 u ) + Bu sinh( 1 u ) 2Au sinh( 2 u ) u 2
2 with u = Bu 4Au Cu . Of course, a similar calculation holds if we replace u v. We apologize for the sloppy notation, in particular the replacement of Killing vectors by matrices, but our main point should be clear: the exponential map in this representation agrees with what we said previously, including the condition t = 1. In the Euclidean case, recall that sl(2, C) su(2) C. We introduce almost = the same basis as above:

1 (w2 1)w

2 i(w2 + 1)w

3 2ww

plus complex conjugates, which we omit from now on. The reader can verify that indeed: i i i [ k , l ] = klm m 2 2 2 so we can, instead of the Killing vector elds, also take the usual Pauli matrices as a representation. Then one has, for a general element w P SL(2, C) and using the same sloppy notation: w = exp[(Aw w2 + Bw w + Cw )w ] i 1 1 = exp[ (Cw Aw )1 + (Cw + Aw )2 + Bw 3 ] 2 2 2 1 Bw Cw 2 = exp 1 Aw 2 Bw = 1 w
w w w w cosh( 2 ) Bw sinh( 2 ) 2Cw sinh( 2 ) w w w w cosh( 2 ) + Bw sinh( 2 ) 2Aw sinh( 2 )

126

with w =

2 Bw 4Aw Cw .

The inverse of Now, in [27], the original map was dened as a map we call from P SL(2, C) into P SL(2, R) P SL(2, R):
3

: exp
j=1

(aj ibj )j

exp 1 (a1 b1 ) 0 (a2 + b2 ) + 2 (a3 b3 ) exp 1 (a1 + b1 ) + 0 (a2 b2 ) + 2 (a3 + b3 )

at least up to some minus signs, which are a matter of convention and which we chose deliberately a little dierent such that is truely the inverse of . Indeed, we see:
3

(ln exp) = (ln exp) = (ln exp) = (ln exp)

j=1

(aj ibj )j

1 (a1 b1 ) 0 (a2 + b2 ) + 2 (a3 b3 ) 1 (a1 + b1 ) + 0 (a2 b2 ) + 2 (a3 + b3 )

1 [(a1 b1 )(1 + i) + (a1 + b1 )(1 i)](w2 1) 2

[(a1 b1 )(u2 1) + (a2 + b2 )(u2 + 1) 2(a3 b3 )u]u [(a1 + b1 )(v 2 1) (a2 b2 )(v 2 + 1) 2(a3 + b3 )u]v

1 + [(a2 + b2 )(1 + i) (a2 b2 )(1 i)](w2 + 1) 2 [(a3 b3 )(1 + i) + (a3 + b3 )(1 i)]w w = (a1 ib1 )(w2 1) + (ia2 + b2 )(w2 + 1) 2(a3 ib3 )w w
3

=
j=1

(aj ibj )j .

We apologize one last time for the notation.

6.4

Handlebodies

This section considers some examples of the application of . First, we briey discuss the BTZ black hole but since we obtain nothing new, we quickly continue with the wormholes we dened in the previous chapters. As we said in the introduction to this chapter, the next section treats some issues and open questions regarding , but we can comfort the worried reader that these issues will not inuence the results of this section. This section is our description of the results from [26, 27].

6.4.1

Analytic continuation of BTZ

The new procedure with gives exactly the same result for the BTZ black hole (with or without angular momentum) as the canonical procedure of section 6.2. This can be seen from the generating Killing vector eld (4.9): the dilatation 127

part with factor r+ remains invariant under and the boost with r becomes a rotation of angle 2r , which easily integrates to (6.9). Note that we adopted t = 2 in section 4.4.3 rather than the convention above with t = 1 of this chapter, hence the extra factors of 2. For BTZ, we thus obtain again the solid torus picture in gure 6.1.

6.4.2

Analytic continuation of static wormholes

In some cases becomes a very simple map, namely if u = v . Indeed, if we consider the diagonal subgroup (a , a ) P SL(2, R) P SL(2, R), then: (a , a ) = a (6.24)

which should be read as a matrix equation. This means that becomes trivial and it maps the diagonal subgroup of P SL(2, R) P SL(2, R) to the real subgroup P SL(2, R) P SL(2, C). Clearly, is an isomorphism for these subgroups. From (6.21) one directly obtains that the diagonal isometries are also precisely the ones that leave y = 0 invariant, so this comprises all the wormhole spacetimes dened in chapter 4. Furthermore, from (6.14) and (6.15), we see that the Euclidean and the Lorentzian groups act in exactly the same way on y = 0, so the surface obtained as H/ is the same in both cases. On the other hand, their extension to nonzero y is generally dierent. So let us see what kind of Euclidean spaces we obtain. If we follow the same procedure as in section 4.2, we can again rewrite the Killing vector elds of H 3 in terms of the new coordinates y and z that are now dened as: y z = y2 + z 2 y = z The Killing vector elds whose orbits lie in the plane y = y = 0 are given by: x xx + z z (x2 z 2 )x + 2xz z

from which it is clear that their action on the boundary z = z = 0 is in fact the same as on y = y = 0. This means that on the boundary H 3 of H 3 , the x-axis is left invariant and there are two images of the surface S = H/: one below and one above the x-axis. So the boundary of the three-manifold we obtain is precisely the Schottky double of S (recall denition 1.5.10): (E )/E with (E ) as in section 6.1.1 and of course E () which is in fact just again. As we promised in the introduction, the corresponding Euclidean manifold is a handlebody with the Schottky double of S as its boundary Riemann surface. Intuitively, it is just the lled Schottky double of S. Example: genus 2 We briey discuss the case where yields a handlebody whose boundary is a genus 2 surface. We will not give any formulas and this discussion is used only to x ideas and to provide an intuitive idea of the handlebodies. In the gures 6.2 and 6.3, we show a genus 2 handlebody. This handlebody is the Euclidean version of a static wormhole spacetime for which the Schottky 128

Figure 6.2: The Euclidean space associated with the trinion spacetime becomes a genus 2 handlebody. On the left, a fundamental domain in H 3 and on the right the handlebody. Again, the trinion that we obtain as y = 0 is shaded, as well as the three horizons. double of the y = 0 slice is a Riemann surface of type (2, 0, 0). This means, by the formula given directly below denition 1.5.10, that this slice is of type (1, 0, 1) or of type (0, 0, 3). So it is either a toroidal wormhole (gure 6.2) or a trinion (gure 6.3). In the guresm, we shaded the invariant surface y = 0 which is totally geodesically embedded in both handlebodies. Note that this is not the only way we could have drawn these surfaces as embeddings in the three-manifolds, for example by twisting the surface in one handle a full turn we would obtain another, inequivalent, embedding. It would be interesting to nd all such embeddings, a subject which is probably related to knot theory. We also indicated the periodic geodesics on the y = 0 slice in the handlebody. These lines are physically important because they are the horizons in the spacetimes obtained from these surfaces. Note how there are three of them for the trinion and only one for the toroidal wormhole. On the left in both gures, we indicated a fundamental domain in H 3 , bounded by hemispheres as well as the boundary of H 3 . The hemispheres are totally geodesic surfaces in the metric (6.1) and are to be glued together pairwise as indicated. On readily sees that the intersection of the closure of the fundamental domain with H 3 is a four-holed sphere whose bounding circles are pairwise identied, which indeed yields a genus 2 surface.

6.5

Issues

Although we just saw that works well for general BTZ and the static wormholes obtained via the diagonal subgroups, there are some issues that need to be resolved if we want to apply it to the more general rotating wormholes of [12, 9, 10]. We discuss these issues in this section.

6.5.1

Conjugation freedom and coordinate dependence

Suppose we have a Lorentzian manifold M 2+1 = AdS3 / for some covering group . If we just naively apply to the covering group without adding any 129

Figure 6.3: Fundamental region and handlebody of the Euclidean toroidal wormhole space. The slice y = 0 is a deformation retract of the handlebody and has a single boundary component, as indicated. The reader might check that the indicated surface in the handlebody has indeed one horizon and is topologically equivalent to the toroidal wormhole of gure 1.6. further information, we run into an ambiguity that originates from conjugation freedom as follows. Any covering group can always be conjugated with an isometry (say ), and a conjugated covering group 1 yields the same quotient manifold (only of a dierent part of AdS3 ). This means that we always have the freedom to conjugate an isometry (group) on either the Lorentzian or the Euclidean side before applying or its inverse. However, applying to the conjugated covering group 1 instead of to may yield a dierent Euclidean space. We use the analytic continuation of the BTZ black hole as an example of what we mean. We work in the four-dimensional embedding coordinates of AdS3 and H 3 and note again that the analytic continuation y iy is equal to V iV by (3.10). The map is then equal to Wick rotating the Killing vector elds (3.6) instead of (6.14) and again absorbing the factors of i in the constants. In particular, the boost V Y + Y V becomes a rotation and the rotation V U U V becomes a boost. Now let us take a closer look at the BTZ black hole in (X, Y, U, V ) coordinates. Including rotation, it is generated by (4.9) which in these coordinates becomes: r+ (U X + XU ) + r (V Y + Y V ) We can use conjugation freedom to rotate U into V and X into Y so that this is equivalent to: r+ (V Y + Y V ) + r (XU + U X ) (6.25) In the Lorentzian case, taking a quotient of a part of AdS3 with respect to the isometry group generated by (6.25), we obtain the same quotient spacetime, namely the BTZ black hole. But whereas U X + XU is invariant under the analytic continuation V iV , we just mentioned that V Y + Y V is not. In terms of Mbius transforo mations the rst becomes a dilatation and the second a rotation. We see that we cannot rotate them into each other as on the Lorentzian side, and in fact 130

maps the Killing vector eld (6.25) to r+ (V Y Y V ) + r (XU + U X ) Taking a quotient yields a solid torus, uniformized in H 3 by the identication (x, y, z) e2r x cos(2r+ )+y sin(2r+ ), y cos(2r+ )x sin(2r+ ), z (6.26)

which is (6.9) with r+ and r interchanged (we again used the convention with t = 2 here). But this new torus is a dierent Euclidean manifold and we conclude that conjugating the covering group might indeed yield dierent resulting Euclidean spaces. Note that we do not claim that conjugating a covering group will always yield dierent Euclidean spaces, but in some cases it does and we need to x this ambiguity. This, in turn, might be done as follows. First of all, this conjugation freedom relates to coordinate dependence in the following way. When we use (or its inverse), we attach a special meaning to the y = 0 slice: it is the only slice invariant under the maps y iy E . However, its location is ambiguous because we can always add an isometry to the transformation (3.10). This means that is coordinate-dependent and the conjugation freedom is a parametrization of the coordinate dependence. So we can x the conjugation freedom by choosing a natural y = 0 slice, which clearly is not a problem for the static wormholes. Also, for generally rotating BTZ, we may simply use the canonical procedure of section 6.2 to overcome the diculty we sketched (in that sense, the BTZ black hole was not that good an example). On the other hand, choosing such a slice might be a nontrivial issue in the case of non-zero angular momentum wormholes. In fact, a similar question was phrased but unanswered in [9], where Barbot asked for a natural spacelike surface in AdS3 whose boundary circle would pass through all the spatially separated limit points of the covering group. Such a slice might be a natural choice for a y = 0 slice. In any case, we need to investigate this a bit further to overcome this ambiguity. However, we do not do so here but we continue instead by mentioning another issue we encountered.

6.5.2

Analytic continuation of subgroups

The issues we are about to discuss arise because we want more than a simple map between Isom0 (AdS3 ) and Isom0 (H3 ). What we actually desire is to use to map subgroups to subgroups. In other words, we want to nd a subgroup E P SL(2, C) given a subgroup of P SL(2, R) P SL(2, R). (We also want that E acts properly discontinuously and xed-point freely on H 3 , but we leave this for a future investigation.) We could in fact think of two possible canonical ways of obtaining a Euclidean subgroup from a Lorentzian one using and both turned out to be problematic if we consider o-diagonal subgroups. First, we can analytically continue all the elements of the group, so we dene E = (). Second, we might analytically continue only the generators of and then using these to generate a subgroup of P SL(2, C). The issues we encountered when trying to apply these methods are outlined below. 131

Is () a subgroup? Let us rst check what we can say about the denition E = (). Since there is no decomposition of P SL(2, C) in two commuting subgroups, whereas P SL(2, R) P SL(2, R) trivially admits such a decomposition, the full isometry groups are not isomorphic. But this means that cannot be a morphism in the general case of rotating wormholes, as it was for the static wormholes. This rises a problem, because in these cases we cannot be (trivially) sure that () is a subgroup of P SL(2, C) instead of a mere subset. Actually, asking for a morphism is actually a bit superuous if we want () to be a subgroup. Since the inverse condition is already satised, namely: ( n ) = ()n (6.27)

which holds in particular for n = 1, we need to check only the composition law. This formally means that () is a subgroup of P SL(2, C) only if we are able to express (1 )(2 ) as of some word of 1 and 2 for all 1 and 2 in . We are, unfortunately, not aware of a proof nor a falsication of this statement so this is still an open question. Is < {(i )} > unambiguously dened? Another viewpoint was taken in [27], where the proposal was made to map only the generators of into elements of P SL(2, C). If is generated by some nite set {i }, so if =< {i } >, we might have to rst map all the generators to P SL(2, C) and use the newly obtained nite set {(i )} P SL(2, C) to generate a new subgroup E < {(i )} > and we use E to take a quotient of H 3 . This is by construction a subgroup of P SL(2, C) so there are no morphism properties to worry about. However, the procedure is ambiguous: as in chapter 2 dierent markings on the same Riemann surface could be dened, we can also choose new generators, for example instead of {1 , 2 , . . .} {1 2 , 2 , . . .}. These two choices yield the same group , so the same quotient manifold, but under the two distinct sets of generators might yield dierent subgroups of P SL(2, C) and so we might obtain dierent Euclidean spaces. This would be problematic since we then would not know which Euclidean space to pick. Our problems are of course solved if the resulting Euclidean spaces are the same after all. If we were to be so lucky, then the image under of the dierent choices of generators should yield conjugate groups. In formulas: < {(1 ), (2 ), . . .} >= < {(1 2 ), (2 ), . . .} > 1 for some P SL(2, C). This means that if we prove that there does not exist an P SL(2, C) such that: W (1 ), (2 ) 1 = (1 2 ) 132

we choose

as well as: W (1 ), (2 ) 1 = (2 ) for any possible words W or W of (1 ) and (2 ), then we ruled out the possibility of conjugate groups as well. However, as for the previous statements, we do not yet have a proof or a falsication. The counterexample below implies that is in any case not generally trivial. An example In this example, we show that it is not always possible to express (1 2 ) as some word of (1 ) and (2 ) for general 1 , 2 Isom0 (AdS3 ). In formulas again: (1 2 ) = W (1 ), (2 ) for any nite word W of (1 ) and (2 ). This counterexample is not all-conclusive but it implies two things. First, we can conclude that dened above is (in this case at least) not the identity. Second, it shows that the above two possible analytic continuations of subgroups are dierent (again, in this case). This is because the result immediately implies that (< 1 , 2 >) is not the same subset of P SL(2, C) as < (1 ), (2 ) >. Our example is similar to the toroidal wormhole from the previous chapter. The static wormhole is purely generated by the diagonal subgroup of P SL(2, R) P SL(2, R) and thus free of these issues, so let us generalize it somewhat and include some rotation in the rst isometry. We thus take 1 = (u1 , v1 ) with u1 and v1 dened as: v1 : v ev v u1 : u eu u

which is the rotating BTZ black hole identication generated by: 1 1 u uu v vv = (u + v )(xx + yy ) (u v )(xy + yx ) 2 2 so with u = v we have J = 0. We also dene 2 = (u2 , v2 ) to be the isometry: u2 : u 1 v2 1 (u + 1) 1 : v 1 (v + 1)

which is just (5.1) with = , which we chose for simplicity. Although we are unable to proof it here, the group they generate, so < 1 , 2 >, probably acts discretely on some region of H 2+1 and if we take the usual quotient we probably get a nice (see chapter 3) three-dimensional manifold. A thorough proof would need application of the results of [9, 10]. We do not do so here, so we consider it a formal example of a subgroup of Isom0 (AdS3 ) without discussing any quotient manifold. We rst analytically continue 1 , then 2 , then their composition 1 2 = (1u 2u , 1v 2v ) and nally show what we promised, so that the latter cannot be obtained as some nite word of the former two. 133

One has: (1 ) = exp u uu v vv 1 = exp [(1 + i)u + (1 i)v ]ww 2 exp[ 1 (1 + i)u + 1 (1 i)v ] 0 2 2 = 0 1 We also calculate: (2 ) = u2 = v2 = This means: (1 ) : w e(1+i)u +(1i)v w 1 . (2 ) : w 1 (w + 1) On the other hand, we have: 1 2 : u v e u 1 e v 1
1 (u+1) 1 (v+1) 1

1 1

which analytically continues to: (1 2 ) = exp u [u2 + (1 eu )u eu (1 1 )] + u v 1 [(1 + i)u + (1 i)v ]w2 w = exp 2 1 + [(1 + i)u (1 eu ) + (1 i)v (1 ev )]ww 2 1 [(1 + i)u eu (1 1 ) + (1 i)v ev (1 1 )]w 2 with u = u cosh1 [ cosh( 2 )] (eu + 1)2 4eu 1 v = cosh1 [ cosh( v )] 2 (ev + 1)2 4ev 1 .

We see that (1 2 ) depends not only on (1 + i)u + (1 i)v , but also on different combinations of u and v . (One may verify this explicitly by writing the corresponding Mbius transformation out completely, picking a (scale-invariant) o quantity like d/c and then taking derivatives with respect to u and v .) However, the only linear combination of s appearing in (1 ) and (2 ) is exactly (1 + i)u + (1 i)v , so this is also he only combination appearing in any word composed of (1 ) and (2 ). Hence, we conclude that (1 2 ) cannot be expressed as a word of (1 ) and (2 ). This means that (< 1 , 2 >) =< (1 ), (2 ) > by which we mean inequality as a subset of P SL(2, C). 134

6.5.3

Summary

We briey summarize our discussion of the ambiguities surrounding . Earlier, we have seen that is a valid invertible map between Isom0 (AdS3 ) and Isom0 (H3 ). It also satises some nice properties like (6.24) and (6.27). On the other hand, there are some open questions that need to be resolved. First of all, we have seen that there is conjugation freedom both the Euclidean and the Lorentzian side. We showed how this freedom is related to choosing a y = 0 slice and indicated a direction that would give us a natural y = 0 slice, although it remains an open question whether this can be unambiguously done in all cases. Second, if we want to use to nd a subgroup of P SL(2, C) acting properly discontinuously on H 3 given such a subgroup of Isom0 (AdS3 ), we should also nd a method of mapping subgroups to subgroups. We proposed two methods of mapping, namely applying to the entire group or only to a set of generators of the group. In both cases, we still have to investigate either whether () is a group or whether < {(i )} > is unambiguously dened. What we have proved already that the two methods are not always equivalent. Note that in the diagonal cases of section 6.4, reduces to an isomorphism between the subgroups and the conjugation freedom can also be xed in a natural way. So the issues we mentioned do not apply in this case, as we promised at the beginning of section 6.4. Remark. This nal remark considers the convex cores of the Euclidean mani folds. Given a Schottky group , the convex core of H 3 / is H 3 / with H 3 the 3 smallest subset of H that contains () on its boundary. We will not explain this concept in detail but assume that the reader is familiar with it, otherwise [43] is a good reference. Using the convex core, we like to show that the cases where is known to work, which are the static wormholes of chapter 4 and both the rotating and the non-rotating BTZ, are indeed special. This is because, only in these cases, the convex core of the obtained Euclidean space is one- or two-dimensional. For general BTZ, the convex core is the quotient of the x-axis and for general static wormholes it is precisely the Nielsen kernel (denition 1.5.12) of the surface obtained at y = 0. So both for BTZ and for static wormholes, the convex core lies entirely in the plane y = 0 and hence is completely invariant under the analytic continuation. On the other hand, for a rotating wormhole, the convex core is generally a lump of three-dimensional space and might not be invariant under ; in any case the metric in it changes signature under the analytic continuation. This might also imply that there simply is no natural y = 0 surface for these spacetimes, something we wished to nd at the end of section 6.5.1 to x the conjugation freedom. We think this shows intuitively how these cases are indeed special and the observation might be useful in understanding whether or not remains valid for a general wormhole spacetime.

6.6

Conclusion

In this chapter, we have made the step from Lorentzian signature to Euclidean signature three-dimensional wormholes. After formulating the well-known re135

sults for the BTZ wormhole, we have seen how a possible analytic continuation can be dened for the wormholes under consideration. In the static case, the corresponding Euclidean signature manifolds turned out to be handlebodies, with the y = 0 slice left completely invariant and embedded in them. In the other cases, there are issues to be solved which we also outlined. Next steps are not hard to identify. First, we would like to solve the ambiguities surrounding in the rotating wormhole case. This means that we should either nd a method of analytically continuing subgroups in an unambiguous way, or obtain a proof that we cannot do so. Also, we need to show how the Euclidean groups act on three-space, because we want to know whether we can use them to take a quotient. Second, there are some open questions concerning the handlebodies as well. For example, we do not yet know which are the possible llings of a Riemann surface: given a Riemann surface S, how many dierent Euclidean three-manifolds exist that have S as their conformal boundary? Also, once we have a handlebody, how many Lorentzian wormhole spacetimes correspond to a handlebody? This last question is probably related to the possible inequivalent embeddings of y = 0 slices inside this handlebody. Perhaps dierent embeddings of such slices are related to dierent llings? And nally, since the boundary of the y = 0 slice is a set of circles embedded in three-space, are there perhaps relations to knot theory? Third, with the new analytic continuation we can also try to answer more physical questions. In particular, we would like to know which temperature and entropy we should associate with these wormholes. There are claims in the literature [28] which sound interesting, so we would like to investigate them further.

136

Chapter 7

Holography
In this chapter, we discuss holography. The main idea behind the holographic principle [42, 41] is that the degrees of freedom of quantum gravity in a spacetime of d+1 dimensions can be captured by a d-dimensional eld theory without gravity. This idea is an extension of the result that the entropy of a black hole is proportional to the horizon area (instead of the enclosed volume) so all degrees of freedom should be encoded on the horizon. In string theory, one concrete example of holography is the AdS/CFTcorrespondence [31, 46, 21] which is a conjectured duality between string theory in d + 1-dimensional asymptotically AdS spacetimes (spacetimes that look like AdS near their conformal boundaries) and a dual d-dimensional quantum eld theory. This eld theory can be thought of as living on the conformal boundary of the spacetime. A frequently used example is string theory on AdS5 S 5 versus the conformally invariant N = 4 super Yang-Mills theory, but the conjecture extends to many other cases. Unfortunately, these conjectured dualities are far from completely understood and currently none of them has been proved. To gain more understanding in the correspondence, one can consider the low-energy limit in which string theory can be replaced by the better understood classical gravity. An account of the already very nontrivial results that can be obtained in this limit can be found in the lecture notes [38]. Below, we will stay in this framework as well. Now, the way the AdS/CFT correspondence works is by using a dictionary to translate results from calculations in a space(time) M to data of a (conformal) eld theory on the conformal boundary M and vice versa. In the classical limit, this dictionary says that the metric on M (plus a suitable choice of coordinate system) translates into an induced metric on M and also the vacuum expectation value of the stress energy tensor T of the eld theory on M [38]. Obtaining the (vev of the) stress-energy tensor of the dual eld theory would be a useful result. This is because we actually do not know at all what the conformal eld theory dual of classical three-dimensional gravity is (if it exists), and knowledge of T is of course helpful in solving this problem. This is why we investigate it below for our Euclidean wormholes, so for handlebodies. We will rst show how the dictionary works, discuss some properties of the coordinates we choose and nally try to nd a suitable coordinate system from which we can obtain T . 137

We henceforth restrict ourselves to the setting of the previous chapter. This means that we assume that the three-manifold under consideration is a handlebody and carries a metric of signature (+++). We also remark that most of the calculations below involve only gravity, altough the results should sometimes be interpreted in the language of the conformal eld theory. This is why we sometimes add a remark to explain the result from the conformal eld theory perspective.

7.1

The asymptotic form of the metric

An important result we will use below is that the asymptotic form of the metric on the spacetimes we consider can always be put in an expansion proposed by Feerman and Graham: ds2 = with g = g(0) + g(2) + 2 g(4) + . . . . (7.2) The conformal boundary is located at = 0. (This is of course the special coordinate system mentioned in the introduction.) In this expansion, it is natural to use as a dening function in the conformal compactication of the spacetime (this function was called in chapter 3). This means that g(0)ij becomes the induced metric on the conformal boundary. Given any such a g(0) and imposing the Einstein equations and the fact that the Weyl tensor vanishes identically, the expansion (7.2) stops at order 2 . Better still, it can be shown [39] that g must be of the form: g(x, ) = g(0) + g(2) + with 2 1 g(2) g(0) g(2) 2 (7.3) d2 1 + gij dxi dxj 42 i {0, 1} (7.1)

and with Tij = Tji following properties:

1 R(0) g(0) + T (7.4) 2 a symmetric tensor on the conformal boundary with the g(2) =
i

Tii

Tij = 0 = R(0)

(7.5) (7.6)

where R(0) is the scalar curvature of metric g(0) induced on the conformal boundary and is the covariant derivative with respect to g(0) . We also raise and lower indices with g(0) . So suppose we have a metric g(0) on a Riemann surface, then near the boundary of the corresponding handlebody there will always be certain charts in which the metric is in the form (7.1) with g given by (7.3). There is considerable freedom in the particular form of the expansion, so in the choice of g(0) and T , since the Einstein equations are (locally) satised for any boundary metric g(0) and for any T satisfying (7.5) and (7.6). However, if we pick g(0) and T completely arbitrarily, we may develop for example conical singularities inside, so for some 138

> 0, or perhaps further away from the boundary in a chart that is not even covered by the (, x) coordinates. If we want to avoid these singularities and require the manifold to be smooth also in the interior then T and g(0) should be related in some sense. We therefore want to nd, for every g(0) , an appropriate T so that the metric in the interior of the handlebody is everywhere well-dened. This is a geometrical problem, and one that we will be concerned with for the remainder of the chapter. Remark. In CFT language, as one might have expected after reading the introduction, T is the vacuum expectation value of the stress-energy tensor of the boundary theory [38]. The boundary metric g(0) acts as a source for T . This implies that nding T as a function of g(0) translates to obtaining T as a function of its source. We have already shown in (7.6) that Tii is not identically zero, so indeed the boundary theory is not conformally invariant [44, Appendix D]. This reects the existence of the conformal anomaly we mentioned in the introduction.

7.2

Complex coordinates

From the Feerman-Graham expansion (7.1) and (7.3), we see that the metric g(x, ) induced on the surfaces obtained for constant is smoothly deformed as a function of and becomes g(0) for = 0 (again using as the natural conformal factor). A deformation of the metric can always be split into a scaling combined with a change in the conformal structure and in this section, we nd this splitting for the -dependent change of the boundary metric g(0) . We from now on assume g(0) is Hermitian, so is of the form: g(0) = 2 dzd z (7.7)

with 2 (z, z ) a real, strictly positive and bounded function in the chart z. From chapter 1, we know that this is in fact very natural: it simply means that we choose the complex structure that corresponds to the conformal structure induced on the boundary by the three-dimensional metric. We will henceforth also use both z and w to denote a complex coordinate; whenever we mean the real coordinate z from (6.1) we will explicitly say so. Remark. We recall that a change in the conformal structure is given by the BersBeltrami dierentials of (2.6). As we saw in chapter 1, given such a dierential of the form (dz)1 d we can deform any Hermitian metric in the following z way: 2 dzd 2 |dz + d|2 z z (7.8) In matrix form, this becomes: 2 2 0 1 1 0 2
1+ 2 1+ 2

(7.9)

Note that the metric becomes singular if || = 1. We start our analysis by taking a closer look at T . In complex coordinates on the boundary, T is still symmetric Tzz = Tzz and imposing a reality condition 139

gives Tzz = T zz . Because of the trace condition (7.6) we see that 42 Tzz = R(0) . But with the ansatz (7.7) the scalar curvature of g(0) is given by: R(0) = 42 z z log 2 if we use the conventions from chapter 3. We therefore have: Tzz = z z log 2 . (7.10)

This is what we need for now; we will have more to say about the free component of T , which is Tzz , below. Substitution in (7.4) yields: g(2) = 1 0 1 Tzz + z z log 2 1 0 0 2 0 Tz z

which we substitute in (7.3) to obtain: g= 2 2 0 1 Tzz z z log 2 + + 2 1 0 Tzz 2 z z log 2 2Tzz z z log 2 Tzz Tzz + (z z log 2 )2 2 2 2 2Tzz z z log 2 8 Tzz Tzz + (z z log )
2

where we dened

We can collect the terms in a single matrix 1 2a 2 Tzz 2 + g = 2 a2 2 1 1 Tzz Tzz 2 + 2 2a 2 a = 1 +

1 2

2a 2

2a 2 Tzz

Tzz Tzz

(7.11)

log 2 = 1 R(0) . 2 z z 2 8 Note the similarity between (7.9) and (7.11). Finally, substituting (7.11) in (7.1) results in: 2 a 2 d2 |dz + d|2 z (7.12) ds2 = 2 + 4 with Tzz = . (7.13) 2a 2 We see that the result of increasing is scaling the metric with a factor a2 and deforming it with the Beltrami dierential (dz)1 d. z Note again that the expansion of the metric in three-dimensional space is fully determined by means of a metric on the conformal boundary, together with the (2,0) form Tzz that also lives on the boundary. There are, as was mentioned in other papers [15, 27], basically two conformal structures at the boundary: the one associated with the boundary Riemann surface and the one associated with the Beltrami dierential . From the above we see that the second one is used to indicate in which direction in Teichmller space we move if we go to u nonzero . We mention again that, if we want the metric to be smooth in the interior, there should be some relation between Tzz and 2 . This relation is discussed in subsection 7.4, after we discuss some properties of the expansion (7.12) in the next subsection. 140

7.3

Further properties of the expansion

Having written the Feerman-Graham expansion in complex coordinates, we now deduce some further properties of it. In particular, we show when the metric becomes invalid and discuss some properties of T . Validity of the metric The metric (7.12) becomes singular whenever: | | = Tzz =1 a 22

Now (only) if 2 |Tzz | > 1 R(0) , this holds for nite positive , namely for: 4 = c 22 |Tzz | z z log 2 (7.14)

The right hand side is independent of the chart we pick from the boundary complex structure, so we cannot change the value of c by changing charts using a holomorphic map w = w(z). In that sense, c is a function on the boundary Riemann surface constructed from the tensors 2 dzd and Tzz dz 2 . z One readily veries that the singularity at c is of rst order whenever |Tzz | = 0 and is of second order otherwise. As we mentioned in the previous section, at c there should exists a coordinate transformation that does away with the singularity. If such a coordinate transformation does not exist, the three-manifold will not solve Einsteins equations there (because of the existence of for example conical singularities). Although this problem can be trivially overcome by adding a new boundary component just before c , we focus ourselves on geodesically complete spaces for which this is not allowed. Holomorphicity of Tzz We still suppose the boundary metric is Hermitian, so of the form (7.7). From the conservation of T : i Tij = 0 together with the fact that g(0)zz = g(0)z = 0, we deduce: z
z Tzz

z Tz z

=0

The nonzero connection coecients of g(0) are z = z log 2 and zz = zz z 2 z log . We have: z Tzz + z Tzz z Tzz = 0 zz

Substituting the values of and of Tzz given in (7.10) yields: z Tzz z z z log 2 + z (log 2 )z z log 2 = 0
1 and using z z log 2 = 4 R(0) 2 :

1 1 z Tzz + z (R(0) 2 ) z (2 )R(0) = 0 4 4 141

1 z Tzz + 2 z R(0) = 0 4 which means that Tzz is holomorphic, namely z Tzz = 0, if the curvature of the metric g(0) is constant. On the other hand, if the metric has non-constant curvature, Tzz is not holomorphic. Weyl transformations of the boundary metric In [37], Skenderis considered an arbitrary rescaling of the metric on the conformal boundary. His results (in our notation) are that there always exists a dieomorphism that retains the form of the metric (7.12) but transforms: 2 e2 2 together with
2 Tzz Tzz + 2z 2(z )2 2(z log 2 )z

which is equivalent to:

(7.15)

(7.16)

We will not need the explicit form of the dieomorphism, but give an example of a special case below. Remark. In fact, the dieomorphism that induces the transformation (7.15) and (7.16) reduces to the identity on the boundary = 0. This means that from the viewpoint of the boundary conformal eld theory, the dieomorphism is just a Weyl rescaling of the boundary metric g(0) . If the theory were conformally invariant, then T would not change and the fact that it does shows again that there is a conformal anomaly. Example of a Weyl transformation We take H 3 as an example. The metric (6.1) can be rewritten in FeermanGraham form as: 1 d2 (7.17) ds2 = 2 + dwdw 4 with w = x + iy a complex coordinate on P1 and = z 2 , by which we mean the real coordinate z of (6.1). Comparing with (7.12), we see that in this coordinate system 2 = 1 and Tww = 0. We should note that dwdw is not a good metric on P1 since it diverges at the point at innity. (To see this, just apply the coordinate transformation w = w 1 and consider the limit w 0.) However, locally (7.17) is as good a Feerman-Graham expansion as any. We now want an explicit coordinate transformation to a primed coordinate system ( , w ) such that the metric becomes (7.12) with 2 = e2 for = (w , w ) some real, smooth and bounded function. By the above result, Tw w is then given by (7.16) with Tww = 0 (and z replaced by w of course). Now, such a coordinate transformation was found in [35, 29] and is given by: e2 (1 + e2 |w |2 )2 e2 w w=w + 1 + e2 |w |2 = 142

(7.18)

Note that the dieomorphism indeed reduces to the identity at = 0. It is straightforward to verify that the metric indeed becomes: ds2 = d 2 2 2 + a |dw + dw | 4 2

with indeed 2 = e2 and, as usual: a = 1 + w w log 2 . 2 2

The deformation of the conformal structure is given by: = with indeed which agrees with (7.16) if 2 = 1 and Tww = 0.
2 Tw w = 2w 2(w )2

Tw w 2a 2

7.4

Fixing the global data

In this section we will show how to nd Tzz as a function of 2 for the handlebodies. First, note that this should be done globally. This is because if we want to avoid for example conical singularities inside the three-manifold, we need the right periodicity of a corresponding angular coordinate (just like the periodicity in Euclidean time for the BTZ black hole). This means that a local analysis, like we did in the previous subsections, is no longer sucient. Nevertheless, this local analysis is an important rst step. This is because the handlebodies are always quotients of H 3 , so locally we can always put the metric in the Feerman-Graham form (7.17), so with 2 = 1 and Tzz = 0. Furthermore, we know that in this metric, the transition functions are exactly the covering transformations, so the elements of the Schottky group extended to isometries of H 3 . All other forms of 2 and Tzz must be dieomorphic to these. However, this result is on itself not enough. The metric (7.17) induces locally the metric g(0) = dzd on the conformal boundary, which, as we said z before, diverges at innity. Furthermore, the transition functions are Mbius o transformations on P1 , and it is easily veried that if: z= then dzd = z aw + b cw + d

This means that the g(0) we obtain from (7.17) is also not invariant under the covering group and thus does not at all descend smoothly to a metric on the boundary Riemann surface. In other words, the covering transformations are isometries of the three-dimensional metric only and not of the two-dimensional metric it induces. 143

1 dwdw. |cw + d|2

This is where the Weyl transformations come in. If we nd a nowhere vanishing function 2 : P 1 R + that satises, for all element of the covering group : (2 ) and diverges only at (), then 2 dzd z is a (1,1) automorphic form (denition 1.3.14) for and descends by construction to a metric on S = ()/. Once we are equipped with 2 , we are done. We start from the metric (7.17) and apply the coordinate transformation (7.18) with e2 = 2 . The new metric is then of the form ds2 = 2 a 2 Tzz d2 + |dz + d|2 z 42 2a 2 (7.19) z z = 2

with the corresponding Tzz given by: 1 2 Tzz = z log 2 (z log 2 )2 . 2 (7.20)

This is the relation we were looking for: we have obtained Tzz as a function of 2 everywhere on the boundary surface of a (smooth) handlebody. We stress that (7.18) only goes through globally if we start from H 3 everywhere: this means that the boundary Riemann surface must be Schottky uniformized. Transformation properties of T Now, let us verify that Tzz so obtained has the right transformation properties under the covering group transformations, so it indeed is an automorphic (2,0) form and descends smoothly to S. Under a general holomorphic coordinate transformation z = z(w), we of course have: 2 (z, z) = Substitution in (7.20) gives: Tzz = Tww dz dw
2

dz dw

2 (w, w)

3 2 w z 3 w z + w z 2 w z

= Tww

dz dw

S(z)

where S(z) is the Schwarzian derivative of w = w(z), introduced in section 2.6. This seems problematic, but recall that (7.20) only works on P1 where S is Schottky uniformized, so the only coordinate transformations we should worry about are the elements of the Schottky group . But these are by denition Mbius transformations, for which the Schwarzian derivative vanishes identically o (see again section 2.6). This implies that Tzz is indeed a (2, 0) automorphic form for on P1 . It hence descends to a (2, 0) tensor on S. It would be interesting to see if this appearance of the Schwarzian can be linked to the quasiconformal mappings of chapter 2. 144

Constant surfaces We will now show that, when the metric is of the form (7.19) (with all the above properties of 2 and Tzz ), then the covering group , acting on the three-space, leaves the slices of constant invariant. This is because not only the two-dimensional metric 2 dzd but also the z three-dimensional metric (7.19) is invariant under the transformations (, z) (, (z)) (7.21)

for all . This in turn follows from the fact that 2 and T are automorphic forms, and a only contains the curvature associated to 2 so a is an automorphic (0,0) form as well. But if the metric is invariant, this means that (7.21) are isometries and furthermore these isometries reduce to the familiar Mbius transo formations on the boundary P1 . By the uniqueness of the extension of these Mbius transformations to isometries of H 3 , we conclude that the elements of o the covering group act as (7.21) also in the bulk. The invariance of the slices of constant under means that if M 3 H 3 /, then the slices M 3 |= are also Riemann surfaces of the same type as the boundary, but (if Tzz is nonzero) with a dierent conformal structure. This result of course holds as long as the metric is in the form (7.19), so until c dened in (7.14). We emphasize that, in the coordinates giving rise to the metric (7.17), the isometries did not leave slices of constant invariant, so it is an additional advantage of this particular form of the metric. Summary Let us give a brief summary from a slightly dierent viewpoint. Suppose we have a Riemann surface S which we want to correspond to the boundary of a handlebody. Then we should Schottky uniformize the Riemann surface on P1 , ll in the sphere and give it the metric (7.17). We then extend the Schottky group to isometries of H 3 and know that the quotient, by construction, becomes a smooth three-manifold whose boundary is S. Then, if we want to obtain the Feerman-Graham form of the metric (globally, or at least everywhere on S), we rst need a metric on S. We lift this metric to an automorphic (1,1) form 2 on P1 and use an asymptotically trivial coordinate transformation (explicitly, equation (7.18)) to obtain Tzz as (7.20). Then, until c , the elements of the Schottky group then leave the slices of constant invariant.

7.5

Constant curvature metric results

A natural choice for 2 , so for g(0) , is the constant curvature metric that exists on any Riemann surface and that we dened in chapter 1. With this choice of 2 , we perhaps might nd a natural form of the corresponding Tzz as well, so it seems a worthwhile direction to be investigated and we do so in this subsection. In order to do this, we should nd a covering map from the uniformization of the boundary Riemann surface in C or H (where we have the constant curvature metric) to the Schottky uniformization (where we can nd Tzz as a function of 145

2 ). Below, we rst treat the cases of genus 0 and of genus 1 separately. Just like in earlier chapters, the sphere is rather trivial and the torus provides us again with a nice and explicit example. Unfortunately, it turns out that it is these are the only cases that can be treated explicitly: once we try to apply the same procedure to higher genus Riemann surfaces we encounter mathematical obstructions. We outline this in the second part of this subsection. We stress that the constant curvature metric, although very natural, is certainly not the only possible choice for a metric on S, as sometimes seems to be the opinion in the literature.

7.5.1

The almost trivial sphere

The sphere is the easiest example, since the Schottky group uniformizing it is just the trivial group. However, as we said before, the metric dzd induced z from (7.17) is not a valid metric on P1 since it diverges near innity. In chapter 1, we came with an alternative: g(0) = dzd z (1 + |z|2 )2

which has constant scalar curvature R(0) = 8. Substituting this metric in (7.20) gives the result: Tzz = 0. For clarity, we give the nal form of the Feerman-Graham metric: ds2 = z (1 )2 dzd d2 + 2 4 (1 + |z|2 )2

where the factor (1 )2 is of course a2 . The validity of this metric can also be explicitly veried using the coordinate transformation (7.18) with e2 = (1 + |z|2 )2 .

7.5.2

Contracting loops on a torus

In [32, 18, 6] the comment was made that, if we have a handlebody whose conformal boundary is a Riemann surface of genus 1, then there are multiple possibilities for the contractible cycle. Basically, a solid torus is created from a boundary torus as follows: we pick any simple loop on the torus we want n m (so in terms of the fundamental group any loop A B with n, m Z relatively prime) and then simply say that this loop bounds a disk of the three-manifold. We show below how this information is encoded in the Schottky uniformization and afterwards in the possible values of Tzz on the boundary torus, so how it is holographically encoded. Using the above information, we start with a torus, parametrized by a H/SL(2, Z), meaning we only x its moduli and do not dene a marking on it. So our torus is the quotient z z+1z+ (7.22)

for z C. We pick the canonical at metric dzd on C, which descends to z a metric on the torus. Since the metric has constant (zero) curvature, Tzz is 146

holomorphic by the above result. It turns out [19] that the only holomorphic quadratic dierentials on a torus are the constants (at least in the charts that are obtained from some domain in the universal covering C). This of course considerably restricts the possible Tzz . The way to choose a contractible cycle is by Schottky uniformizing the torus on P1 . To that end, we apply the coordinate transformation (w is the coordinate on P1 and z the coordinate on C): w = ez with = 2i(a + b)1 for some relatively prime a, b Z. The metric becomes: ds2 = and the identications (7.22) become: which is equivalent to if c, d Z solve adbc = 1. This stems from the fact that (7.26), combined with w e(a +b) w trivially, generates the two transformations (7.25) and vice versa. Of course, this is reminiscent of a relabeling of the generators of the fundamental group as we did in chapter 2 before, but this time in the coordinate w instead of z. So with this coordinate transformation, we have found both a Schottky uniformization (7.26) and a g(0) with 2 = |w|2 . Clearly determines our choice of contractible cycle, which is a + b, so the one that does not correspond to an element of the Schottky group generated by (7.26) but instead is trivial on P1 . We substitute 2 = |w|2 in (7.20) to obtain 1 2w2 so the Feerman-Graham metric becomes: 1 ||2 w d2 |dw + dw|2 ds2 = 2 + 4 |w|2 4 w Tww = w e(c +d) w (7.26) dwdw |w|2 (7.24) (7.23)

w e w e w

(7.25)

with the identication (7.26). Our result is in fact a bit more telling if we transform back to C using (7.23) again: ds2 = 1 2 2 d2 + |dz + d| . z 2 4 4 (7.27)

Of course, (7.23) is branched and thus has only a local inverse, but we can extend this metric with the identications (7.22) since we know it is a metric on the quotient. We emphasize that (7.27) and (7.22) together completely determine the solid torus, and that if 1 Tzz = 2 2 with determined by (7.24), then the metric glues together smoothly in the interior. We conclude that also the contractible cycle is holographically encoded by the value of Tzz on the boundary. 147

7.5.3

Boundaries with negative Euler number

If the boundary is of higher genus, the procedure is similar but unfortunately cannot be made as explicit. We can uniformize a Riemann surface S of genus g > 1 in H as H/F , where its conformal structure is completely determined by the Fuchsian group F . The additional advantage is that we then automatically have a metric of constant negative curvature on S, namely (1.9), which means that Tzz is holomorphic. For the above procedure to work, we need S to be Schottky uniformized, so S = (S )/S for some Schottky group S . Unfortunately, there is no explicit form available of a covering map from H to (S ). This means that we cannot easily transfer the metric of constant negative curvature to (S ). And things get even worse: if is a Schottky group of genus g > 1, then we do not know explicitly any metric on () that smoothly descends to S. In fact, it is easy to see that such a metric should blow up at all the limit points of S on P1 . We already mentioned that the structure of (S ) is a fractal, so probably the explicit form of a complete metric on () that descends to S is very hard if not impossible to nd. But let us come back to the covering map anyway and just dene it as J : H (S ) which is a locally biholomorphic surjective covering map which satises S J = J F . Note that J does not induce an isomorphism between S and F since it is not globally invertible. Remark. As for the torus, there are multiple choices for J that indicate which cycle we contract. It would be interesting to nd out at least how many there are and if they are related. In fact, it can be instructive to compare the results below also with the torus, for which we have (7.23) as the map from the universal covering to P1 , so in that case J = ez . Of course, the metric on C is at so there are also dierences. However, the end result that Tzz = S(J) remains. We again use z as the coordinate on the universal covering H and w as the coordinate on P1 , so w = J(z). Using J, the metric (1.9), which was given by: dzd z (Im z)2 becomes the metric: g(0) = dJ 1 dw
2

dwdw 1 (w))2 (Im J

(7.28)

where the notation J 1 is valid here since J is locally invertible and the dierent branches (since they are related by the isometries of F ) do not imply a dierent form of the metric. Note that this metric has curvature R(0) = 2 (note that J is locally just a coordinate transformation, so it does not change the curvature). With this metric, we obtain again from (7.20): Tww (w) = S(J 1 )(w) (7.29)

which is just the Schwarzian derivative of the local inverse of the covering map. This result is also obtained from a dierent viewpoint in [29]. Substituting 148

(7.28) and (7.29) into (7.12), the Feerman-Graham metric becomes: ds2 = with (w, w) =
1 (1 + 4 )2 dJ 1 d2 + 42 dw 2

|dw + dw|2 (Im J 1 (w))2


2

1 dJ 1 S(J 1 ) 1 2 1 + 4 dw

(Im J 1 (w))2

and periodicities dictated by S whose elements S act on the three-space as (, w) (, S (w)). We can use J 1 to locally transform back to H: ds2 = and we can use S(J 1 ) J = S(J) to obtain (z, z ) = (1 + 1 )2 |dz + d|2 d2 z 4 + 2 2 4 (Im z) dJ dz
2

(7.30)

1 2 1 S(J)(Im z) 2 1 + 4

from which we directly read o that Tzz = S(J). The metric (7.30) is easily seen, just as for the torus, to analytically continue to all of H. The covering group F in this coordinate system of course also acts on z only, so they too leave the slices of constant invariant. Again, the periodicities implied by F together with the metric (7.30) completely determine the three-manifold.

By now, our conclusion should be clear: if the boundary Riemann surface is described as a quotient of its universal covering C or H, and we pick 2 to be the canonical constant-curvature metric then the corresponding handlebody is smooth if Tzz = S(J) with J a covering map from the universal covering to a Schottky uniformization on P1 (in the charts from the universal covering). This result holds of course trivially for P1 as well.

7.6

Conclusion

In this chapter, we considered whether a holographic description of the handlebodies was possible. We have seen how 2 and Tzz are locally free but are constrained by global demands. In particular, we have obtained (7.20) as the constraint if the boundary Riemann surface is Schottky uniformized. Finally, we have shown that in the constant curvature metrics, (7.20) becomes just the Schwarzian of the covering map from the universal covering (P1 , C or H) to P1 . Our results are the basis for a more thorough investigation. Given the more or less explicit values of T we obtained, what more can we say about the dual conformal eld theory? Can we obtain higher-point functions as well? And maybe we can glue a Euclidean and a Lorentzian solution together along their 149

y = 0 slices; this might allow us to see how the eld theories on the dierent Lorentzian boundary components are correlated. These issues hopefully can be addressed in the future; we think that answering them would certainly improve our understanding of the relation between gravity and its description as a conformal eld theory.

150

Conclusion
We briey summarize our main conclusions and the open issues regarding our investigation of the wormholes. More details considering the various subjects can be found at the end of the previous chapters. In this report, we have investigated three-dimensional wormhole spacetimes that are solutions of the Einstein equations in 2+1 spacetime dimensions. The rst three chapters discussed the necessary preliminaries. Afterwards, we have shown in chapter 4 how these wormholes can be obtained as a quotient of some part of AdS3 . We considered the case where this quotient is taken with respect to an isometry group whose elements leave H invariant. The covering group then acts as a Fuchsian group on H and the Riemann surface obtained as H/ completely determines the corresponding spacetime. In this way, we have shown that for any non-exceptional Riemann surface we can obtain such a wormhole spacetime. If the Riemann surface is of type (g, n, m) then such a spacetime has m asymptotic regions that are always separated from the interior by a black and a white hole horizon. These horizons shield asymptotic observers from the nontrivial topology in the interior. Since the metric on the outer regions is the same for every wormhole, we concluded that in this setting it is impossible for asymptotic observers to obtain any information about the topology on the other side of the horizon. In chapter 5, we found coordinate systems for the toroidal wormhole that capture some of the physical data explicitly in the metric. On the other hand, we did not manage to obtain a valid description for all possible toroidal wormholes, so some part of the moduli space remained uncovered. Chapter 6 described the analytic continuation to Euclidean time. We described a new analytic continuation procedure and used it to obtain handlebodies as the Euclidean counterparts of our wormholes. However, there are also open issues regarding the validity of the procedure in the general case. Finally, in chapter 7 we discussed some aspects of holography. We found that the handlebodies can be described holographically by giving a Schottky uniformization of the boundary Riemann surface and saw that the same information was encoded in the stress-energy tensor of the dual conformal eld theory. In the Fuchsian uniformization the holomorphic part of this tensor is just the Schwarzian derivative of a covering map from H to P1 . 151

Outlook
We again summarize some of the more important open questions that are described in detail at the end of the corresponding chapters. First of all, we mention the issues of chapter 5. We should be able to use our description for any wormhole in the moduli space, but have not yet done so. Once this issue is covered, we probably can easily apply the splitting procedure to other wormholes and hence describe all wormholes with similar physical coordinate systems. Second, we have not yet described rotating wormholes. If we want to do so, we need to make the procedure of taking a quotient (of chapter 4) also explicit for quotients with respect to isometries that do not leave H invariant. Once we have a more explicit description, it would be interesting to see how we can cover them with similar coordinate patches as for the toroidal wormhole. Also, in order to analytically continue them to a Euclidean signature manifold, we should solve the issues surrounding the procedure of chapter 6. Third, we would like to capture the nontrivial topology of the wormholes in a dual conformal eld theory. Since this seems impossible in the classical Lorentzian picture, we probably need to add extra elds or strings to the spacetime. Another option would be to glue half a Euclidean solution to half a Lorentzian one, joined together at their y = 0 slice. It would be interesting to see what the CFT looks like in these cases. We conclude by recalling our original motivation as mentioned in the introduction, namely that we like to view this work as setting the stage for a quantum analysis. This means that we will eventually have to investigate issues such as temperatures, entropies and underlying microstates. Results in this direction would be very interesting, in particular if we can take them out of the three-dimensional arena and use them in higher-dimensional cases as well.

152

Acknowledgements
This work could not have been written without the support of many people. I would like to thank at least some of them here. First and foremost, this thesis is founded upon the ideas and support of my supervisor, Kostas Skenderis. I am deeply indebted to Kostas for his quick and solid answers to my frequent questions, for the thorough proofreading and most of all for his company during the project. I would also like to thank Robbert Dijkgraaf for his help on the occasional technical question. Also, I would like to acknowledge Wim and Beata for their help with the better-looking graphics and the cover design. I have always enjoyed the company of my fellow students, both Masters and PhD. Many thanks to Aron, Charles, Erik, Liza, Mark and Michelangelo for making me feel very welcome during my rst days at the ITFA. I am especially grateful to my former roommates Hendrik, Sheer and Dinesh. Naturally, I would also like to thank all the other members of the ITFA, especially Yocklang, Paula and Lotti, for their support and company. Finally, my deepest thanks go to my parents. Their constant support and encouragement have helped me throughout my studies and I feel the mere size of this work is another proof of their patience and kindness to me. Although I will never be able to thank them enough for everything they did, there can only be one appropriate start: I dedicate this thesis to my parents.

153

Bibliography
[1] William Abiko. The Real Analytic Theory of Teichmller Space. Springeru Verlag, 1980. [2] L.V. Ahlfors. Complex Analysis. McGraw-Hill, 1979. [3] L.V. Ahlfors and L. Sario. Riemann Surfaces. Princeton University Press, 1980. [4] Stefan Aminneborg, Ingemar Bengtsson, Dieter Brill, Soren Holst, and Peter Peldan. Black holes and wormholes in 2+1 dimensions. Class. Quant. Grav., 15:627644, 1998, gr-qc/9707036. [5] Stefan Aminneborg, Ingemar Bengtsson, and Soren Holst. A spinning antide Sitter wormhole. Class. Quant. Grav., 16:363382, 1999, gr-qc/9805028. [6] Michael T. Anderson. Geometric aspects of the AdS/CFT correspondence. 2004, hep-th/0403087. [7] M. Baados, M. Henneaux, C. Teitelboim, and J. Zanelli. Geometry of the n 2+1 black hole. Phys. Rev. D, 48:15061525, August 1993, gr-qc/9302012. [8] M. Baados, C. Teitelboim, and J. Zanelli. The black hole in threen dimensional space-time. Phys. Rev. Lett., 69:18491851, 1992, hepth/9204099. [9] Thierry Barbot. Causal properties of AdS-isometry groups. I: Causal actions and limit sets. 2005, math.gt/0509552. [10] Thierry Barbot. Causal properties of AdS-isometry groups. II: BTZ multi black-holes. 2005, math.gt/0510065. [11] Lipman Bers. Automorphic forms for Schottky groups. Advances in Mathematics, 16:332361, 1975. [12] D. Brill. 2+1-dimensional black holes with momentum and angular momentum. Annalen Phys., 9:217226, 2000, gr-qc/9912079. [13] Dieter Brill. Black holes and wormholes in 2+1 dimensions. 1998, grqc/9904083. [14] S. Carlip. Lectures on (2+1) dimensional gravity. J. Korean Phys. Soc., 28:S447S467, 1995, gr-qc/9503024. 155

[15] S. Carlip. Conformal eld theory, (2+1)-dimensional gravity, and the BTZ black hole. Class. Quant. Grav., 22:R85R124, 2005, gr-qc/0503022. [16] Steven Carlip and Claudio Teitelboim. Aspects of black hole quantum mechanics and thermodynamics in (2+1)-dimensions. Phys. Rev., D51:622 631, 1995, gr-qc/9405070. [17] Sean Carroll. Spacetime and geometry An Introduction to General Relativity. Addison Wesley, 2004. [18] Robbert Dijkgraaf, Juan M. Maldacena, Gregory W. Moore, and Erik P. Verlinde. A black hole farey tail. 2000, hep-th/0005003. [19] Hershel M. Farkas and Irwin Kra. Riemann Surfaces. Springer-Verlag, 1980. [20] Otto Forster. Lectures on Riemann Surfaces. Springer-Verlag, 1981. [21] S. S. Gubser, Igor R. Klebanov, and Alexander M. Polyakov. Gauge theory correlators from non-critical string theory. Phys. Lett., B428:105114, 1998, hep-th/9802109. [22] S. W. Hawking and Don N. Page. Thermodynamics of black holes in anti-de Sitter space. Commun. Math. Phys., 87:577, 1983. [23] S.W. Hawking and G.F.R. Ellis. The large scale structure of space-time. Cambridge University Press, 1973. [24] Seungjoon Hyun. U-duality between three and higher dimensional black holes. J. Korean Phys. Soc., 33:S532S536, 1998, hep-th/9704005. [25] Yoichi Imayoshi and Masahiko Taniguchi. An Introduction to Teichmnller u Spaces. Springer-Verlag, 1992. [26] Kirill Krasnov. Holography and Riemann surfaces. Adv. Theor. Math. Phys., 4:929979, 2000, hep-th/0005106. [27] Kirill Krasnov. Analytic continuation for asymptotically AdS 3D gravity. Class. Quant. Grav., 19:23992424, 2002, gr-qc/0111049. [28] Kirill Krasnov. Black hole thermodynamics and Riemann surfaces. Class. Quant. Grav., 20:22352250, 2003, gr-qc/0302073. [29] Kirill Krasnov. On holomorphic factorization in asymptotically AdS 3D gravity. Class. Quant. Grav., 20:40154042, 2003, hep-th/0109198. [30] Olli Lehto. Univalent Functions and Teichmller Spaces. Springer-Verlag, u 1986. [31] Juan M. Maldacena. The large N limit of superconformal eld theories and supergravity. Adv. Theor. Math. Phys., 2:231252, 1998, hep-th/9711200. [32] Juan M. Maldacena and Andrew Strominger. Ads(3) black holes and a stringy exclusion principle. JHEP, 12:005, 1998, hep-th/9804085. 156

[33] Georey Mess. Lorentz spacetimes of constant curvature. IHES preprint M/90/28, April 1990. [34] Subhashis Nag. The Complex Analytic Theory of Teichmller spaces. John u Wiley & Sons, 1988. [35] M. Rooman and Ph. Spindel. Uniqueness of the asymptotic ads(3) geometry. Class. Quant. Grav., 18:21172124, 2001, gr-qc/0011005. [36] Konstadinos Sfetsos and Kostas Skenderis. Microscopic derivation of the Bekenstein-Hawking entropy formula for non-extremal black holes. Nucl. Phys., B517:179204, 1998, hep-th/9711138. [37] Kostas Skenderis. Asymptotically anti-de Sitter spacetimes and their stress energy tensor. Int. J. Mod. Phys., A16:740749, 2001, hep-th/0010138. [38] Kostas Skenderis. Lecture notes on holographic renormalization. Class. Quant. Grav., 19:58495876, 2002, hep-th/0209067. [39] Kostas Skenderis and Sergey N. Solodukhin. Quantum eective action from the AdS/CFT correspondence. Phys. Lett., B472:316322, 2000, hepth/9910023. [40] George Springer. Introduction to Riemann Surfaces. Chelsea Publishing Company, 1957. [41] Leonard Susskind. The world as a hologram. J. Math. Phys., 36:63776396, 1995, hep-th/9409089. [42] Gerard t Hooft. Dimensional reduction in quantum gravity. 1993, grqc/9310026. [43] William P. Thurston. The Geometry and Topology of Three-Manifolds. Princeton University Press, 1980. Lecture notes, http://www.msri.org/publications/books/gt3m/. [44] Robert M. Wald. General Relativity. The University of Chicago Press, 1984. [45] Edward Witten. (2+1)-dimensional gravity as an exactly soluble system. Nucl. Phys., B311:4678, 1988. [46] Edward Witten. Anti-de sitter space and holography. Adv. Theor. Math. Phys., 2:253291, 1998, hep-th/9802150.

157

Das könnte Ihnen auch gefallen