Geometric Aspects of Gauge and Spacetime Symmetries Steffen Christian Martin Gielen Trinity College A dissertation submitted to the University of Cambridge for the degree of Doctor of Philosophy March 2011 2Declaration This dissertation is my own work and contains nothing which is the outcome of work done in collaboration with others, except as specified in the text (section 1.2). It is not substantially the same as any other thesis that I have submitted or will submit for a degree at this or any other university. This thesis is based on the research presented in the following refereed publica- tions: (1) Gary W. Gibbons and Steffen Gielen. The Petrov and Kaigorodov- Ozsva´th Solutions: Spacetime as a Group Manifold. Classical and Quantum Gravity 25 (2008), 165009. (2) Gary W. Gibbons and Steffen Gielen. Deformed General Relativity and Torsion. Classical and Quantum Gravity 26 (2009), 135005. (3) Gary W. Gibbons, Steffen Gielen, C. N. Pope, and Neil Turok. Measures on Mixing Angles. Physical Review D79 (2009), 013009. (4) Steffen Gielen. CP Violation Makes Left-Right Symmetric Extensions With Non-Hermitian Mass Matrices Appear Unnatural. Physical Review D81 (2010), 076003. (5) Steffen Gielen and Daniele Oriti. Classical general relativity as BF- Plebanski theory with linear constraints. Classical and Quantum Gravity 27 (2010), 185017. These papers appear in the text as references [65], [66], [67], [68], and [69] respec- tively. The results of section 5 of (3) were also separately published in (6) Gary W. Gibbons, Steffen Gielen, C. N. Pope, and Neil Turok. Naturalness of CP Violation in the Standard Model. Physical Review Letters 102 (2009), 121802. Steffen Gielen Thursday, 3 March 2011 3Acknowledgments My first and foremost thanks go to my supervisor, Prof. Gary Gibbons, for teaching me so much about physics and mathematics, for many fruitful discussions and sug- gestions, for his encouragement and for his support of all of my plans over the last three years. I have also greatly enjoyed working with, and learning from, Dr. Daniele Oriti, Prof. C. N. Pope and Dr. Neil Turok on different projects that are part of this thesis. Invitations by Dr. Turok and Dr. Oriti that allowed me to spend an intensive week at Perimeter Institute in October 2008 and several months at AEI Golm in 2009/10 deserve special thanks. I am grateful to Trinity College for all its support, both financial and otherwise, over the course of the last four years, without which I would not have been able to complete a PhD at Cambridge. My work has also been supported by an EPSRC studentship. I would like to thank my friends in Cambridge, who have hopefully improved my understanding of physics and made my time more enjoyable, not least through countless hours of ‘foosball’, and my friends in Germany who have kept in touch while I was away for so long, and gave support in difficult times. Going back a bit further in time, I would not have ended up as a PhD student in Cambridge (and certainly not at Trinity) without the encouragement of Dipl.-Phys. Igor Pikovski and the strong support of Profs. Ingo Peschel and Ju¨rgen Bosse. The latter two together with Hermann Schulz (Hochschuldozent) were also influential in shaping my interest in theoretical physics early on. Last but not least, I would like to thank my aunt Odina and uncle Uwe for their support, especially for the first year at Cambridge. 4Geometric Aspects of Gauge and Spacetime Symmetries Steffen Christian Martin Gielen Summary We investigate several problems in relativity and particle physics where symmetries play a central role; in all cases geometric properties of Lie groups and their quotients are related to physical effects. The first part is concerned with symmetries in gravity. We apply the theory of Lie group deformations to isometry groups of exact solutions in general relativity, relating the algebraic properties of these groups to physical properties of the spacetimes. We then make group deformation local, generalising deformed special relativity (DSR) by describing gravity as a gauge theory of the de Sitter group. We find that in our construction Minkowski space has a connection with torsion; physical effects of torsion seem to rule out the proposed framework as a viable theory. A third chapter discusses a formulation of gravity as a topological BF theory with added linear constraints that reduce the symmetries of the topological theory to those of general relativity. We discretise our constructions and compare to a similar construction by Plebanski which uses quadratic constraints. In the second part we study CP violation in the electroweak sector of the standard model and certain extensions of it. We quantify fine-tuning in the observed magnitude of CP violation by determining a natural measure on the space of CKM matrices, a double quotient of SU(3), introducing different possible choices and comparing their predictions for CP violation. While one generically faces a fine-tuning problem, in the standard model the problem is removed by a measure that incorporates the observed quark masses, which suggests a close relation between a mass hierarchy and suppression of CP violation. Going beyond the standard model by adding a left-right symmetry spoils the result, leaving us to conclude that such additional symmetries appear less natural. CONTENTS 5 Contents 1 Introduction 7 1.1 Lie Groups and Homogeneous Spaces . . . . . . . . . . . . . . . . . . 11 1.2 Structure of this Thesis and Conventions . . . . . . . . . . . . . . . . 15 I Deformed and Constrained Symmetries in Gravity 17 2 Deformations of Spacetime Symmetries 18 2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 2.2 Deformation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.3 The Petrov and Kaigorodov-Ozsva´th Solutions . . . . . . . . . . . . . 25 2.4 The Kaigorodov Solution and Lobatchevski Plane Gravitational Waves 42 2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 3 Deformed General Relativity and Torsion 50 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 3.2 Curved Momentum Space . . . . . . . . . . . . . . . . . . . . . . . . 55 3.3 Deformed Special Relativity . . . . . . . . . . . . . . . . . . . . . . . 58 3.4 A de Sitter Gauge Theory of Gravity . . . . . . . . . . . . . . . . . . 60 3.5 Cartan Connections on Homogeneous Bundles . . . . . . . . . . . . . 67 3.6 Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 3.7 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . 74 3.8 Gauge Invariance Broken? . . . . . . . . . . . . . . . . . . . . . . . . 77 3.9 Symmetric Affine Theory . . . . . . . . . . . . . . . . . . . . . . . . . 78 4 General Relativity From a Constrained Topological Theory 80 4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 4.2 Actions for Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 CONTENTS 6 4.3 Constraining BF Theory . . . . . . . . . . . . . . . . . . . . . . . . . 85 4.4 Linear Constraints for BF-Plebanski Theory . . . . . . . . . . . . . . 90 4.5 Lagrangian and Hamiltonian Formulation . . . . . . . . . . . . . . . . 101 4.6 Summary and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . 103 II Gauge Symmetries and CP Violation 106 5 SU(3) and its Quotients 107 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 5.2 The Homogeneous Space SU(3)/U(1)2 . . . . . . . . . . . . . . . . . 111 5.3 Metrics on U(1)2\SU(3)/U(1)2 . . . . . . . . . . . . . . . . . . . . . 112 5.4 Hermitian and Complex Matrices . . . . . . . . . . . . . . . . . . . . 114 6 Naturalness of CP Violation 120 6.1 Introduction and Motivation . . . . . . . . . . . . . . . . . . . . . . . 120 6.2 CP Violation in the Standard Model and Beyond . . . . . . . . . . . 124 6.3 Statistics on the Space of CKM Matrices . . . . . . . . . . . . . . . . 129 6.4 Statistics on the Space of Mass Matrices . . . . . . . . . . . . . . . . 136 6.5 Extension to Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . 151 6.6 Summary and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . 154 Summary and Outlook 157 References 161 List of Figures 176 List of Tables 177 Index 178 Introduction 7 Chapter 1 Introduction “Ich kann es nun einmal nicht lassen, in diesem Drama von Mathematik und Physik – die sich im Dunkeln befruchten, aber von Angesicht zu Angesicht so gerne einander verkennen und verleugnen – die Rolle des (wie ich genugsam erfuhr, oft unerwu¨nschten) Boten zu spielen.” [164] Hermann Weyl It is impossible to overemphasise the importance of symmetry as a fundamental concept in both mathematics and theoretical physics. In mathematics, the study of groups arose as an abstraction of “transformation groups”, invertible maps of a given “point field” into itself [164]. Groups are the mathematical realisation of the concept of symmetry. Felix Klein’s Erlangen Pro- gramme [96] aimed to describe any geometry by its transformation group; as we will detail shortly, the geometries considered in the Erlangen Programme are given by coset spaces of the form G/H = {gH, g ∈ G}, where G is a continuous (Lie) group and H a closed subgroup of G. Klein himself found the concept of groups most char- acteristic of 19th century mathematics [164]. The study of continuous symmetries is, in mathematical terms, the study of Lie groups and homogeneous spaces which nicely combines aspects of group theory and differential geometry. Beyond the direct study of groups, symmetry is an important concept in other parts of mathematics, such as the study of differential equations. In physics, the underlying symmetries of any theory are arguably its most fun- damental property. As an example, the first physical theory in the modern sense, Newtonian mechanics, fundamentally rests on assumptions about the symmetries of Introduction 8 the physical world: A configuration of bodies in (gravitational) interaction will behave in the same way if it is rotated as a whole (space is isotropic); a common motion of all bodies at constant velocity is not observable (relativity of inertial frames). Con- sequently, the laws of Newtonian mechanics are invariant under the action of the semidirect product of the rotation group SO(3) and commutative Galileian boosts, the Euclidean group E(3). When Einstein understood in 1905, in the wake of the Michelson-Morley experiment [118], that it was necessary to modify the concept of relativity of inertial frames, he did so by replacing the Euclidean group by the Lorentz group SO(3, 1) as the fundamental symmetry group. We will see in chapter 2 that this process can be mathematically viewed as a Lie group (or algebra) deformation. The Galilean and Lorentz groups are examples of spacetime symmetries; their primary action on space(time) itself induces an action on all fields according to their tensorial structure. In general relativity, these global spacetime symmetries are made local, although global symmetries also appear in many exact solutions of physical interest, as we shall see in chapter 2. Both global and local spacetime symmetries will play a role in the first part of this thesis. More generally, the equations defining a given theory (or the action, in the usual case that the equations of motion can be derived from an action principle) may also be invariant under a group of transformations of the physical variables which is not associated to a transformation of spacetime. A simple example in field theory is that of a complex scalar field φ with equations of motion invariant under a global rotation φ→ eiαφ. Promoting such a global symmetry to a local symmetry leads to a gauge theory. We will discuss gauge symmetries in both gravity and particle physics in this thesis. Mathematically, the fundamental structures describing local symmetries are prin- cipal fibre bundles over a given manifold which is interpreted as spacetime, where the structure group becomes the “gauge group” of the theory, and a connection on such a bundle is interpreted as a gauge potential. The remaining fields that make up the physical content of the theory then live in associated vector and tensor bundles. While this underlying structure suggests a very close relationship of all fundamental interactions in Nature, there are important differences between Yang-Mills theory, defined in terms of a principal bundle with compact gauge group, and general rela- tivity, where in addition to the connection one has a frame field as a second structure Introduction 9 which is crucial for the geometric interpretation of the theory.∗ In chapter 3, we will study a formulation of general relativity as a Yang-Mills-like theory which tries to exploit this analogy; the usual Lorentz algebra valued connection and the frame field are combined into a connection on a bundle of the Poincare´ group (or its deformation, the (anti-)de Sitter group). This study will elucidate some of the virtues as well as shortcomings of such a formulation. Symmetry is our most important guiding principle towards discovering the laws of Nature. In particle physics, the electroweak model [72, 146, 161] unifying the electromagnetic and weak interactions was discovered by postulating the fundamen- tal symmetries and writing down the most general action invariant under them. A crucial step in completing the present picture of particle physics was the proof by ’t Hooft and Veltman [156] that the mathematical structure of Yang-Mills theory is compatible with the framework of quantum field theory; namely, Yang-Mills theory is renormalisable. By far most attempts to go beyond the standard model of par- ticle physics, such as supersymmetric theories and GUTs, are based on postulating new symmetries. Part of our analysis of chapter 6 will be to discuss possible observ- able consequences of such an existence of additional symmetries beyond those of the electroweak model. General relativity, however, has so far evaded a similar treatment in terms of conventional field theory formulated on a background Minkowski spacetime, and by now it seems unlikely that adding (super)symmetries will be sufficient for a consistent quantum-mechanical treatment, although there has recently been increasing interest in the possiblity that maximal (N = 8) supergravity in four dimensions is perturba- tively finite [22]. String theory is one attempt to go beyond standard quantum field theory by replacing the notion of point particles by extended objects; symmetries (such as dualities) play a fundamental theory in string theory as well. Besides the symmetries that will be important for a consistent fundamental theory of quantum gravity, if we manage to construct (and understand) one, one may wonder whether at low energy some effects of quantum gravity will be manifest in a deformation of spacetime symmetries. This idea, which has been proposed on a purely phenomeno- logical basis [4] as well as a consequence of developments in non-perturbative quantum gravity [6], will be taken up in chapter 3. ∗A second difference is of course the non-compactness of the gauge group for Lorentzian signature, but this seems a less important difference to Yang-Mills theory in the basic structure of the theory – one often considers Riemannian signature as well. Introduction 10 The notion of symmetries we have discussed so far refers to invariance of the action or equations of motion under certain group actions. From a canonical view- point, there is another important class of symmetries, namely those generated by constraints. To obtain the canonical formulation of a given theory, one performs a Legendre transformation, leading to a Hamiltonian pi(x) ≡ δS δq˙i(x) , H ≡ ∫ dd−1xH(x) ≡ ∫ dd−1x ( pi(x)q˙ i(x)− L(x)) , (1.0.1) where S = ∫ dt ∫ dd−1xL(x) is the action defining the theory, and a symplectic struc- ture on the phase space parametrised by (qi, pi), {qi(x), pj(y)} = δij δd−1(x− y) , {qi(x), qj(y)} = {pi(x), pj(y)} = 0 . (1.0.2) The Hamiltonian density H(x) typically splits into a “dynamical” part which is a generic function of the pi and q i, and a “constraint” part of the form λαfα(pi, q i); here λα are Lagrange multipliers, variables that appear in the original action whose canonical momentum is identically zero and that are conventionally not treated as be- longing to the phase space. The functions fα(pi, q i) then act as constraints imposed on the dynamical variables. A subset of these will be first class†, i.e. their Pois- son brackets vanish up to a linear combination of constraints. The Poisson brackets {fα, ·} of the first class constraints then determine vector fields on the phase space whose integral curves are interpreted as describing physically equivalent configura- tions. Although not necessarily associated with connections on fibre bundles, the transformations on the phase space generated by these vector fields are, by their physical interpretation, usually interpreted as “gauge transformations” as well. We will see an instance of such a symmetry in the context of a topological field theory, which is then constrained by a term that is added to the action, in chapter 4. Although our discussions of chapter 4 will remain classical, the motivation for the reformulation of general relativity that we will consider comes from the fact that the quantisation of topological field theories is relatively well understood. One might therefore hope that starting from a topological quantum theory which is then constrained may provide a different route to quantum general relativity. This also provides a bridge to the discussion of chapter 3 in that both consider possible reformu- lations of general relativity at the classical level which are suggested by developments in non-perturbative approaches to quantising gravity. †We are here and in the following using the terminology introduced by Dirac [46]. Introduction 11 In all of our discussions, an important distinction is that between discrete and continuous symmetries. The first systematic study of groups in physics was probably that of discrete groups describing the symmetries of crystals [164]. We shall be mainly interested in continuous symmetries which are described by Lie groups. The spacetime and gauge symmetries described above are of this type. Their differentiable structure means that differential geometry can be used to study them. In particular, we will see that the group structure is encoded in a bracket operation on the tangent space to the group at the identity element. We will also encounter discrete symmetries, such as charge conjugation and parity in particle physics which are represented by the finite group Z2. These will not be studied mathematically, but their physical significance will be discussed. In the spirit of Hermann Weyl’s self-image as a “messenger” between mathemat- ics and physics, we will highlight different examples where a study of the differential geometry of (spacetime or gauge) symmetry groups leads to important physical con- sequences. Let us first introduce the most important aspects of the theory of Lie groups and homogeneous spaces that will be important throughout this thesis. 1.1 Lie Groups and Homogeneous Spaces The study of symmetries is the study of groups. If we consider continuous symme- tries, the corresponding groups have an additional differentiable structure: They are manifolds as well as groups‡. As simple examples one may think of the groups of isometries of Rn,Cn or Hn, denoted by O(n), U(n) and Sp(n) respectively, or more generally, of groups of matrices with continuous parameters which are submanifolds of Rn×n,Cn×n or Hn×n. All examples of relevance in this thesis will be of this type. Such groups are known as Lie groups, named after the Norwegian mathemati- cian Sophus Lie (1842-1899) who initiated their systematic study in the context of systems of differential equations. Lie himself seems to have referred to Lie groups as “continuous groups” which describes appropriately what they are. The study of Lie groups combines aspects of differential geometry with group theory. Two natural actions of a Lie group G on itself are given by the left and right ‡One also requires the group multiplication map (x, y) 7→ x · y and the inversion map x 7→ x−1 to be differentiable. Introduction 12 multiplication maps L : G×G→ G , (h, g) 7→ hg , R : G×G→ G , (h, g) 7→ gh . (1.1.1) By definition of a group these maps act simply transitively: Given a, g ∈ G there is a unique element h such that L(h, g) = a and a unique element h′ such that R(h′, g) = a. The pull-back and push-forward of these maps act on the tangent and cotangent bundle of G. Starting with a basis of the tangent space at the identity element e, one can use the push-forward of left translation to define a basis of vector fields on G, known as left-invariant vector fields. These define a global frame field on G and hence every Lie group is parallelisable. (Right-invariant vector fields can be defined analogously but often left invariance is preferred.) The tangent space at e, or alternatively the set of left-invariant vector fields on G, is known as the Lie algebra g. It inherits a Lie bracket [·, ·] from the commutator of vector fields. For matrix Lie groups, the bracket is given by the commutator of matrices. Lie’s important results are summarised in the following theorem [9]: Theorem 1.1.1 (Lie) (1) For any Lie algebra g there is a Lie group G (not necessarily unique) whose Lie algebra is g. (2) Let G be a Lie group with Lie algebra g. If H is a Lie subgroup of G with Lie algebra h, then h is a subalgebra of g. Conversely, for each Lie subalgebra h of g, there exists a unique connected Lie subgroup H of G which has h as its Lie algebra. Furthermore, normal subgroups of G correspond to ideals in g. (3) Let G1, G2 be Lie groups with corresponding Lie algebras g1, g2. Then if g1 and g2 are isomorphic as Lie algebras, then G1 and G2 are locally isomorphic. If the Lie groups G1, G2 are simply connected, then G1 is isomorphic to G2. Loosely speaking, all of the structure of the groupG is encoded in its Lie algebra. If one is not interested in the topological properties of G and considers only connected and simply connected Lie groups, one can study the geometry of G by studying g which is just a linear vector space together with a bracket structure. For matrix Lie groups, it is usually most convenient to fix a basis of generators Ma and to expand the Maurer-Cartan form in terms of this basis: g−1 dg = σaMa (1.1.2) Introduction 13 determines a basis of left-invariant one-forms σa. Similarly, right-invariant forms can be obtained by computing dg g−1 which is clearly invariant under g 7→ gh. The basis dual to {σa} is then a basis of left-invariant vector fields. By taking the exterior derivative of both sides of (1.1.2) and using d(g−1) = −g−1 dg g−1 for matrices, one can show that the forms σa satisfy the Maurer-Cartan relations dσa = −1 2 Cb a c σ b ∧ σc , (1.1.3) where Ca c b are the structure constants of the Lie algebra defined by [Ma,Mb] = Ca c bMc . (1.1.4) One defines left-invariant tensors accordingly, e.g. left-invariant metrics as generated by symmetrised combinations of σa: Any metric of the form g = habσ a ⊗ σb (1.1.5) for constant symmetric hab is by definition left-invariant. One can phrase this by saying that there is a one-to-one correspondence between scalar products on the Lie algebra g (specified by hab) and left-invariant metrics on G [9]. These metrics are the ones compatible with the left action on G on itself. Lie groups are in some sense special cases of a more general class of manifolds that will be important in the following: In many situations of physical interest, one considers a manifold X which a Lie group G acts on transitively, but not necessarily simply transitively. That is, for any x, x′ ∈ X, there is a (not necessarily unique) g ∈ G with g · x = x′. (Again we take this to be a left action, the case of a right action is analogous.) Then X is called a homogeneous space. For a given x ∈ X, the subgroup Gx = {g ∈ G : g · x = x} leaving x invariant is called the stabiliser, little group or isotropy subgroup of x. Then we have the following result [9] Proposition 1.1.2 Let G × X → X be a transitive action of a Lie group G on a manifold X, and let H = Gx be the isotropy subgroup of a point x. Then: (a) The subgroup H is a closed subgroup of G. (b) The natural map j : G/H → X given by j(gH) = g · x is a diffeomorphism. In other words, the orbit G · x is diffeomorphic to G/H. (c) The dimension of G/H is dimG− dimH. Introduction 14 Since we assume transitivity of the group action, the orbit G · x is the whole of X; one can then identify X with the coset space G/H.§ The choice of x in this identification is arbitrary and corresponds to a choice of origin in X; different choices for x lead to groups Gx related by conjugation. The natural left action of G on G/H is just induced by the group multiplication law. In many applications, an additional assumption is satisfied which simplifies the construction of metrics on a homogeneous space: A given homogeneous space G/H is called reductive if there is a splitting of Lie algebras of the form g = h⊕ x (1.1.6) invariant under the adjoint action of H. Then the subspace x of the Lie algebra g is isomorphic [9] to the tangent space to G/H at H (or any other coset gH). It therefore defines a basis of left-invariant vector fields on X, and one can take the dual basis of a basis for x to construct left-invariant metrics on X. Reductive geometries will be of importance in chapter 3 (with G = SO(d, 1) and H = SO(d − 1, 1)) and in chapters 5 and 6 (with G = SU(3) and H = U(1)2). We will encounter an example of a non-reductive geometry in section 2.4. Homogeneous spaces are the simplest generalisations of (flat) Euclidean geom- etry: In the 19th century, attempts to deduce Euclid’s fifth postulate, the parallel postulate [50], from the other four¶, led to the discovery of elliptic and hyperbolic geometries based on the sphere Sn ' SO(n + 1)/SO(n) and hyperbolic space Hn ' SO(n, 1)/SO(n), the natural counterparts of Euclidean space Rn ' E(n)/SO(n). In Lorentzian geometry, the spaces of constant curvature, Minkowski, de Sitter and anti-de Sitter space, are also homogeneous spaces. Many other known exact solutions in general relativity are also homogeneous spaces; see [154, chapter 12]. Homogeneous spaces appear whenever there are (gauge or spacetime) symmetry groups. Another example would be a Higgs potential spontaneously breaking a sym- metry group G down to a subgroup H; then the space of vacua would be G/H. §This result justifies dropping the distinction between the terms “coset space” and “homogeneous space” as we will do in the following. ¶About 30 such “proofs” were shown to be unsatisfactory in the 1763 dissertation of A. Kaestner’s student Georg Simon Klu¨gel entitled “Conatuum praecipuorum theoriam parallelarum demonstrandi recensio” (Critique of the foremost attempts to prove the theory of parallels) [97, p.274]. Introduction 15 1.2 Structure of this Thesis and Conventions The thesis consists of two main parts. The first part explores symmetries in relativity which are deformed or constrained. In chapter 2 we introduce deformation theory of Lie algebras and investigate its appli- cation to global symmetries (isometries) of certain exact solutions to Einstein’s equa- tions in four dimensions. We will see how the group-theoretic viewpoint of spacetimes in terms of their isometry groups relates to physical properties such as causality vio- lation. Section 2.3 contains calculations and results taken from [65] and are the result of a collaboration with Gary Gibbons, except subsection 2.3.6 which is new. We omit those parts of [65] which are of less relevance to the discussion of symmetry groups. In chapter 3 we turn to a study of local symmetries in relativity, namely those of the tangent space. We investigate how Cartan geometry, a generalisation of Rieman- nian geometry where the flat tangent space is replaced by an arbitrary homogeneous space, can be used to generalise the proposed deformed special relativity (DSR) to a “deformed general relativity” in the same way that (pseudo-)Riemannian geome- try describes general relativity as an extension of special relativity. We analyse the physical predictions of a particular example supposedly describing Minkowski space, and note that it involves a connection with torsion potentially leading to disastrous predictions of charge nonconservation. This analysis serves as another example of the interplay between an algebraic viewpoint on (infinitesimal) symmetries and physical effects. The material presented in chapter 3 is mainly taken from [66] which again resulted from collaboration with Gary Gibbons. In chapter 4 we study yet another possible reformulation of classical general relativity as a topological BF theory with constraints. Such a formulation has been considered before and is of relevance in covariant approaches to quantising gravity, but we give a new version of it where the constraints are linear, removing some of the ambiguity in their solutions. We see how a very large symmetry on the phase space of the theory – the local equivalence of all solutions to the equations of motion – can be constrained to give general relativity, a theory with local degrees of freedom. The results detailed in chapter 4, published in [69], were obtained in collaboration with Daniele Oriti. The second part is concerned with gauge symmetries in particle physics and their influence on CP violation in the electroweak sector. We aim to give estimates for the likelihood of the observed magnitude of CP violation, measured by the Jarlskog invariant J , by constructing a natural measure on the space of Cabibbo-Kobayashi- Introduction 16 Maskawa (CKM) matrices. This space is a double quotient of SU(3) and does not have a clearly preferred measure. After introducing different choices for the measure in section 5, motivated first only by geometric and then also by physical considerations – we will incorporate the observed quark masses into the analysis as well – in chapter 6 we make a detailed study of their physical predictions. We find that once the quark masses are assumed, predictions for CP violation agree well with observation. We then provide another example of how modified symmetries affect physical predictions by assuming an extended left-right symmetry given by a second SU(2) gauge group. This will influence our predictions on CP violation and lead to completely different results. The analysis for the standard model was done in collaboration with Gary Gibbons, Chris Pope and Neil Turok in [67]. Much of the calculations presented in sections 6.3 and 6.4 are taken from this paper; I have omitted some parts of [67] where my original contribution has been limited. The calculations in subsection 6.4.2 are my own and were published in [68]. We close with a brief summary and outline possible directions for future research. Our conventions for general relativity are those of Hawking & Ellis; the signature of a Lorentzian metric is mainly positive; Greek letters are used for spacetime and Latin letters for internal indices. The curvature of a connection is always F = dA+A ∧A without any coupling constants as customary in some of the particle physics literature. In discussions of CP violation in particle physics we follow the book of Jarlskog [91]. 17 Part I Deformed and Constrained Symmetries in Gravity Deformations of Spacetime Symmetries 18 Chapter 2 Deformations of Spacetime Symmetries 2.1 Overview The archetypal examples of groups describing spacetime symmetries in physics are the Galilean and Lorentz groups of classical mechanics and special relativity. The crucial insight of Einstein in 1905 was the necessity of replacing the Galilean group by the Lorentz group in order to have an invariant speed of light c; as we will see shortly, this process is also the archetypal example of a Lie algebra deformation. Historically, the development of a mathematical theory of deformation and contraction of Lie algebras was motivated by advances in theoretical physics, but it is at least conceivable that the replacement of the Galilean group by the Lorentz group had been suggested by mathematicians working on Lie group theory, instead of physicists like Einstein. Since physical theories are to some extent always idealised approximations of a much more complex Nature, the process of deformation of a symmetry group allows one to make statements about the stability of a given theory. This stability point of view was explored in detail in [116]. The passage from non-relativistic to special- relativistic mechanics can be understood as a deformation of an “unstable” to a “robust stable” theory [116]. The point here is that the zero commutators of Galilean boosts constitute “fine-tuning” and any perturbation of the commutation relations leads to noncommuting boosts. On the other hand, given noncommuting boosts, a slight perturbation of the coefficient appearing in the commutator can be undone by redefining the algebra generators, hence one has “robustness” in the theory. Deformations of Spacetime Symmetries 19 In [116], the possible central extension of the Poisson algebra of the phase-space coordinates of Hamiltonian classical mechanics, leading to quantum mechanics, was interpreted as a similar instability of the original algebra. It was then argued that although the resulting Heisenberg algebra admits a deformation, it can be undone by a nonlinear redefinition of the position coordinate. In this chapter, we shall focus on Lie algebra deformations, perturbations of the structure constants which can be expanded in a power series. We are maintaining the structure of a Lie algebra, i.e. of a vector space, and redefinitions of the generators will always be linear transformations. One could consider more general kinds of de- formations, where the Lie algebra structure is lost, either by allowing commutators of two generators to be nonlinear in the generators or by allowing “structure functions”. While the existence or non-existence of deformations of a given Lie algebra is a mathematically well-posed question, the physical interpretation of its answer is less clear in general. Both the generators and the deformation parameters of a given Lie algebra may have different possible interpretations. For example, while the Poincare´ group admits a deformation to the de Sitter or anti de Sitter groups, it was argued in [116] that the deformation parameter R, defining a length scale, may be taken to be infinity if one is only interested in local kinematics, which of course is what happens in general relativity. It is also clear that while linear transformations can formally be applied to the generators in the Lie algebra, it may not appear physically meaningful to take linear combinations of, for instance, translation and rotation generators in the Poincare´ algebra. This point was stressed in [37], who also pointed out that the interpretation of position operators appearing in a “kinematical algebra” acting on single- or multi-particle quantum states is physically rather unclear, since position is not additive in a way that momentum and angular momentum are. The classic cases of the de Sitter group and its contractions were studied in detail in [12], as we will review shortly. One could argue that the mathematical theory behind deformations seems to have predictive power in terms of extensions or generalisations of fundamental theories, but these are only a posteriori predictions [37]. There seem to be two obvious classes of new applications for the concept of deformations of spacetime symmetries: • Situations where the relevant symmetry group is a subgroup of the Poincare´ group, or more generally a group of lower dimension than the Poincare´ group. One example for the former is the proposed “very special relativity” [39], where Deformations of Spacetime Symmetries 20 the fundamental symmetry group is taken to be ISIM(n− 2), consisting of the maximal subgroup of the Lorentz group leaving a null direction invariant plus translations. The possibility of deforming this group to obtain “general very special relativity” was discussed in [62]; a physically acceptable one-parameter deformation was found, similar to the case of the Poincare´ group which can be deformed into the (anti-)de Sitter group. Interestingly enough, this deformed group turns out to be the symmetry group of an asymptotically AdS space- time playing a role in the construction of nonrelativistic hydrodynamics in the context of AdS/CFT: The “Schro¨dinger spacetime” [81] ds2 = r2 (−2 du dv − r2νdu2 + d~x2)+ dr2 r2 (2.1.1) can be viewed as a left-invariant metric on a group manifold which is a subgroup of the deformed group DISIMb(n− 2)[63, app. B]. On a less fundamental level, given a homogeneous space G/H with isometry group G, e.g. as a given solution of general relativity, one may always ask what deformations its isometry group admits, and try to construct spaces with the deformed symmetry. The two examples we give in this chapter describe this kind of situation. In principle, this could be used to facilitate the search for new solutions of general relativity, although one is restricted to homogeneous spaces, and curvature invariants are constant. One may still obtain spaces with interesting causal properties, as exemplified by the BTZ black hole [152] (obtained from anti-de Sitter space by periodic identifications), and by the closed timelike curves that we will see occur in the Petrov and Kaigorodov- Ozsva´th solutions. • Alternatively, one may assume that the symmetry group is larger than the Poincare´ group. In particular, the formalism can be straightforwardly extended to supergroups [52]. It was found in [52] that the Killing superalgebra of the M2- brane in eleven dimensions admits a deformation, which leads to the conjecture that a solution with the deformed Killing superalgebra exists, possibly having an interpretation as a perturbation of the M2-brane. Unfortunately, the search for such a solution does not seem to have been successful∗. Another example is Peter West’s conjecture of E11 as being the relevant sym- metry group for M-theory [162], where one could look for possible deformations. ∗Jose´ Figueroa-O’Farrill, private communication (May 2009) Deformations of Spacetime Symmetries 21 As outlined in section 1.1, the general setup for our considerations is a manifold X which a group of transformations G acts on transitively, which may then be identified with G/H where H is a subgroup of G. When using deformations to look for interesting manifolds “close to” a given one, one considers group deformations of the group G and then identifies the appropriate stabiliser H in the deformed group. It simplifies the interpretation of the manifold one obtains if this stabiliser can be viewed as a deformation of the stabiliser subgroup of the original undeformed group, but in general this need not be the case [62]. By definition, the deformation theory reviewed here can only make statements about Lie algebras. As explained in section 1.1, Lie’s theorems allow us to use the terms “Lie algebra” and “Lie group” interchangeably, meaning by “Lie group” the unique connected, simply connected group specified by a given algebra. In particular, terms such as “Poincare´ group” or “de Sitter group” refer to connected components of the identity. We review the mathematical theory of Lie algebra deformations in section 2.2, and then give two examples for pairs of solutions of general relativity related by such deformations: The Petrov and Kaigorodov-Ozsva´th solutions are discussed in section 2.3. We shall see that they are both geodesically complete, and try to relate their causal properties. We then turn to the Kaigorodov solution in section 2.4, and find it can be deformed into a solution known as a “Lobatchevski plane gravitational wave”. 2.2 Deformation Theory The theory of deformation of Lie algebras was reviewed in [104], where it was con- nected to the perhaps more familiar operation of Lie algebra (Inonu-Wigner [88]) contraction. A Lie algebra g is deformed by redefining the Lie brackets as a power series in a parameter t ft(a, b) = [a, b] + tF1(a, b) + t 2F2(a, b) + . . . , a, b ∈ g , (2.2.1) where the series is required to converge in some neighbourhood of the origin. The Jacobi identity then leads to integrability conditions on the functions Fi at each order in t: ∑ P(a,b,c) ∑ µ+ν=n Fµ(Fν(a, b), c) + Fν(Fµ(a, b), c) ! = 0 , (2.2.2) where the sum is over all cyclic permutations and n = 0, 1, 2, . . . Deformations of Spacetime Symmetries 22 Choosing n = 0 gives back the Jacobi identity for [·, ·], and at linear order (n = 1) one gets ∑ P(a,b,c) F1([a, b], c) + [F1(a, b), c] ! = 0 . (2.2.3) A deformation that only corresponds to a change of basis in the algebra g will be regarded as trivial; these are of the form f(a, b) = S[S−1a, S−1b] , S ∈ GL(n) , (2.2.4) and expanding S in a power series around the identity gives the general form of trivial deformations at each order in t. It is perhaps more common to express these operations in terms of structure constants Ca c b, as defined in (1.1.4) for a given basis. Denoting the structure constants of the deformed algebra, with respect to the same basis, by Cˆa c b, the Jacobi identity for the deformed algebra can be written as Cˆd e [aCˆb d c] = 0 . (2.2.5) One can then contemplate Rn 2(n−1)/2 (where n is the dimension of g) parametrised by coordinates Cˆa c b antisymmetric in the lower indices, and the submanifold of R n2(n−1)/2 defined by the Jacobi identity. A deformation of the given algebra g describes a smooth curve in this submanifold which can be expanded in a power series in a parameter t: Cˆa c b(t) = Ca c b + tAa c b + t 2Ba c b + . . . (2.2.6) At linear order in t the Jacobi identity (2.2.5) holds if Cd e [aAb d c] + Ad e [aCb d c] = 0 . (2.2.7) Then a linear deformation only gives rises to a deformation if the requirement (2.2.5) can be satisfied at each order in t. Under a change of basis in g, expressed as a matrix S, the structure constants will change according to Cˆa b c(t) = S b eCd e f (S −1)da(S −1)f c . (2.2.8) Expanding Sab(t) = δ a b+tM a b+. . ., this means that to first order a trivial deformation can be written as Aa b c =M b eCa e c − CebcM ea − CabeM ec . (2.2.9) Deformations of Spacetime Symmetries 23 One can rephrase these conditions in the language of differential forms by choosing a basis {λa} of left-invariant 1-forms for the original algebra satisfying the Maurer- Cartan relations dλa = −1 2 Cb a cλ b ∧ λc, and also defining Aa = 1 2 Ab a cλ b ∧ λc to be a vector-valued 2-form and Cab = Cc a bλ c to be a matrix-valued 1-form [62]. Here Cb a c are the structure constants of the original algebra and Ab a c is their infinitesimal deformation. Then at linear order, the Jacobi identity is DA := dA+ C ∧ A = 0 , (2.2.10) and a deformation that can be written as A = DS for some S = Sabλ b is regarded as trivial. Hence non-trivial infinitesimal deformations is in one-to-one correspondence with the cohomology group H2(g, g) [53]. At higher order one has to consider the higher cohomology groups. In general, if a non-trivial deformation has been found infinitesimally (i.e. at linear order), it does not necessarily extend to a deformation satisfying the full set of conditions (2.2.2); obstructions to “integrability” of infinites- imal deformations are given by the cohomology group H3(g, g) [53]. In practice, once all infinitesimal deformations have been found, one can verify the full Jacobi identity, and if it is violated, try to add higher order terms to the deformation. Using cohomology theory, one can determine all general deformations of a given algebra, as was done for the Galilean algebra, with and without central extension, in [53]. We will focus on deformations at linear order in the following. Applying this formalism, the authors of [12] classified all possible kinematical groups in four dimensions, by which they meant classifying ten-dimensional Lie al- gebras consisting of “translations”, “rotations”, and “inertial transformations”. This splitting meant singling out three generators (“rotations”) and fixing their commuta- tors with the other generators to assure isotropy of space. They further assumed that parity and time-reversal (defined on the generators) leave the Lie brackets invariant, and that the subgroup of inertial transformations is non-compact. The classification one then obtains consists of eight types of Lie algebras, all of which may be obtained by either contracting the de Sitter algebra so(4, 1) or anti-de Sitter algebra so(3, 2), or alternatively as deformations of the Abelian algebra, as shown in figure 2.1. The possible contractions correspond to different physical limits such as vanishing cosmo- logical constant, infinite speed of light or zero speed of light, as explained in detail in [12]. For instance, the Newton groups describe the symmetries of a non-relativistic cosmological model with nonzero cosmological constant; this was discussed in detail in [64]. Deformations of Spacetime Symmetries 24 (A)dS SO(3, 2), SO(4, 1)ffpara-Poincare´/inh. SO(4) ? ff ? Newton groupspara-Galilei     Poincare´Carroll ?? GalileiStatic (Abelian) ff ff Figure 2.1: Possible contractions and deformations of kinematical groups, according to [12]. Arrows denote possible Inonu-Wigner [88] contractions. Looking at figure 2.1, it could be argued that the introduction of noncommut- ing translations, the most direct effect of a curved spacetime, is very natural from the group deformation viewpoint as the de Sitter and anti-de Sitter groups arise as deformations of the Poincare´ group. For completeness, we mention that in order to obtain the algebra of quantum mechanics from a deformation of that of classical mechanics, as discussed in [116], one either deforms the Poisson algebra of functions on phase-space, replacing the Poisson bracket by a Moyal bracket {f, g}M = {f, g} − ~ 2 4 · 3! ∑ i1,i2,i3 j1,j2,j3 ωi1j1ωi2j2ωi3j3∂i1i2i3(f)∂j1j2j3(g) + . . . , (2.2.11) or alternatively considers the phase-space coordinates as elements of an Abelian Lie algebra which is replaced by the Heisenberg algebra. In the first case one has an infinite-dimensional algebra, in the second case one performs a central extension to- gether with a deformation. The very illuminating discussion in [116] used purely mathematical arguments as well as physical intuition to establish the stability of the Heisenberg algebra, considered as a test for the “robustness” of quantum mechan- ics. We stress again that this is a different notion of deformation, since one is now considering non-linear transformations of the generators. The maximal algebra one could consider in this discussion is the algebra of the Poincare´ group, together with “spacetime coordinates” and an “identity” (a central element) which appear as operators in quantum mechanics. In four dimensions, this Deformations of Spacetime Symmetries 25 is a 15-dimensional group which is naturally deformed into so(6 − t, t), where the original central element ceases to commute with the other generators in general. This maximally extended case is most obviously fraught with interpretational difficulties, as clearly illustrated in [37]. Formally, the deformation process naturally leads to non- commuting spacetime coordinates [116]. This proposal goes back to Snyder [151], whose algebra is however not the same as the one proposed in [116]. The idea of “spacetime being non-commutative” is very interesting physically, however since it is separate from the deformation theory of spacetime symmetries reviewed in this chapter, it will be discussed in more detail in chapter 3. 2.3 The Petrov and Kaigorodov-Ozsva´th Solutions Among the wide range of known exact solutions to Einstein’s field equations (the most comprehensive source is [154]), there are some with interesting symmetry properties. In this section we concentrate on the particularly simple case of a simply transitive group of motions, such that the stabiliser is trivial and there is a one-to-one correspon- dence between spacetime points and elements of the group of motions. Therefore the spacetime is not only a manifold but also a Lie group. One example, and in fact the only example among vacuum solutions without a cosmological constant, is provided by the Petrov solution (2.3.1). An analogous example with negative cosmological constant is provided by the Kaigorodov-Ozsva´th solution (2.3.10). The Petrov solution has a physical interpretation as the exterior solution of an infinite rigidly rotating cylinder. To gain an understanding of the physical proper- ties of the Kaigorodov-Ozsva´th solution, we compute the stress-energy tensor of the boundary theory in the context of the AdS/CFT correspondence from an expansion near the conformal boundary. We then give a matrix representation of the isometry group of the Petrov solution and its algebra and use it to construct left-invariant one-forms on the group which could be used to compute left-invariant metrics. We apply the theory of Lie algebra deformations to relate the Petrov solution to the Kaigorodov-Ozsva´th solution, and discuss the physical interpretation of this result. It is well known that the Petrov solution contains closed timelike curves (CTCs), so an obvious question is whether the Kaigorodov-Ozsva´th solution exhibits similarly exotic properties. Since in the context of possible causality violation by CTCs a central question is whether these can be created by some process in a spacetime which Deformations of Spacetime Symmetries 26 did not exhibit CTCs initially, we give a brief discussion of the possible appearance of CTCs by spinning up a rotating cylinder. This elementary example supports Hawking’s chronology protection conjecture which asserts that the appearance of CTCs is forbidden by the laws of physics [77]. In a more detailed analysis of the physical properties of the Petrov and Kaigorodov- Ozsva´th solutions, we show that they are geodesically complete. We discuss the global causality properties of the two spaces; while the Petrov solution is totally vicious, due to a theorem of Carter, a similar analysis for the Kaigorodov-Ozsva´th solution remains inconclusive. The computation of the stress-energy tensor in AdS/CFT however sug- gests causal pathologies in the Kaigorodov-Ozsva´th solution as well. 2.3.1 Vacuum Solutions With Simply-Transitive Groups of Motions Λ = 0 - The Petrov Solution The Petrov solution is introduced in [154] in the following theorem: The only vac- uum solution of Einstein’s equations admitting a simply-transitive four-dimensional maximal group of motions is given by k2ds2 = dr2 + e−2rdz2 + er(cos √ 3r(dφ2 − dt2)− 2 sin √ 3r dφ dt) , (2.3.1) where k is an arbitrary constant, which shall be set equal to one, and we have rela- belled the coordinates compared to [154]. The solution was first given in [134] and also discussed in [43]. It describes a hyperbolic plane H2 (the (r, z)-plane) with a timelike two-plane attached to each point. The isometry group is generated by the Killing vector fields T ≡ ∂ ∂t , Φ ≡ ∂ ∂φ , Z ≡ ∂ ∂z , R ≡ ∂ ∂r +z ∂ ∂z + 1 2 ( √ 3t−φ) ∂ ∂φ − 1 2 (t+ √ 3φ) ∂ ∂t , (2.3.2) which satisfy the algebra [R, T ] = 1 2 T − √ 3 2 Φ , [R,Φ] = 1 2 Φ + √ 3 2 T , [R,Z] = −Z . (2.3.3) The isometry group contains three-dimensional subgroups of Bianchi types I and VIIh acting on timelike hypersurfaces, and the solution (2.3.1) is Petrov type I [154]. The first three Killing vectors obviously generate translations while the action of the one- parameter subgroup generated by R on spacetime is given by the integral curves of Deformations of Spacetime Symmetries 27 R, solutions of dxµ(λ) dλ = Rµ(x(λ)) . (2.3.4) These integral curves have the form xµ(λ) = ( r0 + λ , z0e λ , φ0e −λ 2 cos √ 3 2 λ+ t0e −λ 2 sin √ 3 2 λ , −φ0e−λ2 sin √ 3 2 λ+ t0e −λ 2 cos √ 3 2 λ ) , (2.3.5) where we label the coordinates by xµ = (r, z, φ, t). The metric components gφφ and gtt become zero at certain values † of r, but as the determinant of the metric in (2.3.1) is always −1, it is possible to extend the coordinates to infinite ranges and the coordinates (r, z, φ, t) define a global chart. The manifold is also time-orientable, as the vector field tµ = (0, 0, sin √ 3 2 r, cos √ 3 2 r) defines a global arrow of time, though this amounts to ∂ ∂t being future- as well as past-directed at some points (as well as spacelike at others). Bonnor [25] pointed out that the solution can be viewed as a special case of the exterior part of a Lanczos-van Stockum solution [103, 155] describing an infinite cylinder of rigidly rotating dust. Since this allows a physical interpretation of (2.3.1), let us give the general solution for an infinite rigidly rotating dust cylinder, which in Weyl-Papapetrou form is given by [157] ds2 = H(ρ)(dρ2 + dz˜2) + L(ρ)dχ2 + 2M(ρ) dχ dτ − F (ρ)dτ 2, (2.3.6) where H, L, M, F are functions of the radial variable ρ containing two parameters a and R, interpreted as the angular velocity and radius of the cylinder respectively. The high energy case aR > 1 2 , which contains closed timelike curves (CTCs), is of interest here, and with the choices R = √ e and aR = 1 the exterior solution is given by H(ρ) = 1 ρ2 , L(ρ) = −2 √ e 3 ρ sin (√ 3 log ρ√ e ) , (2.3.7) M(ρ) = 2√ 3 ρ sin ( pi 3 + √ 3 log ρ√ e ) , F (ρ) = 2√ 3e ρ sin ( pi 3 − √ 3 log ρ√ e ) , where ρ is a radial coordinate in the exterior of the cylinder and so is restricted to ρ ≥ √e , (2.3.8) †these values of r can be shifted by a coordinate transformation φ→ αφ+βt, t→ −βφ+αt and hence have no coordinate-independent significance Deformations of Spacetime Symmetries 28 and χ is an angular coordinate and periodically identified with period 2pi; τ and z˜ are unconstrained. Applying the coordinate transformations z˜ = √ e z , ρ = √ e er , χ = 1 4 √ 3 √ 2e (√ 2 + √ 3z + √ 2− √ 3t ) , τ = 1 4 √ 3 (z − t) (2.3.9) to the line element (2.3.6) indeed gives back (2.3.1). Hence if we adopt the interpre- tation of (2.3.1) as describing the exterior of a spinning cylinder, we restrict the coor- dinates to r ≥ 0 and identify √ 2 + √ 3z + √ 2−√3t with √ 2 + √ 3z + √ 2−√3t+ 2pi 4 √ 3 √ 2e. The general Lanczos-van Stockum solution has three linearly independent Killing vectors ∂ ∂χ , ∂ ∂τ and ∂ ∂z but a fourth Killing vector is only present in the special case aR = 1 because the algebraic invariants of the Riemann tensor are independent of ρ just in this case [26]. Λ < 0 - The Kaigorodov-Ozsva´th Solution A solution of the vacuum Einstein equations with negative cosmological constant which has a simply-transitive four-dimensional group of motions was first given by Kaigorodov [93] and rediscovered by Ozsva´th [125]. It has the line element ds2 = − 3 Λ dr2 + e−2r(dz2 + 2 dt dφ) + e4rdφ2 − 2 √ 2er dz dφ . (2.3.10) This solution was also given in [154]. It is Petrov type III and the metric asymptoti- cally (as r → −∞) approaches that of anti-de Sitter space. Obvious Killing vectors are Z ≡ ∂ ∂z , Φ ≡ ∂ ∂φ , T ≡ ∂ ∂t , (2.3.11) and the metric (2.3.10) has a further isometry r → r + λ, z → eλz, φ→ e−2λφ, t→ e4λt (2.3.12) which is generated by the fourth Killing vector R ≡ ∂ ∂r + z ∂ ∂z − 2φ ∂ ∂φ + 4t ∂ ∂t . The Killing vector fields satisfy the algebra [R,Z] = −Z , [R,Φ] = 2Φ , [R, T ] = −4T . (2.3.13) No analogous solution for a positive cosmological constant exists [125, 154]. Because of the similarity to (2.3.3) one would expect (2.3.13) to arise as a deformation of (2.3.3). Physically both algebras describe the isometries of vacuum solutions of Einstein’s Deformations of Spacetime Symmetries 29 equations, one with Λ < 0 and one with Λ = 0, and hence one might expect one of them to arise as some limit of the other. This spacetime is also time-orientable, as the vector field tµ = (0,− 1√ 2 e3r,−1, 1) defines a global arrow of time. We can express (2.3.10) in coordinates corresponding to Poincare´ coordinates on AdS (with Λ = −3) ds2 = dρ2 + dz2 + dφ2 − dt2 ρ2 − 2ρ dz(dt+ dφ) + 1 2 ρ4(dt+ dφ)2 . (2.3.14) The limit ρ→ 0 in Poincare´ coordinates corresponds to the timelike boundary I of anti-de Sitter space. After setting ρ˜ = ρ2 the line element is ds2 = dρ˜2 4ρ˜2 + 1 ρ˜ ( dz2 + dφ2 − dt2 − 2ρ˜3/2 dz(dt+ dφ) + 1 2 ρ˜3(dt+ dφ)2 ) , (2.3.15) and (2.3.15) is an expansion of the form ds2 = dρ˜2 4ρ˜2 + 1 ρ˜ gµνdx µdxν , gµν(x, ρ˜) = g (0) µν (x) + g (2) µν (x)ρ˜+ g (3) µν (x)ρ˜ 3/2+ . . . (2.3.16) as given in [80]. The coefficient g(3) = −dz ⊗ (dt + dφ) − (dt + dφ) ⊗ dz encodes the stress energy tensor of the boundary dual theory in the context of the AdS/CFT correspondence [15, 80]. It does not satisfy even the null energy condition on the three- dimensional conformal boundary, since g (3) µν nµnν = −2 for the null vector n = ∂∂t + ∂∂z . This presumably reflects causal pathologies of the bulk spacetime that we will try to understand by more conventional means shortly. A general analysis of stationary cylindrically symmetric Einstein spaces was done in [109]. These authors assume the Lewis form of the metric, where a cross term dz dφ would be absent. We did not find it possible to bring (2.3.10) to the Lewis form. The theorem by Papapetrou [127] that any solution with two commuting Killing vectors (one timelike, one spacelike with periodic orbits) can be written in the Lewis form, only applies to solutions of the vacuum Einstein equations without cosmological term. Hence one cannot make a connection with spaces of the Lewis form as was possible for the Petrov solution. 2.3.2 Left-Invariant Forms Since the action of the elements of the group manifold on itself has been given in (2.3.5), we can write down a matrix representation of this group of motions, with a Deformations of Spacetime Symmetries 30 general element given by g =   1 0 0 0 r 0 er 0 0 z 0 0 e− r 2 cos( √ 3 2 r) e− r 2 sin( √ 3 2 r) φ 0 0 −e− r2 sin( √ 3 2 r) e− r 2 cos( √ 3 2 r) t 0 0 0 0 1   . (2.3.17) The group is generated by Z =   0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0   , Φ =   0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0   , (2.3.18) T =   0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0   , R =   0 0 0 0 1 0 1 0 0 0 0 0 −1 2 √ 3 2 0 0 0 − √ 3 2 −1 2 0 0 0 0 0 0   , (2.3.19) so that g = ezZ+φΦ+tT erR and (r, z, φ, t) are coordinates on the group. The generators satisfy the algebra [R, T ] = −1 2 T + √ 3 2 Φ , [R,Φ] = −1 2 Φ− √ 3 2 T , [R,Z] = Z . (2.3.20) This differs from the Killing algebra (2.3.3) by the usual overall minus sign coming from the fact that right-invariant vector fields generate left actions and vice versa. The matrix representation (2.3.17) gives the group multiplication law (r, z, φ, t) · (r′, z′, φ′, t′) = (r + r′, z + erz′, φ+ e− r2 (t′s+ φ′c), t+ e− r2 (t′c− φ′s)) , (2.3.21) where s ≡ sin √ 3 2 r, c ≡ cos √ 3 2 r. The Maurer-Cartan form is g−1dg = e−rR(Z dz + Φ dφ+ T dt)erR +Rdr ≡ Rλ1 + Z λ2 + Φλ3 + T λ4 , (2.3.22) which gives the desired basis of left-invariant one-forms: λ1 = dr , λ2 = e−r dz , λ3 = e r 2 ( cos (√ 3 2 r ) dφ− sin (√ 3 2 r ) dt ) , λ4 = e r 2 ( sin (√ 3 2 r ) dφ+ cos (√ 3 2 r ) dt ) . (2.3.23) Deformations of Spacetime Symmetries 31 We obtain a left-invariant metric on the group ds2 = ηabλ a ⊗ λb = dr2 + e−2rdz2 + er ( cos( √ 3r)(dφ2 − dt2)− 2 sin( √ 3r)dφ dt ) (2.3.24) with η = diag(1, 1, 1,−1), which is the same as (2.3.1) and shows that our chosen coordinates agree with the initial Petrov coordinates. We see how to recover a metric on a group manifold; note that one could obtain this metric by just starting from the algebra (2.3.3). 2.3.3 Deformations of the Petrov Killing Algebra We examine possible infinitesimal deformations of the four-dimensional Lie algebra. UsingMathematica for the computations, equations (2.2.7) give the following con- ditions on linear deformations: 0 = Az r φ + √ 3At r z = Az r t − √ 3Aφ r z = Aφ r t ; 0 = 2Aφ r r + Aφ z z + √ 3At z z = − √ 3Aφ z z + 2At r r + At z z = Aφ z t = Aφ r t ; 0 = Ar r z + 2Aφ φ z + √ 3Az t φ − √ 3At φ z = √ 3Ar r z + 2Az φ t − √ 3Az t t − √ 3Aφ φ z = √ 3Ar r φ + At φ φ − √ 3Aφ t t − Atrr = √ 3Az r φ − Atrz ; 0 = √ 3Ar r z − √ 3Az φ φ − 2Aztφ − √ 3At t z = Ar r z − √ 3Az φ t + 2At t z + √ 3Aφ t z = Ar r φ − √ 3Aφ φ t + √ 3At r r − Attφ = Azrφ + √ 3At r z . (2.3.25) These constraints for Aa c b reduce the number of free parameters from 24 to twelve. We list the most general deformation parameters satisfying (2.3.25) in a table: Table 2.1: Infinitesimal deformations of the Petrov Killing algebra. c = r c = z c = φ c = t Ar c z 2C x1 x2 x3 Ar c φ − √ 3A− B x4 x5 x6 Ar c t −A+ √ 3B x7 x8 x9 Az c φ 0 2B C √ 3C Az c t 0 2A − √ 3C C Aφ c t 0 0 − √ 3B − A B −√3A Deformations of Spacetime Symmetries 32 The parameters x1, x2, . . . , x9, A,B and C can be arbitrary real constants. We need to investigate which of these correspond to trivial deformations. The conditions (2.2.9) mean that trivial deformations can be written as Ar r z = −M rz , Arzz = −M rr , Arφz = − 3 2 Mφz+ √ 3 2 M tz , Ar t z = − √ 3 2 Mφz−3 2 M tz . (2.3.26) The parameters C, x1, x2 and x3 correspond to trivial deformations and can be set to zero by a change of basis. Furthermore, Ar r φ = 1 2 M rφ + √ 3 2 M rt , Ar r t = − √ 3 2 M rφ + 1 2 M rt , Az z φ = −M rφ , Azzt = −M rt (2.3.27) etc., so that we can set A = B = 0, Ar z φ = 3 2 M zφ + √ 3 2 M zt , Ar φ φ = − 1 2 M rr + √ 3 2 Mφt + √ 3 2 M tφ , Ar t φ = − √ 3 2 M rr − √ 3 2 Mφφ + √ 3 2 M tt , (2.3.28) so that we can set x4 = x5 = x6 = 0, Ar z t = 3 2 M zt − √ 3 2 M zφ , Ar φ t = √ 3 2 M rr − √ 3 2 Mφφ + √ 3 2 M tt , Ar t t = − 1 2 M rr − √ 3 2 Mφt − √ 3 2 M tφ , (2.3.29) so that we can set x7 = 0, but must treat x8 and x9 as nontrivial perturbations. After a relabelling of these parameters, the modified Lie algebra is now [R, T ] = −aT − bΦ , [R,Φ] = −1 2 Φ− √ 3 2 T , [R,Z] = Z . (2.3.30) These relations satisfy the full Jacobi identity and so the linear deformation indeed defines a deformation of the Lie algebra. We may modify the matrix representation by setting R =   0 0 0 0 1 0 1 0 0 0 0 0 −1 2 −b 0 0 0 − √ 3 2 −a 0 0 0 0 0 0   . (2.3.31) Deformations of Spacetime Symmetries 33 In the case where a 6= 1 2 , one can always find a linear transformation of the basis vectors Φ and T such that the algebra takes the more symmetric form [R, T ] = a′T + b′Φ , [R,Φ] = a′Φ± b′T , [R,Z] = Z , (2.3.32) with ± depending on the value of b in the original deformation. This means that there are three distinct cases: First case: Positive sign. Then a matrix representation of R is R =   0 0 0 0 1 0 1 0 0 0 0 0 a′ b′ 0 0 0 b′ a′ 0 0 0 0 0 0   (2.3.33) and a general group element looks like g = ezZ+φΦ+tT erR =   1 0 0 0 r 0 er 0 0 z 0 0 ea ′r cosh(b′r) ea ′r sinh(b′r) φ 0 0 ea ′r sinh(b′r) ea ′r cosh(b′r) t 0 0 0 0 1   . (2.3.34) The Maurer-Cartan form is g−1dg = Z e−r dz + Φ e−ar(cosh(b′r) dφ− sinh(b′r) dt) +T e−ar(cosh(b′r) dt− sinh(b′r) dφ) +Rdr (2.3.35) and we can read off the left-invariant forms. The special case a′ = 1, b′ = 3 gives the algebra of the Kaigorodov-Ozsva´th solution (2.3.10), as can be seen by setting Φ′ = Φ − T and T ′ = Φ + T , which amounts to [R,Φ] = Φ + 3T , [R, T ] = 3Φ + T ⇒ [R,Φ′] = −2Φ′ , [R, T ′] = 4T ′ , (2.3.36) which is just the sign-reversed version of (2.3.13). One can recover the metric (2.3.10) by choosing the symmetric matrix hab =   − 3 Λ 0 0 0 0 1 −√2 √2 0 −√2 −1 −1 0 √ 2 −1 3   (2.3.37) Deformations of Spacetime Symmetries 34 and computing the left-invariant metric habλ a⊗λb = − 3 Λ dr2+ e−2rdz2− 2e−2r(dφ2− dt2)+ e4r(dφ− dt)2− 2 √ 2erdz(dφ− dt) , (2.3.38) which after the coordinate transformations φ − t = φ′ and −φ − t = t′ reduces to (2.3.10). Since the matrix (2.3.37) has one negative and three positive eigenvalues, there exists a (vierbein) basis of left-invariant one-forms σa such that ηabσ a⊗σb gives the metric (2.3.10). Second case: Negative sign. R =   0 0 0 0 1 0 1 0 0 0 0 0 a′ b′ 0 0 0 −b′ a′ 0 0 0 0 0 0   , g = ezZ+φΦ+tT erR =   1 0 0 0 r 0 er 0 0 z 0 0 ea ′r cos(b′r) ea ′r sin(b′r) φ 0 0 −ea′r sin(b′r) ea′r cos(b′r) t 0 0 0 0 1   (2.3.39) The Maurer-Cartan form is g−1dg = Z e−rdz+Φ e−ar(cos(b′r)dφ−sin(b′r)dt)+T e−ar(cos(b′r)dt+sin(b′r)dφ)+Rdr (2.3.40) and a left-invariant metric will be given by ds2 = ηabλ a ⊗ λb = dr2 + e−2rdz2 + e−2a′r(cos(2b′r)(dφ2 − dt2)− 2 sin(2b′r)dφ dt) . (2.3.41) The original Petrov algebra is of course the special case a′ = −1 2 , b′ = √ 3 2 . For the metric (2.3.41) the Ricci tensor has non-vanishing components Rrr = −1− 2a′2 + 2b′2 , Rzz = −(1 + 2a′)e−2r , Rφφ = −(1 + 2a′)e−2a′r(a′ cos(2b′r) + b′ sin(2b′r)) , Rφt = (1 + 2a ′)e−2a ′r(a′ sin(2b′r)− b′ cos(2b′r)) , Rtt = (1 + 2a ′)e−2a ′r(a′ cos(2b′r) + b′ sin(2b′r)) . (2.3.42) The manifold is an Einstein manifold only if a′ = −1 2 and b′ = ± √ 3 2 (Λ = 0, Petrov solution) or a′ = 1 and b′ = 0 (Λ = −3, anti-de Sitter space). In the general case the energy-momentum tensor defined by Tµν = 1 8piG Gµν does not satisfy the weak energy condition; without loss of generality assume sin( √ 3r) = 0 and cos( √ 3r) = 1 and choose a timelike vector tµ = (0, 0, t3, t4) (t 2 4 ≥ t23), then Gµνt µtν = (1 + a′ + a′2 − b′2)(t23 − t24)− 2(1 + 2a′)b′t3t4 (2.3.43) Deformations of Spacetime Symmetries 35 can be made arbitrarily negative by letting t3, t4 → ±∞ while keeping t24 − t23 small and positive, unless a′ = −1 2 and b′2 ≥ 3 4 . In the case a′ = −1 2 and b′2 ≥ 3 4 , the Einstein tensor can be written as Gµν = −λgµν + 2λuµuν , λ ≡ b′2 − 3 4 ≥ 0, uµ = (1, 0, 0, 0) , uµuµ = 1 . (2.3.44) Note that uµ is spacelike. This tensor satisfies the dominant energy condition as 2trvr − tava ≥ 0 for any timelike and future-directed t, v. These statements are independent of the choice of the arrow of time, i.e. hold for both t4 < 0 or t4 > 0. Third case: a = 1 2 . Let us introduce a new parameter c, so that [R, T ] = −1 2 T − 2√ 3 c2Φ , [R,Φ] = −1 2 Φ− √ 3 2 T , [R,Z] = Z (2.3.45) for positive b which gives R =   0 0 0 0 1 0 1 0 0 0 0 0 −1 2 −2c2√ 3 0 0 0 − √ 3 2 −1 2 0 0 0 0 0 0   (2.3.46) and a general group element looks like g = ezZ+φΦ+tT erR =   1 0 0 0 r 0 er 0 0 z 0 0 e− r 2 cosh(cr) − 2c√ 3 e− r 2 sinh(cr) φ 0 0 − √ 3 2c e− r 2 sinh(cr) e− r 2 cosh(cr) t 0 0 0 0 1   . (2.3.47) The Maurer-Cartan form is g−1dg = Rdr + Z e−r dz + Φ ( e r 2 cosh(cr) dφ+ 2c√ 3 e r 2 sinh(cr) dt ) +T ( e r 2 cosh(cr) dt+ √ 3 2c e r 2 sinh(cr) dφ ) . (2.3.48) In the limit c→ 0 or b→ 0 this becomes g−1dg = Rdr + Z e−r dz + Φ e r 2 dφ+ T ( e r 2 dt+ √ 3 2 e r 2 r dφ ) . (2.3.49) Deformations of Spacetime Symmetries 36 The only remaining case, namely negative b, is [R, T ] = −1 2 T + 2√ 3 c2Φ , [R,Φ] = −1 2 Φ− √ 3 2 T , [R,Z] = −Z , (2.3.50) which turns the hyperbolic into trigonometric functions. If c 6= 0 one can rescale the coordinate t (for instance) and recover the same left- invariant forms as before, hence this does not give anything new. In the case c = 0 a left-invariant metric is given by ds2 = dr2 + e−2rdz2 + er (( 1− 3 4 r2 ) dφ2 − √ 3r dφ dt− dt2 ) . (2.3.51) The Einstein tensor for this metric can be written as Gµν = 3 16 gµν + 3 8 ( diag (−2, e−2r, 2er, 0)) µν . (2.3.52) This does not satisfy the weak energy condition as Gµνt µtν = − 3 16 (3+ er) < 0 for the timelike vector tµ = (1, 0, 0, 1). Of course, for all three possible cases considered, one could take any constant symmetric matrix with one negative and three positive eigenvalues instead of ηab to construct a left-invariant metric; this gives a 10-parameter family (subject to con- straints) of metrics whose Killing algebras contain one of the deformations of the four-dimensional Petrov Killing algebra as a subalgebra. One of these parameters can be absorbed into an overall scale, and any cross terms dr dz, dr dφ or dr dt can be absorbed into a redefinition of z, φ and t. One is left with a 6-parameter family of metrics, by far most of which will not appear to have a physical interpretation. We have chosen to just give examples; more efficient methods of constructing general metrics with a given isometry group are certainly known. 2.3.4 Spinning Cylinders In the case of an infinite rigidly rotating dust cylinder one could imagine trying to speed up this cylinder by shooting in particles with some angular momentum which enter the interior region on causal curves and increase the angular velocity a, so as to reach and surpass the critical value aR = 1 2 above which CTCs appear. We will show that this is not possible. The interior part of the general van Stockum solution ds2 = H(ρ)(dρ2 + dz2) + L(ρ)dφ2 + 2M(ρ)dφ dt− F (ρ)dt2 , (2.3.53) Deformations of Spacetime Symmetries 37 describing the region ρ < R is given by [157] H = exp(−a2ρ2) , L = ρ2(1− a2ρ2) , M = aρ2 , F = 1 . (2.3.54) As there are closed timelike curves for ρ > 1 a we require aR ≤ 1. The exterior solution is for a < 1 2 , from now on setting R = 1 for simplicity which is no loss of generality, H = e−a 2 ρ−2a 2 , L = ρ sinh(3+ θ) 2 sinh 2 cosh  , M = ρ sinh(+ θ) sinh 2 , F = ρ sinh(− θ) sinh  (2.3.55) where θ(ρ) = √ 1− 4a2 log ρ and  = Artanh√1− 4a2. Note that always−FL−M2 = −ρ2 and so the metric has the right signature for all ρ (this of course is also true as a→ 1 2 ). The point-particle Lagrangian is (a dot denotes differentiation with respect to an affine parameter λ) L = gµν x˙µx˙ν = H(ρ˙2 + z˙2) + Lφ˙2 + 2Mφ˙t˙− F t˙2 , (2.3.56) and since the Lagrangian does not depend on z, φ and t there are three conserved quantities associated with geodesics: P ≡ Hz˙ , J ≡ Lφ˙+Mt˙ , E ≡ F t˙−Mφ˙ . (2.3.57) The Lagrangian for timelike or null geodesics becomes L = Hρ˙2 + P 2 H + 1 ρ2 ( FJ2 − 2MEJ − LE2) = −µ2 ≤ 0 (2.3.58) and we obtain the radial equation( dρ dλ )2 = L Hρ2 ( −ρ 2µ2 L − ρ 2P 2 HL + E2 + 2M L EJ − F L J2 ) = L Hρ2 (E − V +eff (ρ))(E − V −eff (ρ)) , (2.3.59) where we have introduced an effective potential V ±eff (ρ) = M(ρ) L(ρ) J ± ρ √ 1 L(ρ) ( µ2 + P 2 H(ρ) + J2 L(ρ) ) . (2.3.60) This is well-defined for all ρ as H, L, M all remain positive for all ρ. A particle falling in on a geodesic can enter the cylinder if E > V +eff (1) = a 1− a2J + √ 1 1− a2 ( µ2 + P 2ea2 + J2 1− a2 ) ≥ a+ 1 1− a2J = 1 1− aJ . (2.3.61) Deformations of Spacetime Symmetries 38 Any particle entering the cylinder on a geodesic must have J E < 1 − a. In the limit a → 0 the conserved quantities J and E clearly describe angular momentum and energy per mass. We can identify the ratio J E with the angular velocity of an infalling particle at R = 1. If we are considering accelerated observers, equation (2.3.58) still holds, but P, E and J will no longer be conserved quantities. However, only the local values of these quantities at R = 1 will decide about whether or not a particle will be able to enter the interior region of the cylinder. This means that the above considerations also hold for accelerated observers and as any particle entering the cylinder must have J E < 1 2 for a = 1 2 , one cannot speed up the cylinder beyond a = 1 2 using particles on timelike or null curves. 2.3.5 Geodesic Completeness We ask whether the Petrov spacetime, with the radial coordinate r extended to take arbitrary values, is geodesically complete, i.e. whether all timelike and null geodesics can be extended to infinite values of the affine parameter. First we give an example that this need not be possible on a group manifold: Remove the null hyperplane z = t from Minkowski space and consider the half-space z > t, denoted byM−. It is clearly geodesically incomplete. Null translations and boosts (z, t)→ (z + c, t+ c) ; (t+ z, t− z)→ ( λ(t+ z), 1 λ (t− z) ) (2.3.62) act on M−, and together with translations (x, y) → (x + a, y + b) they form a four- dimensional group which acts simply-transitively on M−. For instance, the point (x, y, z, t) = (0, 0, 1, 0) is, by a null translation and a successive boost, taken to (0, 0, 1, 0)→ (0, 0, 1 + c, c)→ ( 0, 0, λc+ 1 2 − 1 2λ , λc+ 1 2 + 1 2λ ) . (2.3.63) There is a one-one correspondence between points in M− and group parameters (a, b, λ, c), where λ > 0. The space M− can be identified with the group G × R2, where G is the unique two-dimensional non-Abelian Lie group. As a second example, introduced in a slightly different context in [76], consider the dilatation group generated by translations and dilatations xµ → xµ + cµ , xµ → ρxµ , (2.3.64) Deformations of Spacetime Symmetries 39 where we denote the generators by Pa and D respectively. Parametrising the group elements by‡ g = ex aPaeλD, the Maurer-Cartan form is g−1dg = e−λdxa Pa + dλD (2.3.65) and hence a left-invariant metric is ds2 = dλ2 + e−2ληµνdxµdxν = 1 ρ2 (dρ2 + ηµνdx µdxν) , (2.3.66) where ρ = eλ. This is the metric of anti-de Sitter space in five dimensions in Poincare´ coordinates, which is geodesically incomplete as these coordinates cover only a patch of the full spacetime. Hence geodesic completeness is a non-trivial property of a group manifold. To show geodesic completeness of the Petrov solution, we need to show that no geodesic reaches infinity for finite values of the affine parameter. Consider the La- grangian L = gµν x˙µx˙ν , (2.3.67) which for the metric (2.3.1) is L = r˙2 + e−2rz˙2 + er ( cos √ 3r(φ˙2 − t˙2)− 2 φ˙ t˙ sin √ 3r ) . (2.3.68) Evidently, from the Euler-Lagrange equations, there are three conserved quantities because the Lagrangian does not depend on z, φ or t explicitly, the conjugate momenta P ≡ e−2rz˙ , j ≡ er(φ˙ cos √ 3r − t˙ sin √ 3r) , E ≡ er(φ˙ sin √ 3r + t˙ cos √ 3r) . (2.3.69) The Lagrangian now takes the form L = r˙2 + e2rP 2 + e−r ( cos √ 3r(j2 − E2) + 2Ej sin √ 3r ) , (2.3.70) and since the Lagrangian is a conserved quantity in geodesic motion the equation( dr dλ )2 = e−r(E2 cos √ 3r − 2Ej sin √ 3r − j2 cos √ 3r)− e2rP 2 − µ2 (2.3.71) is satisfied by any geodesic, where µ2 is positive for timelike, zero for null and negative for spacelike geodesics. For timelike geodesics (µ2 > 0), right-hand side of (2.3.71) ‡The coordinate λ only covers dilatations with ρ > 0, so that one may obtain geodesic complete- ness by adding the disconnected component with ρ < 0. This suggests that geodesic completeness of groups is a topological question that cannot be decided at the level of the Lie algebra. Deformations of Spacetime Symmetries 40 becomes negative for large r, so that r is bounded and we may extend geodesics infinitely. For null geodesics, r˙ is bounded§. So for any finite values of the affine parameter, r˙ remains finite and so do z˙, φ˙ and t˙. Hence the Petrov spacetime is geodesically complete. It will be incomplete if we interpret it as the exterior solution of a rotating cylinder and cut off the region described by the Petrov solution at some value of r. By very similar arguments we can show geodesic completeness of the Kaigorodov- Ozsva´th solution with line element (2.3.14). In this case, the Lagrangian is L = 1 ρ2 ( ρ˙2 + z˙2 + φ˙2 − t˙2 ) − 2ρ z˙(t˙+ φ˙) + 1 2 ρ4(t˙+ φ˙)2 (2.3.72) and the conserved quantities are P ≡ 1 ρ2 z˙− ρ(t˙+ φ˙), j ≡ 1 ρ2 φ˙− ρz˙+ 1 2 ρ4(t˙+ φ˙), E ≡ 1 ρ2 t˙+ ρz˙− 1 2 ρ4(t˙+ φ˙) (2.3.73) so that the radial equation is( dρ dλ )2 = −ρ4(P 2 + j2 − E2)− 2ρ7P (j + E)− 1 2 ρ10(j + E)2 − ρ2µ2 . (2.3.74) The right-hand side becomes negative if µ2 > 0 for both ρ→∞ and ρ→ 0, hence all timelike geodesics are bounded in ρ. For µ = 0 the right-hand side is either constant zero or becomes negative for large r, so null geodesics are bounded from above in ρ. For very small ρ, ρ˙ goes to zero, so as before ρ˙ is bounded for null geodesics. This shows geodesic completeness. Note that in the case of anti-de Sitter space (2.3.74) would be ( dρ dλ )2 = −ρ4(P 2 + j2 − E2)− ρ2µ2 (2.3.75) and depending on the magnitudes of E, j and P the right-hand side blows up as ρ→∞ for some geodesics, which can reach ρ =∞ in finite affine parameter distance. 2.3.6 Causal Properties It was observed by Tipler [157] that the Petrov solution contains CTCs. As shown in [158], its causal behaviour is actually even more pathological: The Petrov solution §If we allow r to take negative values, then for E 6= 0 or j 6= 0 the right-hand side becomes oscillatory for large negative r, taking positive as well as negative values. Hence both timelike and null geodesics are bounded from below in r. For E = j = 0 the right-hand side is either constant zero or always negative. Deformations of Spacetime Symmetries 41 is totally vicious (for a summary of causality conditions see [59]), i.e. for any two points a and b one has a < b and b < a, where “<” is the usual causal relation “can be joined by a timelike curve”. The proof of this statement proceeds by giving the future and past light cones at the identity element of the group manifold, and the observation that the adjoint action of elements of the form exp(λR) rotates the cone around, such that for certain values of λ a future-directed timelike tangent vector is mapped to a past-directed timelike tangent vector. One might hope that a similar statement could be made for the Kaigorodov- Ozsva´th solution, but this turns out not to be case, as is apparent from the deformed group element (2.3.34) or the commutation relations (2.3.13). The action of R on the plane generated by Φ and T is now just a scaling of the generators, which amounts to an opening and closing of the light cones, but no rotation. The assumptions of Theorem 2 in [158] are not satisfied. Apart from the results in [158] which refer to homogeneous spacetimes, we may try to apply a criterion for the causality properties of spacetimes with an Abelian isometry group acting transitively on timelike surfaces which was given by Carter in his 1967 PhD thesis [34]. For the Kaigorodov-Ozsva´th solution, this Abelian isometry group is generated by ∂ ∂z , ∂ ∂φ and ∂ ∂t which act on surfaces {r = constant}. Carter’s result is that a sufficient condition for such a spacetime to be totally vicious is the nonexistence of a one-form in the dual of the Abelian Killing algebra which is everywhere timelike or null. In the present case, such a one-form is of the form ω = Adz+B dφ+C dt, with A, B and C constants. Conversely, if there is such a one-form which is everywhere timelike, the spacetime is virtuous, i.e. causally well-behaved. In the present case, one finds that the norm of such an ω is g˜(ω, ω) = e2r ( A2 + 2 √ 2AC e3r + C(2B + Ce6r) ) , (2.3.76) where g˜ is the contravariant version of (2.3.10). In order to make (2.3.76) non-positive for all values of r, one has to choose A = C = 0 to be left with the null one- form dφ. Put differently, the subgroup generated by ∂ ∂z and ∂ ∂t acts on null surfaces {r = φ = constant}; were these spacelike it would follow that the Kaigorodov-Ozsva´th solution is causal (cf. Proposition 9 in [33]). Carter’s condition for causality violation was used by Tipler in [157]¶ to show the pathological behaviour of the Petrov solution. Here the situation is less clear; we ¶whose wording is slightly unfortunate, since he states the relevant condition on ω as being “everywhere timelike” Deformations of Spacetime Symmetries 42 cannot show directly that in our example spacetimes related by Lie algebra defor- mation have similar causal properties. The strongest indication we have for causal pathologies of the Kaigorodov-Ozsva´th spacetime comes from the computation of the boundary stress-energy tensor in AdS/CFT in (2.3.16). 2.4 The Kaigorodov Solution and Lobatchevski Plane Gravitational Waves As a different example of a homogeneous space which appears as an exact solution in general relativity, we consider the Kaigorodov solution, which has a well-known inter- pretation as the AdS analogue of homogeneous pp-waves on a Minkowski background [139]. It has received recent interest in the context of string theory as the near-horizon limit of an M2 brane [42, 129]. By using Lie algebra deformations we again try to find a spacetime which has a physical interpretation “close to” the Kaigorodov solu- tion. An obvious candidate would be a member of the family of “Lobatchevski plane gravitational waves” discussed by Siklos in [149] as spacetimes admitting a Killing spinor. The expectation that the particular solution considered by Siklos which has a five-dimensional isometry group arises as a deformation of the Kaigorodov solution will be confirmed. Since the isometry group is now five-dimensional and not four- dimensional as in the previous section, one has to consider a quotient G/H. The larger symmetry will place constraints on the general form of a left-invariant metric. 2.4.1 The Kaigorodov Solution The Kaigorodov solution (of Petrov type N) [93] is another example of a homo- geneous space-time which solves the vacuum Einstein equations with a cosmological constant Λ < 0. The form given in [154] is ds2 = −12 Λ dz2 + 10k e2zdx2 + e−4zdy2 − 10u ezdz dx− 2ez du dx . (2.4.1) The five-dimensional Killing algebra of (2.4.1) is spanned by Y ≡ ∂ ∂y , X ≡ ∂ ∂x , U ≡ e−5z ∂ ∂u , Z ≡ ∂ ∂z + 2y ∂ ∂y − x ∂ ∂x , W ≡ ye−5z ∂ ∂u + x ∂ ∂y . (2.4.2) Deformations of Spacetime Symmetries 43 The stabiliser of the origin is the one-parameter subgroup generated by W , which we call H to be consistent with previous notation. The transformations generated by Z are z → z + λ, y → ye2λ, x→ xe−λ , (2.4.3) while W generates y → y + λx, u→ u+ ye−5zλ+ 1 2 xe−5zλ2 . (2.4.4) The Killing vectors have non-zero commutators [W,Y ] = −U , [W,X] = −Y , [Z,U ] = −5U , [Z, Y ] = −2Y , [Z,X] = X , [W,Z] = 3W ; (2.4.5) Y,X and U span a three-dimensional Abelian subalgebra. The generators of the kinematical group will have the sign-reversed algebra [W,Y ] = U , [W,X] = Y , [Z,U ] = 5U , [Z, Y ] = 2Y , [Z,X] = −X , [W,Z] = −3W . (2.4.6) We denote the isometry group by K ‖; it has three-dimensional subgroups of Bianchi types I, II and VIh. The spacetime is naturally represented as a coset space M = K/H . (2.4.7) Represented as matrices, the group K is generated by X =   0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0   , Y =   0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0   , Z =   5 0 0 0 0 0 0 0 0 1 0 0 2 0 0 0 0 0 −1 0 0 0 0 0 0   , U =   0 0 0 0 −1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0   , W =   0 0 −1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0   . (2.4.8) ‖The K stands for Kaigorodov or Killing, according to taste Deformations of Spacetime Symmetries 44 Any element of the group can be parametrised by g = exX+yY+ue 5zUezZewW , (2.4.9) such that (x, y, z, u) are coordinates on the coset spaceM. The Maurer-Cartan form ωMC ≡ g−1dg = eaλa , (2.4.10) where ea are the group generators, gives a basis of left-invariant one-forms on K: λ1 = ez dx , λ2 = e−2z dy − w ez dx , λ3 = dz , λ4 = du− w e−2z dy + 1 2 w2 ez dx+ 5u dz , λ5 = dw + 3w dz . (2.4.11) As usual, one can now construct a left-invariant metric on the group by ds2 = Gabλ aλb , (2.4.12) where Gab is constant and non-degenerate. Any such metric will depend on the coordinate w on W , since the Lie algebra splitting k = m⊕ h (2.4.13) is not reductive: [W,Z] = 3W where one would require [h,m] ⊂ m. (Recall the definition of a reductive geometry given in (1.1.6)). The original Kaigorodov metric can be reconstructed by taking ds2 = 10k(λ1)2 + (λ2)2 − 12 Λ (λ3)2 − 2λ1λ4 + (λ5)2 (2.4.14) = 10k e2zdx2 + e−4z dy2 − 12 Λ dz2 − 2ez dx(du+ 5u dz) + (dw + 3w dz)2 and reducing to the coset by dropping the last term. We summarise the most important physical properties of this solution, as discussed in more detail in [139]: The Kaigorodov spacetime has a curvature singularity at z = ∞, and z = −∞ represents null and spacelike infinity; the metric approaches the AdS metric asymptotically as z → −∞. The patch covered by the coordinates (x, y, z, u) can be extended by changing coordinates to q = e2z, and extending to negative q. The enlarged spacetime splits into two disjoint regions q > 0 and q < 0 which cannot mutually communicate. Then X = ∂ ∂x is a timelike Killing vector for q < 0, and the enlarged spacetime is stationary in this region. This means that gravitational radiation can exist in a stationary spacetime. Deformations of Spacetime Symmetries 45 2.4.2 Deformations of the Lie Algebra In order to find infinitesimal deformations of the Killing algebra (2.4.6), we again consider the Jacobi identity at linear order (2.2.7), again using Mathematica for explicit computations. Imposing these constraints on Aa c b, one is left with 20 free parameters, as shown in table 2.2. Table 2.2: Infinitesimal deformations of the Kaigorodov Killing algebra. c = x c = y c = z c = u c = w Ax c y (2+ η)/6 (2δ + 5τ)/3 0 −γ/4 0 Ax c z −ω + 2P 3α + β −(δ + τ)/3 c2 γ Ax c u σ/3 η/2 0 5/3(δ + τ) 0 Ax c w (c9 − 2c8 − 5c7)/5 c1 (2+ η)/6 α δ Ay c z (6c9 − 7c8 − 15c7)/5 −ω + P (2+ η)/3 β τ Ay c u 0 −2σ/3 0 (5+ 4η)/3 0 Ay c w m c7 σ/3 c5 ε Az c u −6m c8 −5σ/3 ω η Az c w c3 c4 (3c9 − c8)/5 c6 P Au c w 0 −m 0 c9 σ Again it turns out that almost all of the possible deformations should be considered trivial (i.e. correspond to a change of basis in the Lie algebra). Only a one-parameter deformation, which can be given by ω, cannot be obtained by a basis transformation. This infinitesimal deformation satisfies the full Jacobi identity, and we obtain the new algebra [W,Y ] = U , [W,X] = Y , [Z,U ] = (5 + α)U , [Z, Y ] = (2 + α)Y , [Z,X] = (α− 1)X , [W,Z] = −3W . (2.4.15) One would perhaps have expected that at least one possible deformation would be to deform the commutation relations of Z to arbitrary numbers: [Z,U ] = aU , [Z, Y ] = bY , [Z,X] = cX , [W,Z] = dW . (2.4.16) The one-parameter deformation we have obtained is the most general deformation of this type: By a rescaling of Z one can always keep one of the four numbers a, b, c, d Deformations of Spacetime Symmetries 46 fixed. Then the Jacobi identity puts two constraints on changing the coefficients in the three other commutators, leaving a single deformation parameter. Put differently, one is free to replace the deformed algebra (2.4.15) by any one- parameter deformation of the action of Z which satisfies the Jacobi identity, e.g. [W,Y ] = U , [W,X] = Y , [Z,U ] = 5U , [Z, Y ] = (2 + α)Y , [Z,X] = (2α− 1)X , [W,Z] = (α− 3)W . (2.4.17) We will use this last form in the following. In the matrix representation we only have to modify the matrix for Z: Zα =   5 0 0 0 0 0 0 0 0 1 0 0 2 + α 0 0 0 0 0 2α− 1 0 0 0 0 0 0   . (2.4.18) 2.4.3 Construction of the Manifold We are trying to construct a metric on the quotient space M′ = K′/H , (2.4.19) where the stabiliser H is again generated by W . We use the same parametrisation as before: g = exX+yY+ue 5zUezZαewW (2.4.20) By computing the Maurer-Cartan form one finds the left-invariant forms are now λ1 = e(1−2α)z dx , λ2 = e−(2+α)z dy − w e(1−2α)z dx , λ3 = dz , (2.4.21) λ4 = du− w e−(2+α)z dy + 1 2 w2 e(1−2α)z dx+ 5u dz , λ5 = dw + (3− α)w dz . Again, it is not possible to find a left-invariant metric ds2 = Gabλ aλb (2.4.22) which is independent of w. We can, however write any such metric in the form ds2 = G55(dm+ Aµdx µ)2 + G˜µνdx µdxν , (2.4.23) Deformations of Spacetime Symmetries 47 where xµ = (x, y, z, u), and demand that G˜µν , in order to be a metric on the coset space, be independent of w. This reduces the number of free parameters in G from 15 to nine, and the coset metric is (setting G55, which determines the overall scale, to −1) ds2 = p3e z(2−4α) dx2 + p1e−2z(2+α) dy2 + p2dz2 − 2ez(1−2α) dx(p1du+ (5 p1 u− p4)dz) , (2.4.24) where we have defined p1 := G 2 25−G22G55 , p2 = G235−G33G55 , p3 = G215−G11G55 , p4 = G15G35−G13G55 . (2.4.25) It is clear that by shifting u by a constant one can always achieve that p4 = 0. This means the most general metric on the deformed coset space is ds2 = p3e z(2−4α) dx2 + p1e−2z(2+α) dy2 + p2dz2 − 2p1 ez(1−2α) dx(du+ 5u dz) . (2.4.26) One discovers that the determinant of (2.4.26) is det g = −e−2(1+3α)zp31p2 , (2.4.27) and one finds that for Lorentzian signature (−+++) we must have p1 > 0 and p2 > 0. One can verify that the metric, as it should be, has five Killing vectors Y ≡ ∂ ∂y , X ≡ ∂ ∂x , U ≡ e−5z ∂ ∂u , Zα ≡ ∂ ∂z + (2 + α)y ∂ ∂y + (2α− 1)x ∂ ∂x , W ≡ ye−5z ∂ ∂u + x ∂ ∂y , (2.4.28) which satisfy the algebra [W,Y ] = −U , [W,X] = −Y , [Zα, U ] = −5U , [Zα, Y ] = −(2 + α)Y , [Zα, X] = (1− 2α)X , [W,Zα] = (3− α)W . (2.4.29) One finds that the Ricci tensor of (2.4.26) is, in the coordinate basis, Rµν = −3(2 + α) 2 p2 gµν − 5e (2−4α)zp3α(α− 3) p2 δµ xδν x . (2.4.30) There is a cosmological term with Λ = −3(2 + α) 2 p2 , (2.4.31) Deformations of Spacetime Symmetries 48 consistent with p2 = −12/Λ for the original Kaigorodov solution, and a traceless second term which describes null dust. In an orthonormal frame, given by E0 = du+ ez(1−2α) ( − p3 2p1 + 25p1u 2 2p2 + p1 2 ) dx , E1 = du+ ez(1−2α) ( − p3 2p1 + 25p1u 2 2p2 − p1 2 ) dx , E2 = √ p2 dz − 5p1√ p2 u ez(1−2α)dx , E3 = √ p1e −z(2+α)dy , (2.4.32) the nonvanishing components of T ab are T 00 = T 01 = T 10 = T 11 = 5p3α(3− α) p2p21 . (2.4.33) The number (2.4.33) is chosen to be positive so that T ab describes an electromagnetic field; then depending on the value of α the sign of p3 has to be chosen appropriately. Using deformations of the isometry group of the Kaigorodov solution we have constructed a spacetime which belongs to a more general class of “Lobatchevski plane gravitational waves” discussed in [149]. To bring the metric (2.4.26) into the standard form we apply the coordinate transformation q = √ p2 2 + α e(2+α)z , y˜ = √ p1y , u˜ = −p1e5zu . (2.4.34) In these coordinates the metric becomes ds2 = p2 (2 + α)2q2 [ dq2 + dy˜2 + 2dx du˜+H(q)dx2 ] , (2.4.35) with H(q) = p3p α−3 2+α 2 ((2 + α)q) 6−2α 2+α . (2.4.36) In the notation of [154], we have the metric form, ds2 = 3 |Λ|q2 ( dq2 + dy˜2 + 2dx du˜± q2kdx2) (2.4.37) with k = 3−α 2+α and the sign corresponding to the sign of p3. The condition for positive energy given in [154] is equivalent to p3(3− α)α > 0 , (2.4.38) which is consistent with the remarks below (2.4.33). The case k = −1, corresponding to Defrise’s pure radiation solution [44] with enhanced symmetry, is the only value Deformations of Spacetime Symmetries 49 of k which does not correspond to any choice of α. Metrics of the form (2.4.37) are conformal to pp-waves and a more detailed physical interpretation is given in [149]. Similar to the previous case of the Petrov solution, using a deformation of the Killing algebra of a given spacetime we have constructed a solution with similar physical properties, namely in this case a pp-wave on an AdS background which now describes gravitational radiation as well as a null Maxwell field. 2.5 Summary We have seen how the mathematical theory of Lie algebra deformations can relate dif- ferent physical situations, in this case exact solutions of general relativity. The Petrov and Kaigorodov-Ozsva´th solutions, the unique vacuum solutions without and with cosmological constant with simply-transitive four-dimensional isometry groups, were shown to be related by Lie algebra deformations; they were shown to be both geodesi- cally complete. While the Petrov solution displays pathological causal behaviour as first observed by Tipler [157] (for the more general class of Lanczos-van Stockum solutions), we were not able to determine the causal properties of the Kaigorodov- Ozsva´th solution. There is however strong indication that the latter shows causal pathologies since we found an unphysical energy-momentum tensor on the boundary in the context of AdS/CFT. We have to leave a more concrete study of the relation of these two properties to further work. Another interesting question that has been left open is whether the Kaigorodov-Ozsva´th solution can be understood as a certain limit of an exterior solution of a matter distribution, perhaps representing a rotating cylinder as well.∗∗ More generally its physical interpretation which is so far largely unknown deserves further study. In a second calculation, we analysed deformations of the Killing algebra of the Kaigorodov spacetime and saw that one is naturally led to a solution including an electromagnetic field. Although we have not constructed new exact solutions here, we have shown how this more algebraic approach to exact solutions of general relativity can give new insights on physical properties. ∗∗A class of rotating cylinder solutions with negative cosmological constant has been studied in [27], who discuss the appearance of CTCs as well, but leave out certain limiting null cases. Our study strongly suggests that the Kaigorodov-Ozsva´th solution, if related to the results of [27], would be a limiting case not considered in [27]. Deformed General Relativity and Torsion 50 Chapter 3 Deformed General Relativity and Torsion “Seit die Mathematiker u¨ber die Relativita¨tstheorie hergefallen sind, verstehe ich sie selbst nicht mehr.” [54] Albert Einstein 3.1 Introduction In the previous chapter, we have discussed exact solutions of general relativity which are homogeneous spaces G/H. Since these are special cases with more symmetry than one would expect in a generic physical situation, it is natural to investigate a more general geometric description in which the symmetry of a given homogeneous space is only present infinitesimally. The passage from special to general relativity can be understood as making the symmetries of a homogeneous space local: General relativity is formulated in terms of (pseudo-)Riemannian geometry, which makes the symmetries of Euclidean space, which can be thought of as a homogeneous space E(p, q)/SO(p, q) ∗, local, and can thus be formulated on arbitrary manifolds. Physically, this is the incorporation of Einstein’s equivalence principle: At at a given point, any spacetime looks to an inertial observer like Minkowski space. In the vier-/vielbein formalism of general relativity (originally due to Cartan as well as the Cartan connections we will use in this chapter), one chooses a basis of the tangent space at each point, and thus a “moving frame” of ∗For p spacelike and q timelike dimensions. We will set p = d− 1 and q = 1 in the following. Deformed General Relativity and Torsion 51 vector fields, such that the metric in this frame is constant and diagonal. The frame field then encodes the metric information of the manifold. In this chapter, we shall investigate a generalised formulation of general relativity in terms of Cartan geometry which incorporates the symmetries of an arbitrary given homogeneous space G/H on an infinitesimal level, thus providing an extension of both Felix Klein’s Erlangen Programme [96] and (pseudo-)Riemannian geometry, as summarised in the following picture that we essentially take from [148]: Euclidean Geometry - curvature 6= 0 (pseudo-)Riemannian Geometry ? ? generalise transformation group generalise tangent geometry Klein Geometry - curvature 6= 0 Cartan Geometry Figure 3.1: Cartan geometry extends both Klein geometry and (pseudo-)Riemannian ge- ometry. In general relativity, if an orthonormal frame has been found, one can still use general local Lorentz transformations to obtain a different frame with the same prop- erties, and so in this formulation general relativity appears as a gauge theory. The description of general relativity is in terms of an h-valued Ehresmann connection (with H = SO(d − 1, 1), the Lorentz group), together with a soldering form (the frame field) determining how the tangent spaces at different points are “soldered” to the spacetime manifold. The gauge group H acts on the tangent space which is viewed as a vector space Rd−1,1. We will use a similar gauge-theoretic formulation, but of a more general character: Instead of viewing Minkowski space as Rd−1,1, we will view it as a homogeneous space E(d − 1, 1)/SO(d − 1, 1). A description in terms of Cartan geometry will allow us to deform general relativity by replacing E(d− 1, 1) by its deformation, the de Sitter group † SO(d, 1). We effectively try to construct a theory of gravity built on tangent de Sitter spaces which replace the usual (co-)tangent spaces in general relativity. As †The anti-de Sitter group SO(d− 1, 2) could be used equally well, and in fact is used in [153]. Deformed General Relativity and Torsion 52 we saw in the previous chapter, from the viewpoint of Lie group deformations this process is a very natural extension of the standard framework. In addition to the speed of light c, a second scale κ with dimensions of momentum is introduced on a fundamental level. It is tempting to speculate that this scale could be related to quantum gravity, although the basic idea of “deforming” the kinematic algebra has existed at least since Heisenberg, who thought in terms of atomic rather than quantum-gravitational scales. The motivation for investigating this possible generalisation of the usual for- mulation of Einstein-Cartan theory, or general relativity in first order formulation, comes from the recently proposed framework of deformed, or doubly, special rela- tivity (DSR), where one effectively proposes a momentum space that is acted on by the de Sitter, not the Poincare´ group, rather in the spirit of Snyder’s proposal [151]. It seems natural to try to construct a general geometric framework which in- corporates this symmetry on the infinitesimal level. We will use a formulation of Einstein-Cartan theory given by Stelle and West, and interpret it in the concrete ex- ample of Minkowski space. By doing this construction, we hope to shed light on the physical interpretation of DSR, which seems rather unclear at present. In particular, while the noncommuting translations appearing in the de Sitter group are commonly interpreted as noncommuting “spacetime coordinates”, and one passes to a noncom- mutative geometry description to account for this, our discussion will stay within conventional differential geometry. We will see that the description of Minkowski space in this theory includes a connection with torsion instead of a fundamentally noncommutative structure. In order to confront this mathematical framework with physical reality, we note that such a connection with torsion may lead to the usual ambiguities in minimal coupling. We give estimates concerning observable effects and note that observable violations of charge conservation induced by torsion should happen on a time scale of 103 s, which seems to rule out these modifications as a serious theory. Our con- siderations show, however, that the noncommutativity of translations in the Snyder algebra need not correspond to noncommutative spacetime in the usual sense. To set the scene, we start with the Einstein-Hilbert action in four dimensions, without cosmological constant, written in first order (Palatini) form: SEH = 1 16piG ∫ abcd ( ea ∧ eb ∧Rcd) , (3.1.1) where Rcd is the curvature two-form of an so(3, 1)-valued connection one-form ωab. Deformed General Relativity and Torsion 53 One discovers that variation with respect to the vierbein field ea gives the usual vacuum field equation Rica = 0 (where Rica is the Ricci one-form), while variation with respect to ωab gives the condition of vanishing torsion. Thus, this formulation has the advantage over the second order form that the condition of a torsion-free connection does not have to be put in by hand. On the other hand, if one now adds a matter action to the Einstein-Hilbert action, both equations now potentially have a non-zero right-hand side. While the functional derivative of the matter action with respect to the vierbein determines an energy-momentum tensor, one will also have nonvanishing torsion if δSmat δωab 6= 0. (3.1.2) This is the case for matter with internal spin, such as spinor fields. This generali- sation of general relativity which incorporates the possibility of torsion is known as Einstein-Cartan theory. Since torsion does not propagate, i.e. the field equation contains the full torsion two-form instead of contractions of it (as for curvature in Ein- stein’s equations for d > 3), it seems difficult to rule out such a theory experimentally, since one would only measure torsion inside a matter distribution with spin. The basic observation is now the following: Combine the connection ωab and the vierbein ea into a single connection, taking values in the algebra of the Poincare´ group, A =   ωab 1 L ei 0 0   , (3.1.3) where L is a length introduced to get dimensions right, but which is completely arbitrary at this stage. Then consider the Yang-Mills-like action (with now d = 4) S = − L 2 32piG ∫ abcd ( F ab ∧ F cd) = − L2 32piG ∫ d4x 1 4 abcd µνρτF abµνF cd ρτ , (3.1.4) involving the curvature F ab of the connection Aab (Latin indices here run from 1 to 4, and so one projects F ab to its so(3, 1) part in this action). Writing out F ab in terms of the curvature Rab of ωab and the vierbein field e i, one discovers that the action (3.1.4) contains a topological Gauss-Bonnet term plus the Einstein-Hilbert action (3.1.1). Thus, the field equations of Einstein-Cartan theory are recovered. The idea that general relativity may be interpreted as a gauge theory with the Poincare´ group as gauge group is of course far from new. In the usual textbook Deformed General Relativity and Torsion 54 discussions, the significance of a gauge theory is expressed as the promotion of a global symmetry in a given theory to a local one by the introduction of an additional field which appears in gauge-covariant derivatives. This viewpoint was probably first elaborated in detail by Utiyama [159], who showed that the promotion of a global to a local symmetry directly leads to a gauge field, valued in the Lie algebra of the symmetry group under consideration, with the usual transformation properties, and thus geometrically to the well known formulation in terms of connections on principal bundles (see e.g. [120]). Utiyama also considered the Lorentz group as an example and was led to general relativity, but only by making several assumptions, such as the existence of a Riemannian metric and a local Lorentz (vierbein) frame and the symmetricity of the affine connection in its lower indices, without further motivation. It was then realised by Kibble [94] that the more natural gauge group to take is the Poincare´ group, so that the vierbein appears naturally as part of the connection. Kibble showed that one is led to general relativity in the first order formalism, i.e. to Einstein-Cartan theory, and so to a connection with torsion when matter with internal spin, such as fermionic matter, is present. The appearance of torsion when fermions are coupled to gravity had earlier been noted by Weyl [165].‡ The result of MacDowell and Mansouri [110] is essentially that by considering a connection taking values in the algebra so(4, 1) of the de Sitter group, one obtains a cosmological term in addition to the Einstein-Hilbert action (3.1.1). (MacDowell and Mansouri also extended the calculations to supergravity, which we will not discuss further.) We will see that the result is not just a mathematical trick, but has a geometric interpretation in terms of a Cartan connection. The rest of this chapter is organised as follows: We discuss the idea of a curved mo- mentum space in section 3.2 and give a brief introduction into the ideas of deformed (doubly) special relativity (DSR) most relevant to the following discussion in sec- tion 3.3. In section 3.4 we outline how Einstein-Cartan theory can be formulated as a gauge theory of gravity with the de Sitter group SO(d, 1) as gauge group; this theory includes a gauge field that plays a crucial role in what follows. In this section we essentially re-derive the results of Stelle and West, using a different set of coordi- nates which we find more closely related to the DSR literature. This section uses the ‡Rather amusingly, Weyl wondered in [165] whether the “gauge transformations” of electromag- netism should not be called “phase transformations” instead – the term “gauge transformation” had of course been introduced by Weyl himself [163] in the context of local conformal transformations (changes in “gauge”). Deformed General Relativity and Torsion 55 language of gauge theory familiar to physicists; we then give a more mathematical account of the geometry of Cartan connections in section 3.5. We then return to a concrete example: To justify our claim that this geometric framework can be used to generalise the ideas of DSR, we show in section 3.6 how, if spacetime is taken to be Minkowski space, the simplest non-trivial choice of zero section leads to a connection with torsion, providing a geometric interpretation for the noncommuting “coordinates” appearing in the Snyder algebra. We close with a discussion of our results and their possible physical implications, which show that the theory, at least in its given form, is not physically viable. We conclude that there may be different physical interpretations of algebraic commutation relations such as those used in DSR. Since the two most obvious extensions of general relativity are admitting either connections with torsion or non-metric connections, we briefly discuss the theory of a torsion-free non-metric connection, known as symmetric affine theory, in a final section. We will see that it does not fit as well into a description by Cartan geometry as the case highlighted in this chapter. 3.2 Curved Momentum Space It is commonly assumed that quantum gravity sets a fundamental length scale, the Planck scale [137], which can not be resolved by any physical experiment. Different approaches to quantum gravity, such as string theory or loop quantum gravity, in- corporate such a scale. This leads to the idea that some kind of “space discreteness” should be apparent even in a low-energy “effective” theory. The idea of putting quantum mechanics on a discrete lattice§ seems to have been first considered by Heisenberg in the spring of 1930 [130], in an attempt to remove the divergence in the electron self-energy. Because the absence of continuous spacetime symmetries leads to violations of energy and momentum conservation, this approach was not pursued further, but later in the same year Heisenberg considered modifying the commutation relations involving position operators instead [130]. A fundamental length scale is absent in special relativity, where two observers will in general not agree on lengths or energies they measure. Hence the usual ideas of Poincare´ invariance must be modified in some way. Snyder observed [151] that §with spacing equal to the Compton wavelength of the proton, lc ≈ 1.3 fm Deformed General Relativity and Torsion 56 this could be done by deforming the Poincare´ algebra into the de Sitter algebra, i.e. considering the isometry group of a (momentum) space of constant curvature. As we have seen in chapter 2, if one maintains the structure of a Lie algebra and considers deformations of the Poincare´ algebra, replacing it by the de Sitter (or anti-de Sitter) algebra is the unique way of implementing a modified kinematic framework [12]. A d-dimensional de Sitter momentum space with curvature radius κ is defined as the submanifold of a (d+ 1)-dimensional flat space with metric signature (d, 1) by (P 1)2 + (P 2)2 + . . .+ (P d−1)2 − (P d)2 + (P d+1)2 = κ2 , (3.2.1) where κ has dimensions of mass. Its isometry group is generated by the algebra [Mab,Mcd] = ηacMbd + ηbdMac − ηbcMad − ηadMbc , [Xa,Mbc] = ηacXb − ηabXc , [Xa, Xb] = 1 κ2 Mab . (3.2.2) Here Mab correspond to a Lorentz subalgebra of the de Sitter algebra, while Xa ≡ 1 κ Md+1,a are interpreted as (noncommuting) translations. These translations are then interpreted as corresponding to coordinates on spacetime; Snyder thought of operators acting on a Hilbert space. Since the operators X1, . . . , Xd−1 correspond to compact rotations in the (d + 1)-dimensional space, their spectrum is discrete. In this way, one obtains “quantised spacetime”, while maintaining Lorentz covariance. It should be stressed at this point that we are interested in the full isometry group acting on momentum space, i.e. the Poincare´ group in the case of a Minkowski space. Clearly, an equation such as (3.2.1) defines a scale κ already in the (d+1)-dimensional Minkowski space; this scale would however not be invariant under translations of this space. In other words, there is no natural length scale in Minkowski space, whereas there is one in de Sitter space, namely the scale κ defining the manifold. There has been some confusion in the literature (see e.g. [147]) about the way in which Lorentz invariance is deformed in the Snyder algebra and the more recent ideas of DSR that we will review shortly. It should be clear from the commutation relations (3.2.2) that the Lorentz subgroup generated by Mab is not deformed in any way. One can give explicit expressions for the algebra elements by choosing coordinates on de Sitter space (3.2.1). The choice made by Snyder is taking Beltrami coordinates p1 = κ P 1 P d+1 , p2 = κ P 2 P d+1 , . . . , pd = κ P d P d+1 , (3.2.3) whence one has (P d+1)2 = κ4/(κ2 + ηabp apb) to satisfy (3.2.1), and ηabp apb ≥ −κ2, corresponding to an apparent maximal mass if pa were interpreted as Cartesian co- Deformed General Relativity and Torsion 57 ordinates on a Minkowski momentum space. (Up to this point one could in principle have chosen anti-de Sitter instead of de Sitter space. Then this inequality becomes ηabp apb ≤ κ2, which perhaps seems less motivated physically.) A necessary sign choice means that these coordinates only cover half of de Sitter space. In these coordinates, the translation generators Xa = 1 κ ( P d+1 ∂ ∂P a − Pa ∂ ∂P d+1 ) = ∂ ∂pa + 1 κ2 pap b ∂ ∂pb (3.2.4) generate “displacements” in de Sitter space. (In this notation, indices are raised and lowered with ηab, the d-dimensional Minkowski metric, so that pa = ηabp b.) The motivation behind these ideas was to cure the infinities of quantum field the- ory, which evidently arise from allowing arbitrary high momenta (or short distances). In a somewhat similar spirit, Gol’fand suggested [73] to define quantum field theory on a momentum space of constant curvature, using Beltrami coordinates as momen- tum variables. This makes the volume of the corresponding Riemannian space finite and so presumably leads to convergent loop integrals in the Euclideanised theory. The consequences for standard quantum field theory were further explored in [92, 74]. Gol’fand only assumed that κ m for all elementary particles; thinking of quan- tum gravity, one would perhaps identify κ with the Planck scale, whereas the original authors seem to have thought of the Fermi (i.e. weak interaction) scale. The induced metric on de Sitter space in terms of the coordinates pa is gnr = κ2 κ2 + p · p ( ηnr − pnpr κ2 + p · p ) , (3.2.5) where p · p ≡ ηcdpcpd. The metric (3.2.5) becomes singular when p · p → −κ2, and negative definite when extended to what Gol’fand calls the exterior region p ·p < −κ2. In four dimensions, det g = −κ10(κ2 + p · p)−5 , (3.2.6) and the volume element is d4p κ5(κ2 + p · p)−5/2. In Gol’fand’s approach (assuming d = 4 of course), the standard Feynman rules were modified by replacing the addition of momenta p and k at a vertex by (p(+)k)a = κ κ2 − p · k ( pa √ κ2 + k · k + ka ( κ− p · k κ+ √ κ2 + k · k )) , (3.2.7) which corresponds to a translation by k of the vector p. (Again p · k ≡ ηabpakb, etc.) It was also noted that spinors now transform under “displacements” as well, which is Deformed General Relativity and Torsion 58 made more explicit in [92] and [74]. As is well known, five-dimensional Dirac spinors still have four components and the matrix γ5 appears in the Dirac Lagrangian, hence there is no chirality. This alone seems to imply that the original Gol’fand proposal cannot be used for an appropriate model of the known particles. Gol’fand’s approach is very different from more recent approaches to quantum field theory on noncommutative spaces (see e.g. [48]) in that the field theory is defined on a momentum space which is curved, but neither position nor momentum space are noncommutative in the usual sense. 3.3 Deformed Special Relativity The idea that the classical picture of Minkowski spacetime should be modified at small length scales or high energies was re-investigated in more recent times, moti- vated by the apparent existence of particles in ultra high energy cosmic rays whose energies could not be explained within special relativity [5]. The proposed framework of deformed special relativity (DSR) [4] modifies the Poincare´ algebra, introducing an energy scale κ into the theory, in addition to the speed of light c. This leads to a quantum (κ-)deformation of the Poincare´ algebra [113], with the parameter κ associated with the newly introduced scale. It was soon realised [100] that this deformed algebra is the algebra of the isometry group of de Sitter space, and that the symmetries of DSR could hence be obtained by identifying momentum space with de Sitter space, identifying Xa as the generators of translations on this space. The constructions of DSR thus appear to be a resurrection of Snyder’s and Gol’fand’s ideas. We take this observation as the defining property of DSR, and will seek to describe a framework in which momentum space, or rather the (co-)tangent space in general relativity, is replaced by an “internal” de Sitter space. We will see that this can best be done using Cartan geometry. When discussing DSR as a modification of special relativity, we take the view that special relativity is defined as a kinematic framework with preferred inertial systems, related to one another by (proper) Lorentz transformations. That is, one has a flat spacetime on which there exist certain preferred coordinate systems, those in which the metric is diagonal with entries ±1. From this point of view, the choice of coordinates on the internal de Sitter space plays quite an important role if one is looking for a “deformation” of special relativity including an energy scale κ. Such Deformed General Relativity and Torsion 59 a deformation can only arise if the chosen coordinate system reduces to Cartesian coordinates on Minkowski space as κ → ∞. The choice of coordinates is obviously not unique. The generators of the algebra will take different explicit forms when different coordinate systems (on four-dimensional de Sitter space) are chosen. In [100] “natural coordinates” are defined by, in the notation of section 3.2, ¶ g = exp [ pI(MI4 +XI) ] exp [ p4X4 ]O , (3.3.1) where O = (0, 0, 0, 0, κ) is taken to be the origin of de Sitter space in five-dimensional Minkowski space, and MI5 and M45 correspond to translations in space and time. The coordinates one obtains are related to the five-dimensional coordinates by P I = pIe p4 κ , P 4 = κ sinh ( p4 κ ) + ~p2 2κ e p4 κ , P 5 = κ cosh ( p4 κ ) − ~p 2 2κ e p4 κ . (3.3.2) Again, these cover only half of de Sitter space where P 4 + P 5 > 0. The metric in these “flat” coordinates is ds2 = −(dp4)2 + e 2p4κ δIJdpI dpJ . (3.3.3) Slices of constant p4 are flat; to an observer using these coordinates the spacetime appears as expanding exponentially. An illuminating discussion of different coordinate systems and kinematics on de Sitter space is given in [29]. The Magueijo-Smolin model [112] corresponds to the following choice of coordi- nates: p1 = κ P 1 P 5 − P 4 , p 2 = κ P 2 P 5 − P 4 , p 3 = κ P 3 P 5 − P 4 , p 4 = κ P 4 P 5 − P 4 , (3.3.4) The generators of boosts in de Sitter space take the form KI ≡ pI ∂ ∂p4 + p4 ∂ ∂pI + 1 κ pIpJ ∂ ∂pJ , (3.3.5) and translations (not considered by the authors of [112]) would take the form XI = p4 + κ κ ∂ ∂pI + 1 κ2 pIp b ∂ ∂pb , X4 = 1 κ pb ∂ ∂pb + p4 + κ κ ∂ ∂p4 . (3.3.6) ¶Capital Latin indices such as I and J used in this section only run over spatial coordinates (from 1 to 3). Deformed General Relativity and Torsion 60 This choice of coordinates is somewhat peculiar as p4 takes a special role, as is also apparent from the modified dispersion relations presented in [112]. The quantity ||p||2 = ηabp apb (1 + 1 κ p4)2 (3.3.7) is invariant under boosts and rotations in de Sitter space, as would ηabp apb be in Beltrami coordinates. Each DSR model corresponds to a choice of coordinates on de Sitter space, such that all expressions reproduce the expressions for special-relativistic Minkowski co- ordinates as κ → ∞. What Magueijo and Smolin call a “U map” is essentially a coordinate transformation from Beltrami coordinates to a different set of coordi- nates, which becomes the identity as κ→∞. In the remaining sections we shall use Beltrami coordinates. Note that this means we always have p · p ≥ −κ2. 3.4 A de Sitter Gauge Theory of Gravity The most direct implementation of the ideas discussed so far into a framework describ- ing more general spacetimes is replacing the cotangent (or tangent) bundle usually taken as phase space by a general symplectic manifold {P , ω}, which can be locally viewed as a product U ×D of a subset U ⊂M of spacetimeM with de Sitter space D. We want to retain the differentiable structure of a manifold, which we do not assume to be present in a full theory of quantum gravity. We also assume that the structure of momentum space is fixed and in particular does not depend on matter fields, as suggested in [119]. If phase space is described as such a manifold, with a choice of origin in the “tangent” de Sitter space at each point, the appropriate mathematical language is that of fibre bundles. The theory of connections on fibre bundles of this type, called homogeneous bundles in [51], was developed by E´lie Cartan (e.g. in [32]). Adopting this framework means there is now an so(d, 1) connection, instead of an so(d − 1, 1) connection, defining parallel transport on spacetime. It was noted by MacDowell and Mansouri [110] that gravity with a cosmological term in four dimensions could be obtained from a theory of such an so(d, 1) connection by projecting it onto its so(d− 1, 1) part in the action. A more elaborate description in terms of Einstein-Cartan theory was then given by Stelle and West [153]. Their analysis included the gauge field needed to identify the fibres at different spacetime Deformed General Relativity and Torsion 61 points, which will be crucial for the interpretation of the theory. The mathematical side of MacDowell-Mansouri gravity as a theory of a Cartan connection is nicely illustrated in [166]; we follow this article as well as the more computationally based presentation of [153], who use the language of non-linear realisations. An overview over the mathematics of Cartan connections is given in [148]. For clarity we first describe the framework in a language more common to physi- cists; a more mathematical account of Cartan connections on homogeneous bundles is given in the following section 3.5. The usual description of general relativity as a gauge theory of the Lorentz group, known as vier-/vielbein formalism, was reviewed at the beginning of this chapter. Since the tangent bundle is in our description replaced by a homogeneous bundle with a curved “tangent” space, one has to effectively use a “double vielbein” formal- ism, in which spacetime vectors are mapped to vectors in the tangent space to the internal (curved) space by a soldering form (vielbein). The picture we have in mind is that of a de Sitter space rolled along the manifold. One then needs to introduce a new field which specifies the point of tangency, expressed in a given coordinate system on the internal space, at each spacetime point. We denote it by pa(x). This corresponds mathematically to a necessary choice of zero section (see below), and physically to a gauge field. Picking a point of tangency at each spacetime point breaks the gauge group SO(d, 1) down to the Lorentz subgroup SO(d− 1, 1) leaving the point of tangency invariant. In more general terms [167], one has a principal bundle over a base manifold M with gauge group G. Locally, a connection on this bundle can be viewed as a one- form taking values in the Lie algebra g. One then considers the projection of the connection to its g/h part and demands that at any point in the manifold, this be an isomorphism between the tangent space at this point and the algebra g/h. In other words, the g/h part eµ i, written as a matrix, has to have full rank. This of course means that the dimension of g/h must be equal to the dimension of the manifold. The gauge field pa(x) now comes in because the projection of the connection to its g/h part is not canonical; for a given subgroup H ⊂ G viewed as the stabiliser of a particular point in G/H, one could equally well take a conjugate group gHg−1 ' H stabilising a different point in G/H. This will become more explicit shortly when we consider the MacDowell-Mansouri action. The procedure does not depend on the choice of gauge group G, and would be Deformed General Relativity and Torsion 62 equally possible if the gauge group were chosen to be the Poincare´ group E(d− 1, 1); such a formulation for gravity, including copuling to matter, is indeed discussed in [75]. In that case, the homogeneous space G/H would be Minkowski space, viewed as an affine space. As the value of the gauge field pa(x) at a particular spacetime point picks an origin in this affine space, it may then be identified with the vector space R d−1,1: At a given point in Minkowski space, the tangent space to this point can be identified with Minkowski space itself. Although all calculations we will present go through in the case of E(d − 1, 1) as gauge group (resulting in different formulae of course), the formalism may seem a bit redundant since there is no real necessity for the second vielbein field. In the more general case of a homogeneous space G/H which acts as a “tangent space” instead of a flat vector space, it is however necessary. We consider a theory with gauge group SO(d, 1), so that the connection A takes values in the Lie algebra so(d, 1). It can be split as (again introducing a length l on dimensional grounds) A =   ωab 1 l ei −1 l ei 0   , (3.4.1) so that ωab acts as the usual so(d − 1, 1)-valued connection of general relativity and ei as a vielbein one-form. We have simultaneously unified the usual connection and the vielbein, and replaced the (flat) tangent space by a curved “internal” space, such that the de Sitter group and not the Poincare´ group now appears as a gauge group. (Lorentz) indices on ωab and e i are raised and lowered using ηab. A gauge transformation, i.e. a local transformation g(x) taking values in the de Sitter group, can be split as g(x) = s(x)Λ(x), where s(x) changes the zero section, i.e. changes the local identification of points of tangency at each spacetime point, and Λ(x) is a local Lorentz transformation in the vielbein formalism of general relativity which does not mix the ωab and e i parts of the connection. In this notation, the connection transforms under a gauge transformation as A(x)→ A′(x) = Λ−1(x)s−1(x)A(x)s(x)Λ(x)+Λ−1(x)s−1(x)ds(x)Λ(x)+Λ−1(x)dΛ(x) . (3.4.2) One can use this equation to relate the connection A0 corresponding to the triv- ial zero section, where the point of tangency is the origin of the internal space at each spacetime point, pa(x) ≡ (0, 0, 0, 0), to a connection corresponding to any given Deformed General Relativity and Torsion 63 zero section. The physical significance of this is the following. Assume we have fixed pa(x) ≡ (0, 0, 0, 0). Then an action can be defined from the curvature of the connection A (here R is the curvature of the so(d− 1, 1) part of A), F = dA+ A ∧ A =   Rab − 1l2 (ea ∧ eb) 1l T i ≡ 1l (dei + ωij ∧ ej) −1 l Ti 0   . (3.4.3) In four dimensions, the MacDowell-Mansouri action [110, 166] is S = − 3 32piGΛ ∫ abcd ( F ab ∧ F cd) = − 3 32piGΛ ∫ d4x 1 4 abcd µνρτF abµνF cd ρτ , (3.4.4) where the Latin indices run from 1 to 4, and so one projects F to its so(3, 1) part in this action. Apart from a topological Gauss-Bonnet term, the action (3.4.4) is equivalent to the Einstein-Hilbert action with a cosmological term S = 3 16piGΛ 1 l2 ∫ abcd ( ea ∧ eb ∧Rcd − 1 2l2 ea ∧ eb ∧ ec ∧ ed ) , (3.4.5) where we have to identify Λ = 3 l2 (3.4.6) as the cosmological constant.‖ In order to define the projection of F in the action (3.4.4), one has used a splitting so(d, 1) ' so(d, 1)/so(d− 1, 1)⊕ so(d− 1, 1) , (3.4.7) which depends on the gauge field since the subgroup SO(d − 1, 1) leaving a given point in de Sitter space invariant depends on the choice of this point. When the action (3.4.4) is coupled to matter, the so(d, 1)/so(d−1, 1) part ea of the connection appears in a volume element in the matter Lagrangian. By varying the action one obtains the field equations of Einstein-Cartan theory with a cosmological constant Λ = 3/l2. The length scale l, which is so far arbitrary, can be chosen to reproduce the Λ of the observed universe, which means it must be chosen to be very large (the “cosmological constant problem”). By the field equations, one can ‖Note that in the formulation in terms of the Poincare´ group, with connection (3.1.3), no cosmo- logical constant appears and the length scale L remains undetermined. Deformed General Relativity and Torsion 64 determine for a given matter distribution a connection A0 consisting of an so(d−1, 1) connection (ωab)0 and a vielbein e i 0. The MacDowell-Mansouri action reproducing Einstein-Cartan theory with a cos- mological constant includes a gauge choice. We can hence view it as the gauge-fixed version of a more general theory. Since (3.4.2) determines how the connection trans- forms under a gauge transformation, we can generalise a given solution of Einstein- Cartan theory to an arbitrary gauge choice. The extension of the theory to arbitrary configurations of the gauge field, and hence arbitrary choices of tangency points of the internal space to spacetime, is what we call Einstein-Cartan-Stelle-West theory. Any solution of Einstein-Cartan theory, in particular any (torsion-free) solution of general relativity, gives rise to more general solutions of Einstein-Cartan-Stelle-West theory via (3.4.2). We will later see that one can construct an so(d− 1, 1) connection with torsion from a torsion-free one. In (3.4.2), s(x) takes values in the de Sitter group, more precisely in the subgroup generated by “translations” which leaves no point of de Sitter space invariant. The correspondence between Beltrami coordinates pa(x) on de Sitter space and such group elements is given explicitly by s(p(x)) = exp [ pi(x)√−p(x) · p(x) Artanh (√−p(x) · p(x) κ ) κXi ] . (3.4.8) Then the group element s(p(x)) maps (0, 0, 0, 0) to (p1(x), p2(x), p3(x), p4(x)) in Bel- trami coordinates. A different choice of coordinates in the internal de Sitter space would correspond to a different parametrisation of the elements of the subgroup of translations of the de Sitter group. Inserting (3.4.8) into (3.4.2) and setting Λ(x) ≡ e, we obtain ωab(p(x)) = paeb0 lκγ(p) + ( 1− 1 γ(p) ) padpb + ωca0 p bpc p · p + 1 2 ωab0 − (a↔ b) , ei(p(x)) = lκ p · p+ κ2 ( pi pcdp c p · p (1− γ(p)) + dp iγ(p) + (ωib)0p bγ(p) ) +piea0pa 1 + κ 2 p·p(1− γ(p)) p · p+ κ2 + ei0 γ(p) , (3.4.9) where γ(p) ≡ √ p · p+ κ2 κ2 = 1 + p · p 2κ2 + . . . (3.4.10) Because p ·p ≥ −κ2 in Beltrami coordinates, the square root is always real. In the limit p · p→ 0, our parametrisation is the same as that used in [153], and we recover Deformed General Relativity and Torsion 65 their results ωab(p(x)) = ( 1 2 ωab0 + 1 lκ paeb0 + 1 2κ2 ( padpb + ωca0 p bpc ))− (a↔ b) , ei(p(x)) = ei0 + l κ ( − 1 2κ2 pipcdp c + dpi + ωib0 pb ) + 1 2κ2 piea0pa . (3.4.11) Near p = 0, we have ωab(p(x)) = ωab0 +O (p κ ) , ei(p(x)) = ei0 + l κ dpi +O (p κ ) . (3.4.12) As mentioned above, the so(d, 1)/so(d − 1, 1) part of the connection A acts as a vielbein and maps vectors in the tangent space at a point x in spacetime to vectors in the tangent space at p(x) in the internal de Sitter space, given in components with respect to an orthonormal basis at p(x). In order to give their components in the coordinate-induced basis { ∂ ∂pa }, we need another vielbein, which can be obtained from (3.4.9) by setting ω0 = e0 = 0 (corresponding to spacetime being de Sitter space with cosmological constant Λ) and pa(x) = κ l xa, as in [153]. We obtain ln a(p(x)) = κ2 δn a(p · p)γ(p)− papn(γ(p)− 1) (p · p)(p · p+ κ2) , (3.4.13) where n is a coordinate index in the internal space and a denotes a Lorentz index, as before. This vielbein is of course independent of the underlying spacetime. Parallel transport can be defined for the so(d, 1) connection using the notion of development, which generalises the usual covariant derivative. One introduces a de- velopment operator [153] D = d− 1 2 ωabMab − (e · V ) , (3.4.14) where the second term is the usual infinitesimal relative rotation of tangent spaces at different spacetime points, and the last term compensates for the change of point of tangency and hence generates maps from the tangent space at one point of de Sitter space to the tangent space at a different point of de Sitter space. Again one should think of an internal space rolled along spacetime [166]. In components, in our conventions we have (ωabMab) c d = −2ωcd , (3.4.15) and the combination eaVa acts on Lorentz indices as an element of so(d − 1, 1), rep- resenting the map from one tangent space to another in the respective bases. We use Deformed General Relativity and Torsion 66 the result obtained by [153] using the techniques of non-linear realisations, namely that when expressed as an so(d− 1, 1) matrix, l(e · V ) = κ s(p)−1(eaXa)s(p)− s(p)−1 [s(p+ δp)− s(p)] , (3.4.16) where s(p) is defined according to (3.4.8) and δp is determined from the equation [s(p+ δp)]a5 = [ (1 + ebXbκ)s(p) ]a 5 (3.4.17) where only terms linear in ea are kept in δp. An explicit calculation shows that δpa = pa κ (ηbce bpc) + eaκ , (3.4.18) and hence near p = 0, we have δpa = κea, as expected. We find that (e · V ) has components (e · V )bc = κ(ebpc − ecpb)(1− γ(p)) l(p · p) . (3.4.19) One then has a notion of holonomy, mapping closed loops in spacetime into the internal space by development. In particular, if one develops the field p(x) describing the point of tangency around an infinitesimal closed loop at x0, the developed value will in general differ from the original value at x0 [153]: ∆pa(x0) ∝ Tµνi(x0)lai(p(x0)) ∮ xµdxν , (3.4.20) where lai(p(x)) is the inverse of the vielbein (3.4.13) and Tµν i are the components of the torsion tensor T = de + ω ∧ e. This is because by specifying pa(x), one locally identifies the “internal” de Sitter spaces, thinking of a tangent de Sitter space rolled along the manifold. In a generic situation, where the manifold M is not identical to the internal space, such an identification is, strictly speaking, only possible at a point. To understand the geometry, one may picture a sphere being rolled along a plane, as in figure 3.2. In general, when the sphere is rolled around and returns to its initial position on the plane, its point of tangency will not be the original one. This is just the situation for the Lorentzian analogues, Minkowski space and de Sitter space, as we will discuss shortly. The central result we will try to justify in the following is that, starting from Minkowski spacetime, if we assume the internal de Sitter space is rolled along Minkowski space in a non-trivial way, we obtain a connection with torsion. In our interpretation, this is the only way that “coordinates” can act as translations on momentum space, as one normally assumes when associating the Snyder algebra with a noncommutative spacetime. Deformed General Relativity and Torsion 67                     &% '$ Figure 3.2: When the curved internal space is rolled along Minkowski space, a path in spacetime corresponds to a path in the internal space. Because of the curvature of the internal space, a closed path in Minkowski space does not correspond to a closed path in the internal space, which is manifest as torsion. (We have drawn Riemannian spaces, a plane and a sphere, for clarity of presentation.) 3.5 Cartan Connections on Homogeneous Bundles This more mathematical introduction into Cartan connections on homogeneous bun- dles relies mainly on [166], but mentions some additional points which are of impor- tance to our discussion of Einstein-Cartan-Stelle-West theory. The tangent bundle of a manifold needs to be replaced by a fibre bundle whose fibres are homogeneous spaces D ≡ SO(d, 1)/SO(d − 1, 1). This can be achieved by starting with a principal bundle P (M, SO(d, 1)), and considering the associated bundle P = E(M,D, SO(d, 1), P ) = P ×SO(d,1) D (taken as phase space); it can be identified with P/SO(d− 1, 1) by the map ν : P → P/SO(d− 1, 1), [u, a · SO(d− 1, 1)] 7→ ua · SO(d− 1, 1). (3.5.1) Then the structure group SO(d, 1) is reducible to SO(d − 1, 1) if the associated bundle P admits a cross section σ :M→ P [99]; furthermore, there is a one-to-one correspondence between reductions of the structure group and cross sections. This cross section, called a zero section in [135], corresponds to a choice of origin in the momentum space attached to each point. In physicist’s terms, the de Sitter group is spontaneously broken down to the Lorentz group by the choice of points of tangency in the tangent de Sitter spaces at each spacetime point. The bundle reduction depends on the choice of zero section, or rather, its local representation in coordinates as a function M ⊃ U → U × D. This is because the Deformed General Relativity and Torsion 68 embedding of SO(d−1, 1) into SO(d, 1) is not canonical, as the stabilisers of different points in D are isomorphic but related by conjugation. In other words, the mappings appearing in the exact sequence 0→ SO(d− 1, 1)→ SO(d, 1)→ SO(d, 1)/SO(d− 1, 1)→ 0 (3.5.2) are not canonically chosen (cf. the discussion for the affine group in [135]). It is of course possible to choose canonical coordinates such that the function representing the zero section is just x 7→ (x, [e]) ≡ (x, SO(d − 1, 1)) ∈ M × D. However, in general we want to locally identify the fibres at nearby base space points, adopting the viewpoint that there is a single tangent D space which is “rolled along” the manifold. Then we need to retain the general coordinate freedom. (This point is missing in the discussion of [166].) An exact identification is only possible when the connection is flat. Let us assume that coordinates on P have been fixed, and that it is the zero section, and hence the identification of the fibres, that is varied∗∗. After a choice of zero section, there is still a local gauge freedom corresponding to the stabiliser SO(d− 1, 1). We express a given section as s(x), where σ(x) = (x, s(x)) ∈ M×D in our coordinates. The section that corresponds to s0(x) ≡ [e] will be called “trivial”. An so(d, 1)-valued Ehresmann connection A in P is in general not reducible to an so(d−1, 1)-valued connection in the reduced SO(d−1, 1) bundle PR(M, SO(d−1, 1)). It can, however, be pulled back using the inclusion ιx : SO(d− 1, 1)→ SO(d, 1), Λ 7→ s(x)Λs(x)−1 (3.5.3) to a Cartan connection AC on the reduced bundle ††. Of course reducing the connec- tion to an so(d − 1, 1)-valued connection and pulling it back to a Cartan connection are very different operations, since in the latter case one wants the so(d, 1)/so(d−1, 1) part of the pulled-back connection to act as a soldering form, so in particular to be non-singular. We obtain a bundle sequence (cf. [166]) ∗∗one is free to choose an “active” or “passive” viewpoint here ††We assume here that the necessary condition kerA ∩ (ιx)∗(TPR(M, SO(d − 1, 1))) = {0} (see [148]) is satisfied. Deformed General Relativity and Torsion 69 PR(M, SO(d− 1, 1)) - P (M, SO(d, 1)) - P/ιx(SO(d− 1, 1)) ' P ν−1x @ @ @ @ @ @R ? M The reduced bundle PR(M, SO(d− 1, 1)) is mapped into P (M, SO(d, 1)) by p 7→ [p, e] = {(pΛ−1, s(x)Λs(x)−1)|Λ ∈ SO(d− 1, 1)} ∈ PR(M, SO(d− 1, 1))×ιx(SO(d−1,1)) SO(d, 1) . (3.5.4) The connection one-form A on M, induced by the connection A on P , depends on a choice of section τ : M→ P (M, SO(d, 1)). If the zero section σ is fixed, one can choose an arbitrary (local) section τR :M→ PR(M, SO(d−1, 1)) to obtain a section τ ; in local coordinates, σ(x) = (x, s(x)) , τR(x) = (x,Λ(x)) −→ τ(x) = (x, s(x)Λ(x)s(x)−1 · s(x)) = (x, s(x)Λ(x)) . (3.5.5) For practical computations, it is often useful to first consider the trivial section. The induced connection corresponding to this section, denoted by A0(x), is related to the connection for a general section by A(τ(x)) = Λ−1(x)s−1(x)A0(x)s(x)Λ(x) + Λ−1(x)s−1(x)ds(x)Λ(x) + Λ−1(x)dΛ(x) . (3.5.6) Once the zero section s(x) has been fixed, there is still the freedom of SO(d − 1, 1) transformations, corresponding to different choices of Λ(x) in (3.5.6). These are the standard local Lorentz transformations in the vielbein formalism of general relativity. The choice of zero section induces a local splitting of the so(d, 1) connection, according to so(d, 1) ' so(d, 1)/so(d− 1, 1)⊕ so(d− 1, 1) . (3.5.7) This splitting is invariant under the adjoint action of SO(d− 1, 1), thus the different parts of the connection will not mix under SO(d − 1, 1) transformations. We are considering a reductive geometry (cf. (1.1.6)); recall that the existence of such a splitting g = x⊕ h (3.5.8) Deformed General Relativity and Torsion 70 invariant under the adjoint action of H, implies that we may identify g/h with the subspace x of g. In the present case, this is of course the subspace spanned by the generators Md+1,a of SO(d, 1). For the de Sitter group, just as for the Poincare´ and anti de Sitter groups, one can say more: The Lie algebra so(d, 1) is equipped with a Z2 grading; the subspaces in (3.5.8) satisfy [h, h] ⊂ h , [h, x] ⊂ x , [x, x] ⊂ h (3.5.9) and the homogeneous space G/H is a symmetric space [167]. The fact that de Sitter space viewed as a homogeneous space is a symmetric space is directly responsible for the splitting of the curvature F of the Cartan connection A into curvature and torsion of the so(d − 1, 1) connection ω that we have observed above [167]. A general Cartan connection A = ω + e, where ω takes values in h and e takes values in x, has curvature F = R + 1 2 [e, e] + dωe , (3.5.10) where [e, e] denotes taking the wedge product on the form part and the commutator on the Lie algebra part, i.e. [e, e]iµν = Cj i k ( ej ∧ ek) µν = Cj i k ( ejµe k ν − ekµejν ) (3.5.11) if Cj i k are the structure constants as in (1.1.4). The middle term generically takes values in h and x, so that one has a splitting Fh = R + 1 2 [e, e]h , Fx = 1 2 [e, e]x + dωe . (3.5.12) For a symmetric space, however, [e, e]x vanishes [167], and we obtain the result used in the construction of the MacDowell-Mansouri action (3.4.4). Because we assume A to be a Cartan connection, the so(d, 1)/so(d−1, 1) part acts as a soldering form, corresponding to the standard vielbein of general relativity; in particular, eµ i is an invertible matrix. The soldering form maps vectors in the tangent space TxM at a point x in spacetime to vectors in the tangent space Tp(x)D at p(x) in the internal de Sitter space, given in components with respect to an orthonormal basis at p(x). The vielbein that maps between the components of a vector in the orthonormal basis and the coordinate-induced basis is given in (3.4.13). Deformed General Relativity and Torsion 71 3.6 Synthesis The notion of development along curves in spacetime is central to the interpretation of Einstein-Cartan-Stelle-West theory, because it allows “spacetime coordinates” to act as translations in the internal de Sitter space. The situation described by DSR, where noncommuting translations on a curved momentum space are interpreted as noncommuting spacetime coordinates, here corresponds to a Minkowski spacetime with an internal de Sitter space rolled along this Minkowski space. The gauge field pa(x) specifies the points of tangency of the internal space at each spacetime point, and we have chosen Beltrami coordinates on de Sitter space which look like Cartesian coordinates on Minkowski space near the “origin” of de Sitter space. Since the internal space has a natural scale κ and we needed to introduce a natural scale l in spacetime, we choose the gauge field to be pa(x) = κ l xa (3.6.1) in a vicinity of the origin of spacetime which is now taken to be Minkowski space, where xa are the standard Minkowski coordinates such that the connection vanishes in general relativity. In general a closed path in spacetime will not correspond to a closed path traced out on the internal space, hence such an identification is only local and, strictly speaking, only valid the origin of Minkowski spacetime. On dimensional grounds, the effects of torsion scale as x l or p κ . For (3.6.1) to be well-defined, we must guarantee that x · x ≥ −l2, so l should be large in Planck units. We will comment on the significance of the scale l at the end of this section. It should perhaps be emphasised that the gauge field pa(x) does not represent physical momentum, but determines the point of tangency of the internal space we have introduced which is to some extent arbitrary. Tangent vectors to the original spacetime can be mapped to tangent vectors to the internal space via the vielbein. The physical interpretation of motion in an internal “momentum” space which is related to motion in spacetime seems obscure, but if coordinates are to act as translations in the internal space, the two must be connected in some way. In this sense, we are constructing the minimal non-trivial gauge field which leads to observable effects, and an alternative interpretation of noncommuting generators Xa in the Snyder algebra. In our interpretation, different points in the internal de Sitter space do not repre- sent different values for physical four-momentum. Hence we avoid problems with the physical interpretation of DSR, such as the “spectator problem” of noncommutative momentum addition and the “soccer ball problem” of how to describe extended ob- Deformed General Relativity and Torsion 72 jects, given that any momentum would appear bounded by the relation p·p ≥ −κ2. In our framework, tangent vectors representing a particle’s (or extended body’s) velocity remain vectors and as such live in an unbounded space with commutative addition. As explained before, we can use equations (3.4.9) to obtain the connection com- ponents ω and e corresponding to this choice of our gauge field; we set ω0 = 0 and (eµ a)0 = δµ a and substitute (3.6.1) to get ωµ ab = ( xaδµ b − xbδµa ) x · x+ l2(γ(x)− 1) l2(x · x)γ(x) , (3.6.2) eµ i = 1 (x · x)(x · x+ l2) ( xixµ(x · x− 2l2(γ(x)− 1)) + δµi(x · x)2l2γ(x) ) and ∂νeµ i − ∂µeνi = ( xνδµ i − xµδνi ) l2(2l2(γ(x)− 1)− 3(x · x))− (x · x)2 (x · x)(x · x+ l2)2 , ων ibeµb − ωµibeνb = (2l 2 + x · x) (x · x+ l2(γ(x)− 1)) (x · x)(x · x+ l2)l2γ(x) ( xνδµ i − xµδνi ) , (3.6.3) where now γ(x) ≡ √ x · x+ l2 l2 , (3.6.4) which gives a non-zero torsion Tµν i = ( xνδµ i − xµδνi ) 1 l2 √ x·x l2 + 1 . (3.6.5) Interestingly enough, for the choice of zero section (3.6.1) the scale κ drops out of all expressions. Expressed in coordinates on the internal space, one has Tµν i = ( pνδµ i − pµδνi ) 1 lκ √ p·p κ2 + 1 . (3.6.6) The quantity Tµν i will be multiplied by an infinitesimal closed loop ∮ xµdxν to give the difference in the value p(x) caused by development along this loop. In momentum coordinates, this is equal to l κ ∮ pµdpν , and the effect of going around the developed curve in the internal space is (near x = 0 or p = 0) proportional to κ−2, just as was suggested by (3.2.2). Expressing Minkowski space in the usual coordinates, together with the (local) identification pa(x) = κ l xa, in this framework gives a connection with torsion. De- veloping a closed curve in spacetime in the internal space will give a curve that does not close in general, which is the effect of noncommuting translations in the internal space. Deformed General Relativity and Torsion 73 The reader may wonder how the “deformation” of the Minkowski solution de- scribed here is manifest in a metric. We can define a metric by the usual expression gµν = e a µe b νηab . (3.6.7) This metric would not determine the connection, but could be used to define distances in the spacetime in the usual way. Then, from (3.6.2), we get gµν = ηµν 4 1 + x·x l2 + xµxν (x · x) ((x · x) + l2)2 . (3.6.8) It should be stressed that the connection on spacetime is not the Levi-Civita con- nection of this metric, as we are working in a first order formulation where metric and connection are independent. There is a factor of 4 because of a term in (3.4.12) which does not necessarily go to zero as p → 0. With the identification (3.6.1), the soldering form always gets a contribution eµ i(x) = (eµ i)0 + δµ i +O (x l ) . (3.6.9) The limit κ → ∞ is now identified with the limit l → ∞, in which we recover the (rescaled) Minkowski metric. In deriving the expressions (3.6.2) we started with Minkowski space, which clearly solves the field equations of the Einstein-Cartan theory for an energy-momentum tensor cancelling the cosmological constant term, and vanishing internal spin. In changing the zero section, we then performed a SO(d, 1) gauge transformation, under which the curvature F transformed as F (s(x)) = s−1(x)F (x)s(x). (3.6.10) This is a general SO(d, 1) rotation which mixes up the so(d−1, 1) and so(d, 1)/so(d− 1, 1) parts of the connection and the curvature. Hence, the resulting connection will no longer solve the original field equations, but the field equations for an energy- momentum tensor which has also undergone a SO(d, 1) transformation. This mixes up the energy-momentum and internal spin parts, combining them into an element of the Lie algebra so(d, 1), the interpretation of which seems obscure at least. A comment is in order with regard to physical units. In addition to the energy scale κ, which is perhaps naturally identified with the Planck scale, the identification of lengths with momenta, necessary in the framework presented here, requires the choice of a unit of length l which is not necessarily connected to the scale κ. It may Deformed General Relativity and Torsion 74 well be that it is instead the cosmological constant which sets this length scale, leading to an astronomical scale instead of a sub-atomic one. And indeed, some more recent approaches to quantum gravity (e.g. [58]) use the product GΛ as a dimensionless parameter in a perturbative expansion. A fixed positive Λ also seems to be required in non-perturbative approaches to quantum gravity [6]. Then the cosmological constant may play the role of a fundamental parameter in quantum gravity. 3.7 Summary and Discussion It has been argued that the algebra of DSR describes the symmetries of a semiclas- sical limit of (a generic theory of) quantum gravity (see e.g. [6]). If this claim is taken seriously, one has to give an interpretation of the noncommuting translations appearing in the algebra, and usually they are supposed to represent a spacetime with a fundamentally noncommutative structure [111]. Alternatively, one may view the apparent noncommutativity as an artefact of the finite resolution of lengths [70]. However, there are fundamental difficulties in associating these noncommuting op- erators directly with coordinates on spacetime, as position is not additive in a way that momentum and angular momentum are [37]. Furthermore, as also pointed out in [37], a proposed noncommutativity of spacetime of the form (3.2.2), proportional to angular momentum or boost generators, and hence vanishing at a given “origin”, seems deeply at odds with any idea of (even Galilean) relativity. This would also be an obvious criticism of the framework presented in this chapter, when taken as a theory that is supposed to describe the real world. What we have shown here, is that using the framework of Einstein-Cartan-Stelle- West theory, one reaches a different conclusion from the usual one: The noncommu- tativity of translations on a momentum space of constant curvature is interpreted as torsion of a connection that solves the equations of Einstein-Cartan theory with a modified energy-momentum tensor that mixes with the spin tensor. If one takes this seriously, one is led to conclude that there is an effect of torsion induced by quantum gravity, whose effects would however only become measurable over distances compa- rable to l, a length scale presumably associated with the cosmological constant. No such effect appears in de Sitter space with an appropriate cosmological con- stant, or indeed any vacuum solution of the theory. Vacuum solutions are then just described by the Poincare´ algebra, and hence undeformed special relativity. It is curi- Deformed General Relativity and Torsion 75 ous that the curvature in the internal “momentum” space is reflected in torsion when Minkowski spacetime is considered as spacetime, which would however seem rather unnatural as it is not a vacuum solution any more. This result does however seem to be the most natural interpretation of the DSR algebra in a differential-geometric framework capable of describing more general spacetimes. Any non-zero energy-momentum tensor, however, will lead to a connection having torsion. In theories such as Einstein-Cartan theory, this leads to well-known problems when trying to couple the gravitational field to Maxwell fields, for instance, as there is no unambiguous procedure of minimal coupling. This is because the statement that the exterior derivative is independent of the choice of connection, (dA)µν ∝ ∂[µAν] = ∇[µAν] , (3.7.1) is true precisely when torsion vanishes. Using an so(d − 1, 1) connection, this is apparent from d(eiAi) = ∇(eiAi) = Ai∇ei − ei ∧∇Ai = −ei ∧∇Ai + AiT i (3.7.2) where ∇ei = dei+ωij ∧ ej etc. One has two different candidates for the field strength F , namely ei ∧ ∇Ai and d(eiAi), with possibly observable differences between these choices, although it could be argued that F = dA is the only meaningful choice because it preserves gauge invariance [21]. In the framework of Einstein-Cartan-Stelle-West theory, gauge fields should be coupled to gravity via development, i.e. replacing F = dA by F = DA. We compute from (3.4.9) and (3.4.19) that development can be expressed in terms of ω0 and e0 by D = d+ ω − (e · V ) = d+ ω0 + 2(p⊗A e0)κ(γ(p)− 1) l(p · p) =: d+ ωeff , (3.7.3) where ⊗A is an antisymmetrised tensor product, (U ⊗A V )ab ≡ U [aV b] ≡ 12(UaV b − U bV a). Parallel transport is effectively described by the connection ωeff , whose torsion is in general non-zero. One can give an explicit formula for the torsion which is however rather complicated and does not seem to give much insight; to linear order in pi, one has T i = l κ (Rib)0p b − 1 2lκ ei0 ∧ (e0jpj) + 1 2κ2 ( piej0 ∧ dpj − pjei0 ∧ dpj ) +O(p2). (3.7.4) If we assume a universal relation of internal momenta and spacetime lengths of the form p ∼ κ l x, the second and third terms seem to give contributions of order x/l2. The Deformed General Relativity and Torsion 76 first term is proportional to the local curvature of ω0, R i b = dω i b+ω i j∧ωjb, contracted with xb. Note that it is the Riemann tensor, not the Ricci tensor, that appears, so that propagating degrees of freedom of the gravitational field are included. This first term should in realistic situations, even in vacuum, give the dominant contribution. Assuming that minimal coupling is achieved through the development operator D, or equivalently by using the effective connection which has torsion, one would couple vector or matter fields (using Dψ for spinors) to torsion, breaking gauge invariance. Such an effect of course leads to the absence of charge conservation, and this should be experimentally observable in the presence of a non-trivial gravitational field, i.e. in regions where spacetime is not exactly de Sitter. Let us recall that in standard tensor calculus one uses the identity [∇µ,∇ν ]Mλρ = RµνλσMσρ +RµνρσMσλ − Tµσν∇σMλρ (3.7.5) which gives for an antisymmetric Mλρ when contracted [∇λ,∇ρ]Mλρ = −gµλgνρTµσν∇σMλρ, (3.7.6) to establish that the right-hand side of Maxwell’s equation ∇λFλρ = 4piJρ satisfies a continuity equation in the absence of torsion. With torsion present, one has then for any region R ∫ ∂R d3x √ h nλJλ = 1 4pi ∫ R d4x √ g (−gµλgνρTµσν∇σFλρ) . (3.7.7) Effects become important when the size of the region R is comparable to the length scale of torsion. As an example consider the Schwarzschild solution, which has Kretschmann scalar RabcdR abcd ∼ r 2 S r6 , (3.7.8) so roughly Rabcd ∼ rSr−3. Assuming that the origin of the x coordinate system corresponds to the centre of the Earth, we would, on the surface of the Earth, measure a torsion of order rSR −2, where R is the radius of the Earth. With rS ∼ 10−2 m and R2 ∼ 1013m2, the length scale for effects of torsion would be about 1011 m. The other two contributions, given that l ∼ 1026 m, would be much smaller. Although this crude estimate suggests that effects will be very small, even tiny violations of charge conservation should have been observed experimentally. For a discussion of experimental tests of charge conservation and possible extensions of Maxwell theory Deformed General Relativity and Torsion 77 in Minkowski space, see [102]. Processes such as electron decay on a length scale of 1011 m, or a time scale of 103 s, can clearly be ruled out. The example presented here shows that the correct physical interpretation of purely algebraic relations, such as the commutators of the Snyder algebra, may not be the seemingly obvious one. We conclude that the physical motivation for assuming spacetime is “noncommutative” may not be as clear as often assumed. 3.8 Gauge Invariance Broken? When discussing the issue of possible consequences of broken gauge invariance, we must bear several points in mind. The idea that an asymmetry between the proton and electron charges could have interesting astrophysical consequences goes back to Lyttleton and Bondi [108], who argued that a charge difference, and hence a net charge of the hydrogen atom, of 10−18 elementary charges, might explain the observed expansion of the universe by electrostatic repulsion. This idea was proposed in connection with Hoyle’s ideas of a universe in a steady state, which required continuous production of material via a “creation field” [85], and a modification of Maxwell’s equations was proposed to accommodate charge nonconservation. From Hoyle’s perspective, however, the steady state model was incompatible with expansion of the universe by electrostatic repulsion, and should lead to electrostatic attraction instead [86]. There seems to be no need for the electron and proton charges to be of equal magnitude to maintain gauge invariance. However, if the universe as a whole is not neutral, but it is homogeneous, gauge invariance must be broken. Hence the two issues are closely related. Modern laboratory experiments [115] give a bound of 10−21 elementary charges on the difference of electron and proton charge; astrophysical considerations give bounds of 10−26 elementary charges using the isotropy of the cosmic microwave background [30], or 10−29 elementary charges by considering cosmic rays [124]. Recently, an interesting proposal to measure net charges of atoms and neutrons, sensitive to 10−28 elementary charges, was put forward [8]. From a theoretical viewpoint, if gauge invariance is broken, it is natural to assume a nonvanishing photon mass. One then considers Einstein-Proca theory, an outline of which can be found in [79]. The photon may also be charged. Here, experimental bounds on the charge are 10−29 elementary charges using pulsars [141], and possibly Deformed General Relativity and Torsion 78 10−35 elementary charges from CMB isotropy [30]. Experimental bounds on violations of gauge invariance in electrodynamics are very tight, and hence any theory predicting torsion which is coupled to electromagnetism faces severe problems when confronted by experiment. In the framework of Einstein- Cartan-Stelle-West theory, it is possible to maintain gauge invariance by choosing F = dA, but using the development operator is the most natural choice. 3.9 Symmetric Affine Theory If Einstein-Cartan theory is considered as the extension of general relativity which allows for torsion, there is an analogous extension which allows for a non-metric con- nection. This theory can be formulated in terms of a torsion-free gl(n,R) connection and is known as symmetric affine theory. It is equivalent to standard general relativity with a massive vector field, known as (nonlinear) Einstein-Proca theory [79]. One could attempt to embed this theory into a theory of a connection taking values in the algebra of the affine group a(n,R)‡‡, A =   ωab 1 l ei 0 0   , (3.9.1) where now ωab is not constrained by ω ab = −ωba. Geometrically, this means that the connection does not preserve the lengths of vectors under parallel transport. The corresponding curvature of A (R is the curvature of the gl(n,R) part), F = dA+ A ∧ A =   Rab 1 l T i ≡ 1 l (dei + ωij ∧ ej) 0 0   , (3.9.2) would then be constrained by demanding that T i ≡ 0. This seems rather unnatural from the perspective of Cartan geometry. Furthermore, the length scale l is now completely arbitrary as it does not appear in the gl(n,R) part of the curvature any more, just as for a connection taking values in the Poincare´ algebra. ‡‡For a comprehensive review of general theories of this type, see [78]. Deformed General Relativity and Torsion 79 One proceeds by considering Lagrangians that only depend on the Ricci tensor, which is a one-form obtained by contracting the components of the Riemann curva- ture, written in the basis of one-forms given by the vielbein ei: Rica = Riciae i, Ricia = Rji j a, (3.9.3) where the curvature two-form is Rab = 1 2 Rij a be i ∧ ej. (3.9.4) One then splits the Ricci tensor into symmetric and antisymmetric part, symmetrising over a component (with respect to the given basis) index and a gl(n,R) index. The antisymmetric part can be interpreted as a spacetime two-form iea ( Rab ∧ eb ) , (3.9.5) where iea is interior multiplication with the vector ea, defined by being dual to the one-forms eb: eb(ea) = δ b a. (3.9.6) No such construction is possible for the symmetric part, which is normally more relevant in concrete constructions. The splitting itself seems depend on the choice of basis. General Relativity From a Constrained Topological Theory 80 Chapter 4 General Relativity From a Constrained Topological Theory “What, then, is time? If no one ask me, I know; if I wish to explain to him who asks, I know not.” [11] St. Augustine 4.1 Overview In this chapter, we shall investigate another possible formulation of general relativity in four dimensions, in terms of a constrained topological theory. Here the term “topological” means that the theory has a very large symmetry: All of its solutions are locally gauge equivalent, and there are no local degrees of freedom. Globally, not all solutions are necessarily gauge equivalent if there are topological obstructions. Starting from such a theory, we will use constraints that restrict the possible gauge transformations and hence allow for a less trivial theory, namely one with local degrees of freedom that are those of general relativity. The topological theory we consider is usually known as BF theory, deriving its name from the action S = ∫ Bab ∧ Fab[A] , (4.1.1) where F is the curvature of a G-connection A and B is a g-valued two-form (in our case G will be a rotation group). The equations of motion are F ab = 0 , ∇Bab = 0 , (4.1.2) General Relativity From a Constrained Topological Theory 81 where ∇ is a covariant exterior derivative involving the connection A. All flat con- nections are locally gauge equivalent, and the (local) equivalence of all solutions to the second equation can be seen [13] by noting that the transformation Bab 7→ Bab +∇ηab , (4.1.3) where ηab is a g-valued one-form, is a symmetry of the action (4.1.1). This transfor- mation is generated by the equation F ab = 0, in a canonical sense, and should be viewed as a generalised gauge transformation. It is a symmetry on the phase space of the theory. Locally, any solution to ∇Bab = 0 is of the form ∇ηab, so any solution is gauge equivalent to Bab = 0 in this generalised sense. The motivation for considering BF theory is that general relativity in three di- mensions is just of this form: The first order formulation can be written as S = ∫ abc ( ea ∧Rbc[ω]) , (4.1.4) and hence general relativity in three dimensions is topological. Since the quantisation of this system is relatively well understood (see [168] for Witten’s formulation as a Chern-Simons theory, and the book [31] for an overview over different routes to quantisation), and seems to have many conceptual aspects such as diffeomorphism and Lorentz invariance in common with the higher-dimensional case, one might hope to gain understanding of higher-dimensional general relativity from studying BF theory. On the other hand, there is of course a fundamental difference between a system with an infinite number of (local) degrees of freedom and one with only (usually a finite number of) global degrees of freedom, which makes it rather unclear whether this approach will be fruitful. It has been shown [56] that in any number of dimensions d ≥ 4 general relativity can be formulated as a constrained BF theory, where one adds to the BF action constraints quadratic in the field Bab, restricting it to be a wedge product of one- forms ea that can be interpreted as a frame field. In four dimensions, where one has to enforce Bab = abcde c ∧ ed (4.1.5) to reproduce the first order GR action (3.1.1), this formulation (in slightly different complex form) was first studied by Plebanski [138]. This is also the situation most studied in the spin foam approach to quantum gravity, where one tries to quantise general relativity by quantising a BF-type theory and then implementing constraints. General Relativity From a Constrained Topological Theory 82 In this chapter, we shall investigate a formulation which is in the spirit of the Plebanski formulation, but involving only linear constraints, of the type used re- cently as a new idea in the spin foam approach to quantum gravity. We identify both the continuum version of the linear simplicity constraints used in the quantum discrete context and a linear version of the quadratic volume constraints that are nec- essary to complete the reduction from the topological theory to gravity. We illustrate and discuss also the discrete counterpart of the same continuum linear constraints. Moreover, we show under which additional conditions the discrete volume constraints follow from the simplicity constraints, thus playing the role of secondary constraints. In the general context of the thesis, we see how a very large symmetry present in the original action (4.1.1) can be constrained appropriately so as to give a theory with rather different physics – one with local degrees of freedom. This symmetry acts neither on spacetime nor on fibre bundles, but on a symplectic manifold, namely the phase space of a dynamical system. 4.2 Actions for Gravity It should not be surprising that the equations of general relativity can be derived from several different action principles, leading to equivalent classical theories (in the case of pure gravity, at least). The statement that given equations of motion arise from different actions is the functional equivalent of saying that there are many functions with the same stationary points. Apart from constructing different actions as functionals of the same variables, one can also express the same physical content in terms of different variables (by performing a Legendre transformation, or other- wise), leading to even more possibilities for putting the same physics into a different mathematical form. This introduces several possible sources of ambiguity into quantisation; Feynman’s path integral approach to quantum mechanics is based on the fact that off-shell con- figurations, i.e. those that do not solve the equations of motion and are not stationary points of the action, contribute to transition amplitudes as well. Hence different func- tionals of the same variables, while giving the same equations of motion, may give different quantum theories. For formulations in terms of different variables it is even less clear whether the resulting quantum theories are related or not. One may see this as a problem or as a virtue; when a classical theory such as general relativity seems General Relativity From a Constrained Topological Theory 83 to resist conventional quantisation techniques, one may hope that one of the possi- ble classical reformulations offers more promise for quantisation. The most obvious example of this is Ashtekar new variables [10] as a possible new route to quantis- ing general relativity in four dimensions. In three dimensions, a number of consistent quantisation schemes for general relativity are known, leading to (at least apparently) inequivalent theories [31]. But already classically, in the case of general relativity equivalent formulations for pure gravity can give non-equivalent theories when coupled to matter. We en- countered an instance of this in chapter 3 when discussing Einstein-Cartan theory; in contrast to conventional general relativity, where torsion is set to zero by fiat, in Einstein-Cartan theory matter can act as a source of torsion. This is a non-equivalence on the level of equations of motion which can therefore be tested more easily by ex- periment. In this chapter we shall consider a reformulation of pure general relativity only. Hilbert- Palatini L(eµa , ωabµ ) - δS δω = 0⇒ ω = ω(e) Einstein- Hilbert L(eµa) 6 ? Legendre transform ADM Hamiltonian H(eia, pibj) Hilbert-Palatini Hamiltonian H(ωabi , piiab) 6 ? Legendre transform - solve 2nd class constraints BF-Plebanski (real form) L(ωabµ , Babµν ,Ξ) 6 ? δS δΞ = 0 ⇒ B = B(e) - δS δB = 0, δS δΞ = 0⇒ B = B(ω),Ξ = Ξ(ω) SO(3, 1)→ SO(3,C) CDJ Lagrangian L(η, ωabµ ) Ashtekar Hamiltonian H(Eai, Aai)- complex canonical transform @ @ @ILegendre transform Figure 4.1: Lagrangians and Hamiltonians for general relativity in four dimensions. (Es- sentially taken from [131].) In figure 4.1, essentially taken from the review [131], we show a few possible action principles for general relativity and their interrelations.∗ The most commonly used ∗Figure 4.1 is by no means meant to be a complete display of all known actions for general General Relativity From a Constrained Topological Theory 84 ones are the Einstein-Hilbert action [82], the Palatini first order formulation, and its modification proposed by Holst [83]. This last one is of special interest because it is the classical, covariant starting point for the canonical quantisation leading to loop quantum gravity [145]. Even if not the only possible useful one [150], a particularly popular action in covariant approaches to quantising gravity [122], like the spin foam [133] and group field theory approach [121], is the formulation as a constrained BF (or Plebanski [138, 101]) theory. Here one starts from topological BF theory [84] in four spacetime dimensions, and adds suitable constraints on the two-form B variables of the theory such that, on solutions of these constraints, the action reduces to the Palatini or Holst action for general relativity. We will summarise the idea behind this formulation in the following. In the original Plebanski formulation the constraints on the B variables are quadratic, and so are the discrete constraints that are then implemented in the spin foam models based on a simplicial discretisation. On the other hand, the most recent developments in the spin foam and group field theory approach to quantum gravity are based on a linear set of discrete constraints, which can be shown to be slightly stronger, in the restrictions they impose on the original BF configurations, than some of the original discrete quadratic constraints. Once more, we will detail this construction in the following. Here we investigate whether a formulation in terms of linear constraints is also possible in the classical continuum theory, and what it implies. We will see that the replacement of the so-called diagonal and cross-simplicity constraints with linear con- straints at the continuum level is relatively straightforward, after one has introduced new variables na forming a basis of three-forms at each point. One then needs addi- tional constraints corresponding to the volume constraints. We will see that one can linearise these constraints too. We then give a discrete version of these linear volume constraints (thinking of spacetime which is discretised in a triangulation), which bears a striking resemblance to the so-called “edge simplicity” constraints of [47]. We note that only certain linear combinations of the volume constraints one would naively write down are necessary to constrain the bivectors ΣAB(4) sufficiently. Similarly to the quadratic case, we will also see that when linear diagonal and off-diagonal con- straints hold everywhere in a 4-simplex, and one also imposes “closure” constraints on both bivectors Σab(4) (associated to triangles) and normals na(,) (associated to relativity; for instance, an interesting variant of the (complex) Plebanski formulation has recently been explored by Eyo Eyo Ita III (Ph.D. thesis, University of Cambridge (2010)). General Relativity From a Constrained Topological Theory 85 tetrahedra), the sufficient set of linear combinations of the linear volume constraints follows. This additional four-dimensional closure constraint on the normal vectors has, to the best of our knowledge, not been considered or implemented as an ad- ditional condition in the spin foam literature yet, although it does appear in some first order formulation of Regge calculus [35], and it plays also a role in the discrete analysis of [40]. 4.3 Constraining BF Theory Let us briefly review what is known at the classical continuum and discrete level, concerning the Plebanski formulation of classical gravity. We limit our considerations to the covariant, Lagrangian context, and to a very small subset of the available results, those which have been already of direct relevance for quantum gravity model building, especially in the spin foam context. For recent results in the canonical Hamiltonian setting, both continuum and discrete, see [28, 47, 160]. In this chapter we consider the Einstein-Hilbert-Palatini-Holst [82, 83] Lagrangian (without cosmological constant), a modification of the action (3.1.1) used earlier, SEHPH = 1 8piG ∫ R×R ( 1 2 abcd e a ∧ eb ∧Rcd[ω] + 1 γ ea ∧ eb ∧Rab[ω] ) , (4.3.1) where spacetime is assumed to be of the form R × R so that a (3+1) splitting can be performed, ωab is a G-connection one-form (the gauge group G is SO(3, 1) or SO(4), or an appropriate cover), Rab its curvature, and ea is an R4-valued one-form representing an orthonormal frame. The term involving γ, known as the Holst term, is not relevant classically; it is, up to a total derivative†, proportional to T a ∧ Ta, and hence does not modify the classical equation of motion T a = 0. When (4.3.1) is coupled to matter, the Holst term leads to a rescaling of the coupling of the matter spin density to torsion. Though not relevant classically, the Holst term is of fundamental importance in loop quantum gravity (LQG), where γ is known as the Barbero-Immirzi parame- ter, and more generally for any canonical formulation of gravity, as it modifies the symplectic structure of the theory. If one introduces a g-valued two-form, Bab = 1 8piG ( 1 2 abcd e c ∧ ed + 1 γ ea ∧ eb ) , (4.3.2) †For a discussion of the total derivative term see [117]. General Relativity From a Constrained Topological Theory 86 then the action (4.3.1) becomes‡ S = ∫ Bab ∧Rab[ω] + λαCα[B] , (4.3.3) i.e. it takes the form of a topological BF theory with additional constraints Cα which enforce that Bab is indeed of the form (4.3.2), and that are enforced by means of Lagrange multipliers λα. As said, BF theory without constraints is topological. Its equations of motion imply that ωab is flat and the covariant exterior derivative of Bab vanishes. Having no local degrees of freedom, the quantisation of such a theory is therefore rather simple and quite well understood. Inspired by this classical formulation, the main issue when trying to construct a quantum theory related to quantum gravity, in four dimensions, is then the correct implementation of appropriate constraints that lead to (4.3.2) for some set of one-forms ea, either at the level of quantum states or in a path integral formulation. Indeed, the bulk of the work in the spin foam approach [49, 57, 105, 133] (as well as in the group field theory formalism [16, 121, 123]), in recent years, has been devoted to this task. These constraints are also the subject of this chapter. To simplify the following calculations, we introduce another two-form field Σab, Σab ≡ 1 1− sγ2 ( Bab − γ 2 abcdB cd ) , (4.3.4) where s is the spacetime signature, s = −1 for G = SO(3, 1) and s = +1 for G = SO(4) (and we assume γ2 6= s)§. This is a linear redefinition which simplifies the constraint (4.3.2), Σab = 1 8piγG ea ∧ eb , (4.3.5) but leads to more terms in the action. The translation of all calculations from one set of variables to the other is usually straightforward. The traditional way to enforce the restriction (4.3.5), the one matching the origi- nal classical Plebanski formulation of gravity, was to add quadratic simplicity con- straints¶ to the action [45, 143], abcdΣ ab µνΣ cd ρσ = V µνρσ , (4.3.6) ‡We use indices from the beginning of the Greek alphabet such as α to denote abstract indices not associated with certain transformation groups. §For uniformity of the discussion, we shall in the following talk about “time” and use the label 0 even when the gauge group is SO(4) and the signature Riemannian. ¶“Simplicity” because a two-form that can be written as a wedge product of one-forms is called simple. General Relativity From a Constrained Topological Theory 87 where V can be expressed in terms of Σab by contracting (4.3.6) with µνρσ, to give: V = s 24 µνρσabcdΣ ab µνΣ cd ρσ. This is itself a reformulation of the original Plebanski con- straint, which would read: µνρσΣabµνΣ cd ρσ = V  abcd , (4.3.7) and is equivalent to the first under assumption that V 6= 0 everywhere. The version (4.3.6) has the advantage of permitting a much simpler discretisation and thus a more straightforward implementation within the spin foam formalism. Under the same assumption V 6= 0, there are the following four classes of solutions to (4.3.6): either Σab = ±ea ∧ eb or Σab = ±1 2 abcdE c ∧ Ed (4.3.8) for some set of one-forms Ea or ea (in the following, we reserve ea for one-forms satisfying the first relation, viewed as encoding the metric; since we consider pure gravity, the factor 8piγG can obviously be introduced by rescaling). One would like to select only the first class of solutions Σab = +ea ∧ eb, which, when substituted in the BF action, gives the Holst action (4.3.1). Classically, this is not a severe problem. As shown in [143], non-degenerate initial data of a solution of the form Σab = +ea∧eb generically remain within the same branch of solutions. The situation in the quantum theory, where one necessarily has contributions from all branches, is less clear. More troublesome, if V = 0, the field Σab does not permit a straightforward geometric interpretation at all. Since in the region of the phase space where V = 0, the theory is less constrained, and hence has more degrees of freedom, these non- geometric configurations should be expected to be dominating in a path integral [143], unless measure factors are such that this is avoided. Spin foam models are usually defined in a piecewise flat context, and spin foam amplitudes are defined for given simplicial complexes [133]. Therefore one is interested in identifying a discrete version of the above constraints that could be imposed at the level of each complex. The version (4.3.7) of the simplicity constraints admits only a rather involved discrete counterpart [45] and, upon quantisation, leads to the Reisenberger model [45, 142], which has so far received only limited attention. The discrete analogue of the constraints (4.3.6) led instead [45, 143] to the con- struction of the Barrett-Crane model [17], in the case in which the Barbero-Immirzi parameter is excluded from the original action (γ → ∞). The construction is ini- tially limited to a single 4-simplex, the convex hull of five points in R4 (R1,3, in the Lorentzian case) with the topology of a 4-ball, whose boundary is triangulated by the General Relativity From a Constrained Topological Theory 88 five tetrahedra identified by the five independent subsets of four such points, while subsets of three points identify the four triangles belonging to each of these five tetra- hedra, each of the triangles being shared by a pair of tetrahedra (see figure 4.2). One p1 p2 B B B B B B            Q Q Q Q Q p5 p3 p4          B B B B B B B B B         Z Z Z Z Z Z Z Z Figure 4.2: A 4-simplex is generated by five points. then associates a Lie algebra element (bivector) Σab4 ∈ so(4) ' ∧2R4 (similarly in the Lorentzian case) to each triangle 4 in a given triangulation by integrating the two-form Σab over 4. The task is then to constrain appropriately these Lie algebra variables (or their quantum counterpart) following the continuum treatment. It is useful to split the set of continuum equations (4.3.6) into two sets. Out of the 21 equations (4.3.6), one first identifies and imposes those 18 which have zero on the right-hand side (the “diagonal” and “off-diagonal” simplicity constraints), abcdΣ ab µνΣ cd µν = abcdΣ ab µνΣ cd µρ = 0 ∀µ, ν, ρ (no sum over µ, ν) . (4.3.9) This corresponds to the case if one or two of the indices of the two fields Σ coincide. At the discrete level, this translates into two triangles on which the same fields are discretised which either coincide or at least share a single edge, and thus belong to the same tetrahedron. Thus all bivectors Σab4 are required to satisfy abcdΣ ab 4Σ cd 4 = 0 (diagonal simplicity constraint) and abcdΣ ab 4 Σ cd 4′ = 0 for all 4,4′ sharing an edge (cross-simplicity constraint). These two sets of equations can be imposed at the level of each tetrahedron in the 4-simplex. The remaining three equations (the “volume” constraints) are equivalent to the requirement that: abcdΣ ab 01Σ cd 23 = −abcdΣab02Σcd13 = abcdΣab03Σcd12 ∝ V (Σ) , (4.3.10) General Relativity From a Constrained Topological Theory 89 and can be imposed at the discrete level as the requirement that, for each 4-simplex: abcdΣ ab 4 Σ cd 4′ = V for all 4,4′ not sharing an edge (volume constraints) (4.3.11) where V is defined by the above equation, and is interpreted, on the solutions of the constraints, as the volume of the 4-simplex. An additional condition on the bivectors is usually considered, namely the “clo- sure” constraint, which states that the sum of four bivectors corresponding to the faces of one tetrahedron is zero: ∑ 4⊂, Σab4 = 0 . (4.3.12) This constraint can be understood in two ways. One can either view it as the condition that the triangles described by the variables Σab4 close to form a tetrahedron [14], or as a consequence of the equations of motion. In a topologically trivial region such as the interior of a tetrahedron, a flat connection can be set to zero by a gauge transformation. Then using Stokes’ theorem, the integral over the equation dΣab = 0 can be written as ∫ , Σab = 0, which is the closure constraint. The canonical counterpart of this condition is the so-called Gauss constraint, which generates local gauge (rotation) transformations and is to be imposed on the quantum states of the theory. The same picture appears in three spacetime dimensions, where there are no sim- plicity constraints and one directly deals with a su(2) connection one-form ωa and an su(2)-valued (using su(2) ' R3) one-form ea. Here the equation dea = 0 is integrated over a (spacetime) triangle to give a closure constraint. The vectors associated to the edges of the triangle add up to zero, and thus have a consistent geometric in- terpretation as edge vectors in R3. In this sense, an n-form with vanishing exterior derivative and appropriate internal indices can be given a geometric interpretation as describing n-simplices closing up to form an (n + 1)-simplex. We shall encounter another instance of this statement later on. The closure constraint, being linear in the Σ’s and local in each tetrahedron, is obviously easier to impose at the discrete level, and in the quantum theory, than the volume constraints. Thus it is a useful fact that it can indeed be imposed instead of them. More precisely, it can be shown [105] that the volume constraints in each 4-simplex are implied if one has enforced the diagonal and cross-diagonal simplicity constraints, plus the closure conditions everywhere, i.e. in all the tetrahedra of the General Relativity From a Constrained Topological Theory 90 4-simplex (in general, i.e. for non-degenerate 4-simplices, involving tetrahedra be- longing to different “time slices”). From a canonical perspective, this observation is usually phrased as an interpretation of the volume constraints as “secondary con- straints” required to guarantee conservation of the other constraints (including the Gauss (closure) constraint) under time evolution. After a period of investigations, several potentially worrying issues have been put forward regarding the Barrett-Crane model [1, 19] (for a more recent analysis of the geometry of the Barrett-Crane model, see [16]), and have given impetus to the development of alternative spin foam models [49, 57]. These models are known to have nice semiclassical properties [18], and, importantly, generalise the spin foam setting to include the Barbero-Immirzi parameter at the quantum level (for an early attempt, see [107]), and thanks to this allow for a more direct contact with the canonical loop quantum gravity. Their study is still somewhat preliminary, but the above properties make them promising candidates for a quantum theory related to gravity. One of the central features of the new models is the replacement of the quadratic simplicity constraints (4.3.9) by linear constraints of the form na(,)Σ ab(4) = 0 ∀4 ⊂ , , (4.3.13) where na is the normal associated to the tetrahedron , and 4 is any of the faces of ,. It can be shown that these are lightly stronger than the discrete diagonal and off- diagonal quadratic simplicity constraints, and remove some of the discrete ambiguity in the solution for Σab: out of the classes of solutions (4.3.8), one can restrict to (the discrete version of) Σab = ±ea∧eb only. For a geometric analysis of these conditions in the discrete setting, see [40, 49, 57], and for a proof that the same discrete conditions can also lead to the Barrett-Crane model, see [16]. 4.4 Linear Constraints for BF-Plebanski Theory The purpose of this chapter is to investigate whether a formulation in terms of linear constraints is also possible in the classical continuum theory, and what it implies. Let us work backwards, at first. Assume that the two-form field Σab is of the form Σab = Ea ∧ Eb, and that the “frame field” (not necessarily associated with a metric) Ea is non-degenerate, i.e. that the matrix (Eaµ) is invertible. It follows that EρaΣ ab µν = δ ρ µE b ν − δρνEbµ . (4.4.1) General Relativity From a Constrained Topological Theory 91 In order to make a connection to the discrete setting it is more convenient to work with exterior powers of the cotangent bundle only (n-forms can be integrated over n- dimensional submanifolds). Hence we multiply (4.4.1) by ρστυ and insert the relation ρστυE ρ a = (detE µ a ) adefE d σE e τE f υ , which is true for invertible matrices, obtaining adefE d σE e τE f υΣ ab µν = (detE a µ) ( µστυE b ν − νστυEbµ ) . (4.4.2) One can define the three-form naστυ ≡ na[στυ] ≡ adefEdσEeτEfυ , so that (4.4.2) take the form naστυΣ ab µν = (detE a µ) ( µστυE b ν − νστυEbµ ) . (4.4.3) naστυ can be interpreted as a 3D volume form for the submanifold parametrised by (xσ, xτ , xυ) embedded in 4D spacetime, whose internal index gives the normal to this submanifold. If Ea are a basis of one-forms at each spacetime point, then na are a basis of three-forms at each spacetime point, and so one can choose to work either with one or the other set of variables. Clearly Ea can be reconstructed from na: 1 6 ρστυnaστυ = s(detE a µ)E ρ a = s 3 √ det ( 1 6 νστυnbστυ ) Eρa . (4.4.4) This means that the set of variables na(x) define a co-tetrad frame at any point of the spacetime manifold (for the discrete analogue of the above, see [40]). 4.4.1 Linearised Diagonal and Off-Diagonal Constraints So far we have just rewritten the equation we want to obtain for Σab. Let us now consider the implications of imposing (4.4.3) as constraints, where we restrict to those with zero right-hand side, i.e. those for which {µ, ν} ⊂ {σ, τ, υ}. These are half of the equations (4.4.3). This will identify the continuum analogue of the linear simplicity constraints. Claim 4.4.1 For a basis na of three-forms, the general solution to naστυΣ ab µν = 0 ∀{µ, ν} ⊂ {σ, τ, υ} (4.4.5) is Σabµν = GµνE [a µ E b] ν , (4.4.6) where Eaµ is defined in terms of naστυ as in (4.4.4), and so in particular non-degenerate, and Gµν = Gνµ and Gµµ = 0. (Obviously, as the variables Σ ab and Ea, the “coeffi- cients” Gµν are spacetime dependent.) General Relativity From a Constrained Topological Theory 92 Proof. First note that we can rewrite (4.4.5) as adefE d σE e τE f υΣ ab µν = 0 (4.4.7) with Eaµ defined by (4.4.4). Then E a µ by assumption defines a basis in the cotangent space, so that Σabµν = G χξ µνE a χE b ξ (4.4.8) for some coefficients Gχξµν with G χξ µν ≡ G[χξ][µν]. Substituting this into (4.4.7), we get 0 ! = adefE a χE d σE e τE f υG χξ µνE b ξ = χστυ det(E a µ)E b ξG χξ µν , (4.4.9) and since det(Eaµ) 6= 0 and Ebξ form a basis of (the internal) R4, this implies that χστυG χξ µν = 0 ∀{µ, ν} ⊂ {σ, τ, υ} . (4.4.10) It follows that Gχξµν = 0 unless {χ, ξ} = {µ, ν} and so Gχξµν ≡ δ[χµ δξ]ν Gµν .  By a linear redefinition eaµ = λµE a µ one might try to set some of the Gµν to a given value (usually ±1, but one might prefer ± 1 8piγG ), but it is clear that one needs two conditions on the Gµν for this to be possible. In the discrete context, one sets na(,) = (1, 0, 0, 0) for each , by a gauge trans- formation.‖ One could use some of the gauge freedom here to restrict the form of na: This amounts to finding a convenient parametrisation for the coset space GL(4)/SO(3, 1). Let us make the (usual) assumption that the normal to hypersur- faces {t = constant} is indeed timelike. Then one can use the boost part of SO(3, 1) to set na123 = (C, 0, 0, 0). The remaining SO(3) subgroup can then be used to make the (3× 3) matrix ni0στ , where i ∈ {1, 2, 3}, upper diagonal, so that one has the form naστυ ∼   ∗ ∗ ∗ ∗ 0 ∗ ∗ ∗ 0 0 ∗ ∗ 0 0 0 ∗   . (4.4.11) Clearly, when this form of naστυ is assumed, integrating the three-form na over a region where t is constant gives a vector in R4 that only has a time component. Its magnitude specifies the three-dimensional volume of such a region. ‖This seems to involve an implicit assumption, namely that there is a non-degenerate normal to each tetrahedron, as well. General Relativity From a Constrained Topological Theory 93 4.4.2 Linearised Volume Constraints As in the quadratic case, further constraints, in addition to the linear simplicity constraints (4.4.5), are needed to complete the identification Σab = ±ea ∧ eb. First of all, one can show the following. Claim 4.4.2 Under the assumption that all Gµν are non-zero, the necessary and sufficient conditions for the existence of a linear redefinition eaµ = λµE a µ, such that either Σab = cea ∧ eb or Σab = −cea ∧ eb, where c is a given positive number, are G12G03 = G01G23 = G13G02( 6= 0) . (4.4.12) Proof. Set c = 1; the extension to arbitrary c amounts to a further rescaling by√ c. Then the required redefinition is possible if and only if there exist λ0, . . . , λ3, such that either Gµν = λµλν for all µ 6= ν or Gµν = −λµλν for all µ 6= ν. Clearly (4.4.12) are necessary. They are also sufficient: Take λ1 = √∣∣∣∣G12G13G23 ∣∣∣∣, λ2 = sgn ( G12G13 G23 ) G12 λ1 , λ3 = sgn ( G12G13 G23 ) G13 λ1 , (4.4.13) which solves the equations for G12, G13 and G23 with sgn ( G12G13 G23 ) specifying the overall sign. The remaining three equations for G01, G02 and G03 are then solved by the two relations (4.4.12) and λ0 = sgn ( G12G13 G23 ) G01 λ1 . (4.4.14)  The assumption Gµν 6= 0 is necessary: One solution to (4.4.12) is G12 = G23 = G13 = 0 with the other Gµν non-zero, which cannot be expressed as Gµν = ±λµλν . Further constraints, in addition to the linear simplicity constraints (4.4.5), are needed to complete the identification Σab = ±ea ∧ eb. One possibility is to use the quadratic volume constraints (4.4.12). Take the three volume constraints (4.3.10), abcdΣ ab 01Σ cd 23 = −abcdΣab02Σcd13 = abcdΣab03Σcd12 , (4.4.15) and substitute the solution Σabµν = GµνE [a µ E b] ν of (4.4.5). This gives precisely (4.4.12). The non-degeneracy assumption needed for (4.4.12) is then the usual one, namely V 6= 0 in (4.3.6). General Relativity From a Constrained Topological Theory 94 This shows that imposing the linear version of the diagonal and off-diagonal sim- plicity constraints (4.4.5) together with the quadratic volume constraints (4.3.10) and a non-degeneracy assumption on Σab implies that Σab = ±cea ∧ eb (4.4.16) for some set of one-forms ea, where c > 0 can be chosen at will. Thus, linearising the diagonal and off-diagonal simplicity constraints means that two of the four types of solutions for Σab are removed, but on the other hand one needs to introduce a basis of three-forms na at each spacetime point, which is put in as an additional variable. One also still has to assume V 6= 0. There is also generically no evolution of initial data with V 6= 0 into a degenerate Σab with V = 0 and a non-geometric interpretation (this is part of the discussion of [143]). The geometry of the spacetime manifold is specified by ea and not by Ea which is only used to determine normals in the constraints. Alternatively, one might prefer to use a linear version of the volume constraints as well. Consider the original equation (4.4.3) naστυΣ ab µν = (detE a µ) ( µστυE b ν − νστυEbµ ) , (4.4.17) which was equivalent to Σab = Ea ∧ Eb for a basis of one-forms Ea. So far we only considered one half of these equations, namely those with {µ, ν} ⊂ {σ, τ, υ}. The other half have the form naντυΣ ab µν = (detE a µ)µντυE b ν , no sum over ν , (4.4.18) with µντυ 6= 0. One way to read these equations is as the requirement on the left-hand side to be totally antisymmetric in (µ, τ, υ): naντυΣ ab µν = naνυµΣ ab τν = naνµτΣ ab υν , no sum over ν . (4.4.19) We could again try to turn the argument around and impose (4.4.19) as constraints on a g-valued two-form Σab together with the linear simplicity constraints (4.4.5). Substituting the solution Σabµν = GµνE [a µ E b] ν of the linear simplicity constraints into (4.4.19) gives (after diving by a non-zero factor 1 2 det(Eaµ)) µντυGµνE b ν = τνυµGτνE b ν = υνµτGυνE b ν . (4.4.20) General Relativity From a Constrained Topological Theory 95 For µντυ 6= 0 this would imply Gµν = Gτν = Gυν . By Claim 4.4.2, imposing (4.4.19) for one fixed ν, say ν = 0, is generically not sufficient: If we know that G01 = G02 = G03 6= 0, we still have the condition G12 = G13 = G23 , (4.4.21) so that one needs more conditions of the form (4.4.19). These will then imply that all Gµν are equal, Gµν = c ′ for some c′ that could be positive, negative, or zero. One can absorb |c′| by an overall redefinition, so that one has Σab = ±c ea ∧ eb , (4.4.22) for any chosen c, as before. Note that here it is possible, if c′ = 0 at a point, that all ea are zero this point and so Σab = 0 as well. While this is a very degenerate geometry, it is still a geometry. While the conditions (4.4.19), imposed for all values of ν, are therefore sufficient to complete the identification of the two-form field Σab as ±cea∧eb, note that (4.4.19) is a massively redundant set of constraints: In order to obtain at most five relations on the coefficents Gµν (two relations (4.4.12) if all Gµν are nonzero), we are imposing eight vector equations! We have not exploited the fact that (4.4.20) is a multiple of one of the vectors Ebν , which are by assumption linearly independent. We could add several of the equations (4.4.19) for different ν, instead of considering all equations for different ν separately. Let us try to impose ∑ ν ∑ {µ,υ}6∈{ν,τ} naντυΣ ab µν = 0, τ ∈ {0, 1, 2, 3} fixed. (4.4.23) Again substituting the solution Σabµν = GµνE [a µ E b] ν of the linear simplicity constraints into (4.4.23), we obtain 1 2 det(Eaµ) ∑ ν ∑ {µ,υ}6∈{ν,τ} µντυGµνE b ν = 0, τ fixed, (4.4.24) which implies, by linear independence of the Ebν , that indeed Gµν = Gυν for all τ 6∈ {µ, ν, υ}. It is then sufficient to impose the constraint (4.4.23) for three different choices of τ , say τ = 0, 1, 2, so that we only need three vector equations instead of eight. By absorbing the constant Gµν = c ′ (all Gµν are equal) we rescale all Ea by the same factor to obtain the variables ea that will have the physical intepretation of frame General Relativity From a Constrained Topological Theory 96 fields encoding the metric geometry of spacetime. While in the case of quadratic vol- ume constraints the one-forms Ea, or alternatively the three-forms na, only specified the normals to submanifolds {xµ = constant}, for linear volume constraints they can be directly interpreted, up to a position-dependent normalisation, as specifying an orthonormal basis in the cotangent space. Note that this implies that one can assume a convenient normalisation for the one-forms Ea. Instead of just assuming non-degeneracy det(Eaµ) 6= 0, one could fix det(Eaµ) = 1. This is no restriction of the physical content of the theory as the E a, for both linear and quadratic volume constraints, only have a geometric interpretation after rescaling. One could then interpret Eaµ as a map into SL(4,R). For linear volume constraints, the relation between the normalised one-forms Ea and the variables ea that are interpreted as frame fields is a single function on spacetime which may be viewed as a “gauge” in the sense of Weyl [163]. In contrast to the case of the quadratic volume constraint, no non-degeneracy assumption on the two-form Σab is needed to enforce simplicity. One might get Σab = 0 in some region as a solution to the constraints, in which case the action for this region will be zero. This is analogous to a metric with vanishing determinant in general relativity and, in contrast to the requirement V 6= 0 outlined above, not an additional issue. Notice, however, that one still has to assume that the tetrad field Ea and, equivalently, the co-tetrad field na are non-degenerate, in order for the simplicity and volume constraints to imply (4.4.16). Failing this, one gets solutions of the constraints that admit no proper geometric interpretation. In the end, writing the action for BF theory in terms of Σab, S = ∫ Bab ∧Rab = ∫ Σab ∧Rab + γ 2 abcdΣ cd ∧Rab , (4.4.25) we substitute (4.4.16) into this action, which gives (setting c = 1 8piγG ) S = 1 8piG ∫ R×R σ(x) ( 1 2 abcde a ∧ eb ∧Rcd + 1 γ ea ∧ eb ∧Rab ) . (4.4.26) One is left with a field σ(x) that can take the values ±1, but in the classical theory one may again argue that if σ = 1 everywhere on an initial hypersurface, there will be no evolution into σ = −1. What we obtain is first order general relativity where one uses (det e) = ±| det e| instead of | det e| as a volume element in the action. If σ is continuous as classical fields usually are assumed to be, this differs from the action with | det e| by an overall sign at most. General Relativity From a Constrained Topological Theory 97 To summarise, we have identified both a linear version of the quadratic simplicity constraints and a linear version of the (quadratic) volume constraints in the con- tinuum, which can be used to reduce topological BF theory to 4D gravity in the continuum. We have found also that both linear versions are slightly stronger (i.e. more restrictive) than the corresponding quadratic constraints, so that the resulting constrained theory is likely to be closer to gravity at the quantum level than the one in which quadratic constraints are implemented. We now discuss the discrete counterpart of the constraints found above. 4.4.3 Discrete Linear Constraints and their Relations The discrete analogue of (4.4.5) is just the linear constraint used in [49, 57], as desired: na(,)Σ ab(4) = 0 ∀4 ⊂ , . (4.4.27) One could write down also a discrete version of (4.4.19), obtained in the natural way, demanding that within the same 4-simplex na(,)Σ ab(4′) = na(,′)Σab(4′′) (4.4.28) whenever 4′ 6⊂ , and 4′′ 6⊂ ,′ and the edge shared by 4′ and , is the same as that shared by ,′ and 4′′. In the following we adopt the notation of [105], where the tetrahedra in a given 4-simplex are labelled by A,B,C,D,E, so that triangles are represented by AB,AC, etc., and edges by combinations ABC,ABD, etc. The orientation of the triangles and tetrahedra in (4.4.28) is then fixed by the signs of the permutations of the letters, na(,A)Σ ab(4BC) = −na(,B)Σab(4AC) = na(,C)Σab(4AB), etc. (4.4.29) In analogy to the continuum case, it will be sufficient to impose, instead of the full set of conditions (4.4.29), certain linear combinations of (4.4.29) to complete the geometric interpretation of the bivectors Σab(4). The discrete analogue of the three continuum equations (4.4.23), where the index e was kept fixed, is to pick one of the tetrahedra and add those six equations out of (4.4.29) which involve triangles belonging to this tetrahedron. Starting with A, we impose the constraint ∑ {i,j}63A na(,i)Σ ab(4Aj) = 0, (4.4.30) General Relativity From a Constrained Topological Theory 98 and the equivalent conditions for the tetrahedra B to E, thereby needing to satisfy only five instead of 20 volume constraints. The above discrete formulation of the linearised volume constraints resembles strongly the edge simplicity constraints studied, in a canonical setting, in [47], and it imposes indeed the same restriction on the discrete data. However, it does not match exactly any of the various expressions given for these edge simplicity constraints in [47]. The correspondence between the two, therefore, deserves to be studied in more detail, given also that edge simplicity constraints have been shown to be crucial for the kinematical phase space of BF theory (and of loop gravity) to reduce to that of discrete gravity, in accordance with what we find here in a covariant setting. In spin foam models such as [49, 57], as we mentioned earlier, only the diagonal and off-diagonal simplicity constraints, but no quadratic volume constraints (4.3.10) are imposed. This is because in the discrete setting, one can use the closure con- straint (4.3.12), imposed in all the tetrahedra in a 4-simplex, to relate the (quadratic) simplicity constraints to the volume constraints, so that if the former are imposed everywhere the latter follows. Since the quadratic simplicity constraints follow from the linear ones, as can be easily checked, this argument is still valid if one uses linear simplicity constraints. One might hope that the sufficient set of linear volume constraints (4.4.30) would also follow from the linear simplicity constraints and the closure constraints. This is almost the case, but not quite. In fact, one more constraint should be added to simplicity and closure imposed in the five tetrahedra in the 4-simplex. This is a “4D closure” constraint of the form na(,A) + na(,B) + na(,C) + na(,D) + na(,E) = 0 , (4.4.31) where ,i are the (appropriately oriented) tetrahedra of a given 4-simplex. Just as for the usual closure constraint (4.3.12), there are two ways to understand why such a constraint must be imposed. Recall that if one demands the triangles described by discrete variables Σab4 close to form a tetrahedron, they have to satisfy (4.3.12). Alternatively, one can start with the continuum field equation ∇(ω)[µ Σabνρ] = 0, where ∇(ω) is the covariant derivative for the connection ωab, set the (flat) connection to zero by a gauge transformation, and integrate this over an infinitesimal 3-ball (whose triangulation is a tetrahedron). The new constraint (4.4.31) seems to be the analogous statement that tetrahedra close up to form a 4-simplex. By Hodge duality ∧1R4 ' ∧3R4 and any internal cov- General Relativity From a Constrained Topological Theory 99 ector na(,) can be mapped to a three-form; unlike for two-forms, any three-form can be written as E1 ∧ E2 ∧ E3 for some eα. Demanding that the tetrahedra described by these three-forms form a closed surface is then (4.4.31). Thus the simplicial geo- metric reasoning goes through also for this new constraint. In terms of the equations of motion of the theory, on the other hand, the only argument for the need of this constraint is the following. If ∇(ω)[µ Σabνρ] = 0 and we assume that Σab = ±ea ∧ eb, then it follows that ∇(ω)[µ naνρσ] = 0 as well. Integrating this equation (with the connection again set to zero) over a 4-ball (which can be thought of as our 4-simplex) whose boundary is a 3-sphere, triangulated by tetrahedra, then leads to (4.4.31). We then however have to assume simplicity of Σab. A more direct derivation of (4.4.31) from the equations of motion would be desirable. A B B B B B B B            Q Q Q Q Q E C D          B B B B B B B B B         Z Z Z Z Z Z Z Z Figure 4.3: Dual picture of a 4-simplex; points represent tetrahedra, lines represent trian- gles shared by adjacent tetrahedra. The role of this constraint, anyway, is the following. Consider the closure con- straint Σab(4AB) + Σab(4AC) + Σab(4AD) + Σab(4AE) = 0. (4.4.32) Contracting with na(,B) gives, using the linear simplicity constraint na(,B)Σ ab(4AB) = 0, na(,B)Σ ab(4AC) + na(,B)Σab(4AD) + na(,B)Σab(4AE) = 0. (4.4.33) Alternatively, one may start with the 4d closure constraint and contract with ΣAB(4AB) to get, again using the linear simplicity constraints, na(,C)Σ ab(4AB) + na(,D)Σab(4AB) + na(,E)Σab(4AB) = 0. (4.4.34) In total one obtains 20 + 10 = 30 equations of this kind that can be used to express some of the combinations na(,)Σ ab(4) in terms of others. Substituting the result- ing expressions into the five discrete volume constraints (4.4.30) one finds that the General Relativity From a Constrained Topological Theory 100 equations (4.4.30) indeed follow from the relations (4.4.33) and (4.4.34). We have seen in the continuum that the summed constraints (4.4.23) are sufficient to identify ΣAB = ±eA∧eB, and hence we find that in the discrete case the situation is analogous to the case of quadratic constraints in that a sufficient set of volume constraints can be viewed as secondary. To see more clearly what happens in both our construction and in the case of quadratic constraints analyzed in [105], note that in our linear case one could use the 3d and 4d closure constraints to express the variables na(,E) and Σ ab(4AE),Σab(4BE), Σab(4CE),Σab(4DE) in terms of the others. Taking the linear simplicity constraints into account, one is then left with twelve independent combinations na(,)Σ ab(4), just as in the continuum. In the continuum, we saw that one can impose the three additional constraints (4.4.23) on the twelve contractions naµτυΣ ab µν to complete the identification Σab = ±ea ∧ eb. In the discrete case, one has the following three addi- tional conditions coming from linear cross-simplicity constraints: 0 = na(,E)Σ ab(4AE) = na(,B)Σ ab(4AC) + na(,B)Σab(4AD) + na(,C)Σab(4AB) +na(,C)Σ ab(4AD) + na(,D)Σab(4AB) + na(,D)Σab(4AC) (4.4.35) and similar ones coming from na(,E)Σ ab(4BE) = 0 and na(,E)Σab(4CE) = 0. These are precisely the analogue of the continuum constraints (4.4.23). Similarly, in the case of quadratic simplicity constraints, one can use 3d closure to eliminate Σab(4AE),Σab(4BE),Σab(4CE),Σab(4DE). Then one observes that ad- ditional quadratic cross-simplicity constraints give expressions such as 0 = abcdΣ ab(4AE)Σcd(4BE) = abcdΣ ab(4AC)Σcd(4BD) + abcdΣab(4AD)Σcd(4BC) (4.4.36) which are equivalent to the desired (two) volume constraints. All of this is an exercise in solving a system of linear equations for which there might be a more simple and elegant description, but the upshot is the following. The sufficient set of linear volume constraints (4.4.23) does indeed follow from the linear simplicity constraints and the closure constraints, once one also imposes a four-dimensional closure constraint on the normals to tetrahedra that seems very natural in light of their geometric interpretation. Just as in the formulation in terms of quadratic simplicity constraints [105], the volume constraints can be viewed as General Relativity From a Constrained Topological Theory 101 secondary constraints that imply conservation of the simplicity constraints in time, or put differently, the volume constraints follow if the simplicity constraints hold everywhere. Once more this strenghtens the relationship between the discrete linear volume constraints we have identified and the edge simplicity constraints of [47]. 4.5 Lagrangian and Hamiltonian Formulation Let us briefly outline the Lagrangian formulation of 4D gravity resulting from our linear constraints added to BF theory. One adds the linear simplicity and volume constraints to the action of BF theory using Lagrange multipliers: S = ∫ d4x ( 1 4 µνρσΣabµνRabρσ[ω] + γ 8 µνρσabcdΣ ab µνR cd ρσ[ω] + Ξ µνστυ b naστυΣ ab µν ) , (4.5.1) where the Lagrange multiplier field Ξµνστυb satisfies Ξ µνστυ b ≡ Ξ[µν][στυ]b , and µτυΞ µνντυ b = 0 (no sum over ν) . (4.5.2) Indeed, varying with respect to Ξb then gives back the constraints naστυΣ ab µν =  0, {µ, ν} ⊂ {σ, τ, υ},µτυf bν , ν = σ (for some f bν) . (4.5.3) Note that the second line corresponds to the set of constraints (4.4.19) and not to the summed version (4.4.23), and that it is clearly sufficient for the geometric inter- pretation of Σab. The field equation from varying with respect to the connection ω is the usual ∇(ω)[µ Σabνρ] = 0 , (4.5.4) where ∇ is the covariant derivative for the connection ωab. The remaining equations involve the Lagrange multipliers, as would be expected: 1 4 µνρσRabρσ[ω] + γ 8 µνρσabcdR cd ρσ[ω] + Ξ µνστυ [b na]στυ = 0 , Ξ µνστυ b Σ ab µν = 0 . (4.5.5) We have seen that the constraints imply that Σab = ±ea ∧ eb, and when substituting this back into the action one will recover general relativity, modulo the possible sign ambiguity we have already discussed. We leave a complete Hamiltonian analysis of this theory to future work. However, we note a feature of the theory that follows directly from the use of linear constraints, and from the introduction of the additional variables na. General Relativity From a Constrained Topological Theory 102 As in unconstrained BF theory the initial dynamical variables will be the spatial part of the connection ωabk and its conjugate momentum P k ab ≡ 12ijkΣabij. We also saw that the equation of motion ∇(ω)[µ Σabνρ] = 0 is unaffected by the constraints. Hence there will be Gauss constraints of the form Gcd ≡ ∂iP cdi + ωceiP edi + ωdeiP cei (4.5.6) on the canonical momenta. Their role is to generate G gauge transformations. Looking at the action (4.5.1), one would already require that (4.5.6) should be modified to generate gauge transformations on the normals naστυ; (4.5.1) is only invariant under gauge transformations if the three-forms naστυ are transformed. The need for such a modification is also seen if one computes Poisson brackets between the linear simplicity and Gauss constraints. Define “smeared” constraints C[Ξ] := ∫ Ξij,στυb naστυijkP abk , G[Λ] := ∫ ΛcdGcd . (4.5.7) One then finds that {C[Ξ],G[Λ]} = − ∫ δC[Ξ] δP abk δG[Λ] δωabk = − ∫ Ξij,στυB naστυijk [ ΛadP bd k − ΛbdP adk ] = −C[Λ · Ξ]− ∫ Ξij,στυB Λ adnaστυijkP b d k , (4.5.8) where (Λ ·Ξ)ij,στυd = ΛdbΞij,στυb . The first term alone would imply that G[Λ] generates gauge transformations, but the second term is an unwanted extra piece. For G to be a generator of gauge transformations, it must be first class (i.e. commute with other constraints up to linear combinations of constraints). We can remedy this by adding the variables naµνρ to the phase space, together with their conjugate momenta pi aµνρ. Now we can define a new Gauss constraint G ′cd ≡ Gcd − n[cµνρpid]µνρ . (4.5.9) Then, computing the Poisson brackets of the new Gauss constraint with C[Ξ], one finds {C[Ξ],G ′[Λ]} = {C[Ξ],G[Λ]} − ∫ Ξij,στυB Λ cancστυijkP abk = −C[Λ · Ξ] , (4.5.10) as desired. We have however increased the number of phase space variables at each point by 32. General Relativity From a Constrained Topological Theory 103 A similar reformulation of the Gauss constraint, leading to a relaxation of the gauge invariance properties of spin network states, has been already suggested by the Hamiltonian analysis of the Plebanski theory [28], and it has been advocated in the loop quantum gravity context in [3, 106] as well as spin foam and group field theory context [2, 16, 123]. 4.6 Summary and Outlook We have investigated a formulation of classical BF-Plebanski theory where the con- straint Σab = ±ea ∧ eb, needed to reproduce general relativity in four dimensions, starting from topological BF theory, is imposed through constraints linear in the bivector field Σab. The discrete counterpart of a part of these linear constraints (the “simplicity constraints”), in fact, has proven very useful in the spin foam approach to quantum gravity [49, 57]. The corresponding continuum constraints have been easily identified, and can indeed be used to replace the quadratic “diagonal” and “off-diagonal” parts of the simplicity constraints appearing in the Plebanski formulation. As in the discrete case, one needed to introduce a new set of variables na which are assumed to form a basis of three-forms at each point of spacetime, and are slightly stronger than the quadratic constraints: they eliminate two of the four sectors of solutions that are present for quadratic constraints. In the second part of the analysis we found that the quadratic volume constraints of the Plebanski formulation, needed to complete the identification Σab = ±ea∧eb, can also be replaced by linear constraints, which again are stronger than their quadratic analogues. They do not require an additional non-degeneracy assumption on Σab. However, a non-degeneracy assumption on the three-forms na is still necessary, and only when this is imposed one can hope to eliminate all “non-geometric” degenerate configurations for Σab, which are feared to dominate the quantum theory in the case of quadratic volume constraints. Also, while for quadratic volume constraints the variables na merely specify normals to submanifolds {xµ = constant} and hence can be independently rescaled arbitrarily at each point, for linear volume constraints they directly specify, up to an overall rescaling, the frame field encoding the metric geometry, i.e. an orthonormal basis in the cotangent space at each spacetime point. We have then analysed the discrete (simplicial) translation of the linear constraints General Relativity From a Constrained Topological Theory 104 we identified. In the context of spin foams, the quadratic volume constraints follow from imposing the (quadratic) diagonal and off-diagonal simplicity constraints every- where together with closure constraints on the discrete variables Σab(4). We have shown a similar property for the linear volume constraints. If (linear) diagonal and off-diagonal simplicity constraints and closure constraints for both bivector variables Σab(4) and normals na(,) are imposed everywhere, a sufficient set of linear volume constraints follows. This means that “non-geometric” bivector configurations can- not appear if the additional closure constraint on the normals holds, and the same normals are assumed to be non-degenerate. We have not performed a complete Hamiltonian analysis of the resulting linear constrained BF action for gravity, but only noted that the use of linear simplicity and volume constraints immediately requires a modification of the usual Gauss constraint to generate a transformation of the normal 3-form variables na alongside that of the Σ’s; a similar relaxation of the Gauss constraint, which translates at the spin foam and discrete gravity level into a closure constraint for simplices, and in the canonical quantum gravity context into a generalisation of spin network states, has been suggested on more than one occasion in the literature [2, 3, 16, 106], even if its proper implementation at the quantum level has not been yet developed. On the classical level, therefore, a full Hamiltonian analysis of the constraints would be highly desirable. This would involve adding momenta for the components Σab0i , which are Lagrange multipliers in unconstrained BF theory, as well as those for the normals na we have introduced, so that all variables can transform nontrivially under G gauge transformations generated by a modified Gauss constraint, as shown. Still at the classical level, but with obvious implications for the quantisation, one aspect of our construction that deserves further work is the relation between the discretised linear volume constraints we have found and the edge simplicity constraints used in [47]. As noted, the two sets of constraints appear to be very similar, and their role in the classical theory is the same, in particular, they remove (partly) the non- geometric configurations from the configuration space (or phase space) of the theory and appear as “secondary” in the sense specified above. So it natural to conjecture that one is simply a reformulation of the other. The implications for the quantum theory are not only due to the dominant role that non-geometric configurations may play in the quantum theory, if not removed, but also in the fact that one discrete formulation of these constraints can actually be simpler to implement in a spin foam General Relativity From a Constrained Topological Theory 105 context than the other. The possible use of our findings in the spin foam and group field theory con- text, and more generally in any quantisation based on the formulation of gravity as a constrained BF theory, are in fact most interesting. In particular, it seems to be important to explore how a closure constraint on normals could be implemented into existing spin foam models, given that we found it to be necessary for the full imposition of the geometric constraints on the variables of topological BF. A conve- nient setting to do so could be the GFT formulation of [16], since there the simplicial geometry and the contact with classical actions is brought to the forefront. 106 Part II Gauge Symmetries and CP Violation SU(3) and its Quotients 107 Chapter 5 SU(3) and its Quotients 5.1 Introduction The Lie group that will play a central role in this second part of the thesis is the special unitary group SU(3). In the standard model of particle physics, this group plays two important roles: Firstly as the (colour) gauge group of quantum chromodynamics (QCD), the version of Yang-Mills theory describing the strong interactions, secondly as the group of flavour symmetry which relates different generations of quarks, fa- mously introduced in Gell-Mann’s “Eightfold Way” [60]. We shall in the following focus on the electroweak sector of the standard model [72, 146, 161], where the gauge group is SU(2) × U(1)/Z2 ' U(2) [132], and possible extensions of it, where one adds another SU(2) group acting on “right-handed” fields. We are interested in CP violating processes which can change quark flavour, which are again described by an element of SU(3), the Cabibbo-Kobayashi-Maskawa (CKM) matrix. Ambiguities in the definition of the bases with respect to which this matrix is defined mean that all matrices obtained from a given element of SU(3) by left or right multiplication by an element of the maximal Abelian subgroup U(1)2 should be regarded as equiva- lent to the original matrix. Hence the space of CKM matrices is the double quotient U(1)2\SU(3)/U(1)2. In chapter 6 we will make statements about statistics of the Jarlskog invariant J , a measure of the magnitude of CP violation, which assume a choice of measure on this quotient space. In this chapter we study the geometry of SU(3) and its quotients to motivate different possible choices of measure. We will also discuss measures on the space of Hermitian and complex 3×3 matrices whose construction relies on the knowledge of measures on quotients of SU(3), and SU(3) and its Quotients 108 which become important if one considers the space of quark mass matrices. Let us first introduce coordinates on SU(3) that we find convenient. The Lie algebra su(3) is generated by (i times) the Gell-Mann matrices λ1 =   0 1 0 1 0 0 0 0 0   , λ2 =   0 −i 0 i 0 0 0 0 0   , λ4 =   0 0 1 0 0 0 1 0 0   , λ5 =   0 0 −i 0 0 0 i 0 0   , λ6 =   0 0 0 0 0 1 0 1 0   , λ7 =   0 0 0 0 0 −i 0 i 0   , λ3 =   1 0 0 0 −1 0 0 0 0   , λ8 = 1√3   1 0 0 0 1 0 0 0 −2   . (5.1.1) To make the passage from SU(3) to the single or double quotients straightforward, we note that any element of SU(3) can be written as U = AL V AR , (5.1.2) where AL = e i 2 (3p−q)λ3+ i √ 3 2 (p+q)λ8 , AR = e itλ3+i √ 3rλ8 , (5.1.3) and V = eixλ7 e−iwλ3 eiyλ5 eiwλ3 eizλ2 , (5.1.4) where the ranges of the Euler angles x, y, z and the complex phases w, p, q, r, t are 0 ≤ x , y , z ≤ pi 2 , 0 ≤ w , p , q , r , t < 2pi . (5.1.5) Explicitly, the matrix V takes the form V =   cos y cos z cos y sin z e−iw sin y − cos x sin z − eiw sin x sin y cos z cos x cos z − eiw sin x sin y sin z sin x cos y sin x sin z − eiw cos x sin y cos z − sin x cos z − eiw cosx sin y sin z cos x cos y   , (5.1.6) whereas the diagonal matrices AL and AR are AL = diag(e 2ip, ei(q−p), e−i(p+q)) , AR = diag(ei(r+t), ei(r−t), e−2ir) . (5.1.7) It should be clear the orbits of the left action of the subgroup U(1)2 are parametrised by (p, q), while the orbits of the right action are parametrised by (r, t). Therefore one SU(3) and its Quotients 109 may use (x, y, z, w) to parametrise the double quotient U(1)2\SU(3)/U(1)2. Indeed, (5.1.6) is commonly used in the literature as a parametrisation of the CKM matrix. We may now obtain left-invariant forms on SU(3) by computing the Maurer- Cartan form U−1dU = iλa σa . (5.1.8) The resulting expressions are somewhat involved, and we use the usual complex nota- tion to slightly shorten the expressions (taking real and imaginary parts one recovers the σi which are all real), σ1 + iσ2 = ie 2it ω − e2it sin y cos(2z)(sinw dx+ cosw sin(2x) dq) + 1 4 e2it sin(2z) Ω , σ3 = 1 4 cos(2z) Ω + dt+ sin y sin(2z) (sinw dx+ cosw sin(2x) dq) , σ4 + iσ5 = ie i(3r+t+w) cos z dy − cos y sin z(iei(3r+t) dx+ ei(3r+t) sin(2x) dq) +ei(3r+t+w) cos y sin y cos z (3 dp− dw + cos(2x) dq) , σ6 + iσ7 = ie i(3r−t+w) sin z dy + cos y cos z(iei(3r−t) dx+ ei(3r−t) sin(2x) dq) +ei(3r−t+w) cos y sin y sin z(3 dp− dw + cos(2x) dq) , (5.1.9) σ8 = √ 3 4 [−dp+ 4 dr + dw + 2 dq cos(2x) cos2 y + (3 dp− dw) cos(2y)] , where we defined the 1-forms ω = dz + dx cosw sin y − 2 dq cos x sinw sin x sin y , Ω = 3 dp+ dw + dq cos(2x)(cos(2y)− 3) + (3 dp− dw) cos(2y) . (5.1.10) A general left-invariant metric on the group will have the form ds2 = gabσaσb (5.1.11) for some constant symmetric matrix gab. Depending on the choice of gab, the resulting metric will have a certain subgroup of SU(3) as isometries acting by right multipli- cation on SU(3). The possible cases are detailed in [41]. The most symmetric case is the bi-invariant metric, invariant under SU(3)L × SU(3)R, which results from taking gab proportional to δab. Since we will be mainly interested in the measure, we note that for any choice of gab the (left-invariant) volume form is, up to an (irrelevant) overall constant, µ SU(3) l.inv. ∝ σ1 ∧ σ2 ∧ σ3 ∧ σ4 ∧ σ5 ∧ σ6 ∧ σ7 ∧ σ8 (5.1.12) = 3 √ 3 2 sin(2x) sin y cos3 y sin(2z) dx ∧ dy ∧ dz ∧ dw ∧ dp ∧ dq ∧ dr ∧ dt , SU(3) and its Quotients 110 which is independent of the complex phases w, p, q, r and t. Clearly left invariance fixes the volume element up to an overall constant: Pick a group element g (say the identity), then the space of 8-forms at g is one-dimensional. The volume form at any other group element is then determined to be the push-forward of the volume form at g under left translation. The volume form (5.1.12) will be the central quantity in many of the following calculations. One could of course have started demanding right-invariance, computing the right- invariant forms from dU U−1 = iλa τa . (5.1.13) If the left- and right-invariant measures are identical the group is called unimodular. Compact and semisimple groups are unimodular. Hence we can use left and right invariance as interchangeable conditions on a natural measure on SU(3). If G is compact and semisimple, (the negative of) the Killing form is a bi-invariant Riemannian metric which is also an Einstein metric [9]. It may therefore be regarded as the most natural choice of metric on such a group. For a (complex) matrix Lie group this metric is defined by ds2 = Tr(dg dg†) . (5.1.14) It can also be viewed as the induced metric on the group G, interpreted as a sub- manifold of Cn×n with flat metric. (The same considerations are of course true for a group of real matrices.) We will use this metric as the most natural choice later on. It does give the same measure as any left-invariant metric, as we have noted. For completeness, we give its form explicitly for SU(3): ds2 = 1 2 Tr ( dU dU † ) = δabσaσb = 3 dp2 + dq2 + 3 dr2 + dt2 + 3 2 (3 cos(2y)− 1) dp dr +3 cos2 y (cos(2z) dp dt+ cos(2x) dq dr) + 1 2 [cos(2x) cos(2z) (cos(2y)− 3) + 4 sin(2x) sin(2z) sin y cosw] dq dt − sin2 y (3 dp− 3 dr + cos(2x) dq − cos(2z) dt)dw +2 sin y sinw (sin(2z) dt dx− sin(2x) dq dz) +dx2 + dy2 + dz2 + sin2 y dw2 + 2 sin y cosw dx dz . (5.1.15) SU(3) and its Quotients 111 5.2 The Homogeneous Space SU(3)/U(1)2 The coset space SU(3)/U(1)2 is a reductive geometry as explained in section 1.1: The splitting of su(3) into an Abelian subalgebra u(1)2 and the orthogonal subspace is invariant under the adjoint action of SU(3). We may hence identify an appropriate subset of generators of SU(3) and use the corresponding one-forms to define metrics on the homogeneous space SU(3)/U(1)2. The one-forms σ3 and σ8 correspond to the Gell-Mann matrices generating the Abelian subgroup U(1)2. We can construct left-invariant metrics on SU(3)/U(1)2 from the remaining six left-invariant forms σ1, σ2, σ4, . . . , σ7. The left group action of SU(3) on the homogeneous space again guarantees the existence of a unique (up to a constant) left-invariant volume form, and we find that this unique volume form is µ SU(3)/U(1)2 l.inv. ∝ σ1 ∧ σ2 ∧ σ4 ∧ σ5 ∧ σ6 ∧ σ7 (5.2.1) = 3 √ 3 2 sin(2x) sin y cos3 y sin(2z) dx ∧ dy ∧ dz ∧ dw ∧ dp ∧ dq , where we note that, due to the SU(3) decomposition (5.1.2), we may parametrise the quotient SU(3)/U(1)2 by the SU(3) coordinates (x, y, z, w, p, q). This is of course the same volume form as the one obtained by starting with the left-invariant volume form on SU(3) and integrating over the coordinates r and t, or by just considering a slice {r = constant, t = constant} in SU(3) with the measure (5.1.12). (We shall see in the next section that this need not be true if one reduces to the double quotient U(1)2\SU(3)/U(1)2.) There is an additional possible source of non-uniqueness of the measure, namely a possible ambiguity of the definition of the same manifold as different homogeneous spaces X = G/H = G′/H ′. In practice, this does not seem to lead to different results for the measure. It should be noted that the homogeneous space considered here, where the sub- group is the maximal Abelian subgroup, the “maximal torus”, of a compact group, is what is known as a flag manifold. Such a manifold is naturally a Ka¨hler manifold. For any G-invariant complex structure on a flag manifold, there is a unique G-invariant Ka¨hler-Einstein metric [9]. This metric may then be regarded as the most natural choice on such a homogeneous space. An explicit construction for such metrics on spaces SU(M)/U(1)M−1, where U(1)M−1 is the maximal torus of SU(M), is given in [136]. The case M = 6 might be of interest for neutrino mixing; here the analogue of the CKM matrix is the Maki-Nakagawa-Sakata matrix [114], usually taken to be SU(3) and its Quotients 112 an element of the single quotient U(1)2\SU(3). In the see-saw mechanism one adds very heavy right-handed neutrinos, and the most general mixing matrix would be an element of U(1)5\SU(6). The construction of [136] would be the starting point of an analysis of this quotient, which is beyond the scope of this thesis. 5.3 Metrics on U(1)2\SU(3)/U(1)2 The double quotient U(1)2\SU(3)/U(1)2 does not have a natural SU(3) group action, because the left and right actions of SU(3) on itself do not commute with the Abelian subgroup U(1)2 and hence do not induce a group action on the double quotient: If g and g′ are in the same equivalence class, [g] = [g′], then g = hg′h′ for some h, h′ ∈ U(1)2. For a group action to be defined on the double quotient, one would need [j · g] = [j · g′], j · g = k(j · g′)k′ = k(j · h−1gh′−1)k′ for some k, k′ ∈ U(1)2. This can only be done for those j that commute with h−1. It is therefore less straightforward than in the previous cases to find a geometrically motivated measure. The simplest way of reducing the left-invariant measure on the group SU(3) or the homogeneous space SU(3)/U(1)2 to the double quotient would be to just integrate (5.2.1) over the coordinates p and q, which is trivial since the volume form (5.2.1) is independent of p and q. One would therefore essentially use the same measure for the double quotient. Alternatively, one might start from a metric on either the group or the homoge- neous space, reduce it in some way to the double quotient, and then determine the associated volume form, but this introduces two possible sources of ambiguity; the choice of metric and the choice of method of reduction. We will see that indeed the measure on the double quotient is vastly non-unique. For example, one may be tempted to identify the double quotient with the sub- manifold p = q = r = t = 0 of SU(3), take a symmetric metric on SU(3), say the bi-invariant metric (5.1.15), and set p = q = r = t = 0. This approach was indeed followed by Ozsva´th and Schu¨cking [126], and leads to an appealingly simple form of the metric ds2 = dx2 + dy2 + dz2 + 2 sin y cosw dx dz + sin2 y dw2, (5.3.1) but seems not to be justified geometrically: It corresponds to a gauge choice, a choice of section in the orbits of U(1)2 × U(1)2 over the double quotient, and introduces SU(3) and its Quotients 113 symmetries (three commuting Killing vectors) into the metric which should not be present. It seems much better motivated geometrically to do a Kaluza-Klein reduction of a metric, i.e. to project it orthogonally to the orbits of U(1)2. One writes a given metric on SU(3) (say) as ds2 = hij(x) (dy i + Aiµ(x)dx µ)(dyj + Ajν(x)dx ν) + g˜µν(x)dx µ dxν , (5.3.2) and identifies g˜µν as a metric on the quotient space parametrised by (x µ). To understand what is going wrong in (5.3.1), we compare with the simpler case of SU(2). The bi-invariant metric on SU(2) is ds2 = (dψ + cos θ dφ)2 + dθ2 + sin2 θ dφ2 , (5.3.3) where ∂ ∂φ generates the left action of U(1) and ∂ ∂ψ generates the right action of U(1). Projecting the metric orthogonally to the orbits of right translations, a` la Kaluza- Klein, gives the round metric ds2 = dθ2 + sin2 θ dφ2 (5.3.4) on SU(2)/U(1) = S2. By contrast, simply setting dψ = 0 (the analogue of the construction of Ozsva´th and Schu¨cking) instead gives the flat metric ds2 = dθ2 + dφ2 . (5.3.5) The round metric (5.3.4) is invariant under SO(3). The flat metric (5.3.5) appears to be invariant under the Euclidean group, with ∂ ∂θ and ∂ ∂φ having the appearance of translations, but these are only local symmetries since φ is a periodic coordinate and θ lies in an interval. The example of SU(2) also illustrates the difference between taking the left- invariant measure on the homogeneous space and a measure obtained by Kaluza-Klein reduction of a metric to a biquotient. The biquotient U(1)\SU(2)/U(1) = U(1)\S2 is just an interval. Its metric becomes, after performing another Kaluza-Klein reduction of (5.3.4), ds2 = dθ2 . (5.3.6) The measure would be dθ, and not dθ sin θ as obtained by integrating a function f(θ) over the coordinate φ. It is apparent from this simple example that there are inequivalent ways of calculating integrals of a function on a right quotient that is invariant under the left group action; namely, one can either reduce the metric to SU(3) and its Quotients 114 obtain a measure on the double quotient or take the measure on the single quotient and integrate out the left phases. Performing a Kaluza-Klein reduction of the bi-invariant metric (5.1.15) we find a rather complicated metric on the double quotient, whose determinant gives the measure √ det g˜ ∝ sin 2x sin 2z sin y cos2 y · ( (sin2 2x+ sin2 2z) sin2 y + 1 8 (5 cos 2y − 3) sin2 2x sin2 2z + 1 2 sin 4x sin 4z sin3 y cosw + 1 8 (3 cos 2y − 5) sin2 2x sin2 2z sin2 y cos2w )− 1 2 . (5.3.7) One might think that, starting from a more general (“squashed”) left-invariant metric on the homogeneous space SU(3)/U(1)2, there might be some which after Kaluza-Klein reduction give rise to a simpler expression for the measure on the biquo- tient, but we have only found more complicated ones. 5.4 Hermitian and Complex Matrices In chapter 6 we will require measures on the space of either all Hermitian or all complex 3× 3 matrices which will appear as the space of quark mass matrices in the standard model of particle physics or certain extensions of it. These will involve the left- (or right-)invariant measure on the homogeneous space SU(3)/U(1)2, as we will see shortly. Let us start with the space of 3× 3 Hermitian matrices. It has a natural group action by U(3) by conjugation, corresponding to a change of basis. We would like to determine a metric invariant under this action. Any Hermitian matrix can be diagonalised, i.e. written in terms of a real diagonal and a unitary matrix: MH = U †DU , (5.4.1) It should be clear that U is only defined up to left multiplication by elements of U(1)2; U should be regarded as an element of the homogeneous space U(1)2\SU(3) ' SU(3)/U(1)2. The most symmetric measure on the space of Hermitian matrices is induced by the bi-invariant metric ds2 = Tr(dMH dMH) = Tr (dD dD) + 2Tr (( dU U †D )2 − (dU U †)2D2) (5.4.2) SU(3) and its Quotients 115 which is clearly invariant under the action of U(3) by conjugation. Define right- invariant one-forms τa on SU(3) by dU U † = iλa τa , (5.4.3) this becomes∗ [with D ≡ diag(D1, D2, D3)] ds2 = Tr (dD · dD)− 2τaτbTr (λa[D, λb]D) = dD21 + dD 2 2 + dD 2 3 + 2 { (D1 −D2)2(τ 21 + τ 22 ) + (D1 −D3)2(τ 24 + τ 25 ) +(D2 −D3)2(τ 26 + τ 27 ) } . (5.4.4) The corresponding volume form is (D1−D2)2(D1−D3)2(D2−D3)2 dD1∧dD2∧dD3∧ τ1∧ τ2∧ τ4∧ τ5∧ τ6∧ τ7 . (5.4.5) We have explained above that any left- or right-invariant measure on the coset U(1)2\SU(3) must be given by (5.2.1) and hence there is no need to compute the τa explicitly. We obtain a Riemannian measure DMH := (D1 −D2)2(D1 −D3)2(D2 −D3)2 sin(2x) cos3 y sin y sin(2z) dD1 dD2 dD3 × dx dy dz dw dr dt (5.4.6) on the space of Hermitian 3×3 matrices. Note that we use coordinates (x, y, z, w, r, t) on U(1)2\SU(3) according to the decomposition (5.1.2). The measure (5.4.6) favours strongly non-coinciding eigenvalues. From (5.4.1), it is apparent that each Hermitian matrix with three distinct eigen- values is associated with six different elements of R3 × U(1)2\SU(3), related by the action of the discrete group S3: MH = U †DU = (U †P−1)PDP−1(PU) =: U˜ †D˜U˜ , P ∈ S3 , (5.4.7) where S3 is the symmetric group of degree 3 (the dihedral group of order 6, sometimes denoted by D3 or D6) which permutes the canonical basis vectors of R 3. The set of matrices with coinciding eigenvalues has zero measure and hence can be ignored in the present discussion. Thus we need to consider the space R3× (U(1)2×S3)\SU(3) instead†, restricting the coordinates on the quotient U(1)2\SU(3) to an appropriate range to pick one of ∗Compare with the corresponding result for real matrices given in [71] †Note that the space of Hermitian matrices is not topologically identified with this space; we have discarded a subset of Hermitian matrices which has zero measure for our construction. SU(3) and its Quotients 116 the six matrices related by the S3 action. We can use the fact that the S3 action permutes the rows of an SU(3) matrix to demand that the elements of the third column (cf. (5.1.6); note that absolute values will not depend on rephasing, i.e. right multiplication by an element of U(1)2) satisfy the relation | sin y| ≤ | sin x cos y| ≤ | cos x cos y| , (5.4.8) which restricts the coordinates x and y to 0 ≤ y ≤ arctan(sin x) , 0 ≤ x ≤ pi 4 . (5.4.9) Using the right-invariant measure (5.2.1) on U(1)2\SU(3), we see that this region has precisely one-sixth of the total volume: pi/2∫ 0 dz pi/4∫ 0 dx arctan(sinx)∫ 0 dy sin 2x cos3 y sin y sin 2z pi/2∫ 0 dz pi/2∫ 0 dx pi/2∫ 0 dy sin 2x cos3 y sin y sin 2z = 1 6 . (5.4.10) Notice that while the total volume of (U(1)2 × S3)\SU(3) will be finite, the in- tegrals over the R3 part will diverge; the eigenvalues of arbitrary Hermitian matrices can of course be arbitrarily large. One would have to introduce some kind of cutoff to make integrals over the space of Hermitian matrices convergent, as we will do later on. The space of 3 × 3 complex matrices can be discussed along similar lines. Here, one has a left and right action by independent elements of U(3). We should therefore determine a metric invariant under both of these group actions. An arbitrary complex matrix can be represented as Mc = U † LDUR , (5.4.11) where UL, UR are unitary and D = diag(D1, D2, D3) is real diagonal. UL and UR are only defined up to simultaneous left multiplication by a diagonal unitary matrix UL → AUL , UR → AUR , A ∈ U(1)3 (5.4.12) which so that one has the correct number of real parameters, only 18 instead of 21. Since here the signs of the elements Di can be changed by a different choice of UL and UR, and one can permute the Di as for Hermitian matrices, there are additional discrete ambiguities given by elements of the group S3 × Z32. We will see shortly SU(3) and its Quotients 117 that the measure, similar to the measure for Hermitian matrices, vanishes whenever elements of D coincide up to sign, so we can restrict to matrices with D21 6= D22 6= D23 6= D21. Hence we can identify this subspace of the space of 3× 3 complex matrices with R3 × (U(1)3 × S3 × Z32)\(U(3)× U(3)) ' R3+ × (U(1)3 × S3)\(U(3)× U(3)). The metric we will use to determine a measure is ds2 = Tr(dMc dM † c ) , (5.4.13) which clearly is invariant under M → OMO′ for O,O′ ∈ U(3). To evaluate this, use dMc = −U †L dUL U †LDUR + U †L dDUR + U †LDdUR , dM †c = −U †R dUR U †RDUL + U †R dDUL + U †RDdUL (5.4.14) and [D, dD] = 0 to obtain Tr(dMc dM † c ) = Tr(dD dD)− Tr(D2(dUL U †L)2)− Tr(D2(dUR U †R)2) +2Tr(DdUL U † LDdUR U † R) . (5.4.15) We introduce right-invariant one-forms dULU † L = iλa τ a L , dURU † R = iλb τ b R , (5.4.16) where λ1, . . . , λ8 are the Gell-Mann matrices, and λ9 = √ 2 3   1 0 0 0 1 0 0 0 1   , (5.4.17) so that iλa are a basis for the Lie algebra u(3). Then Tr(dMc dM † c ) = Tr(dD dD)+Tr(D 2 λ(aλb))(τ a Lτ b L+τ a Rτ b R)−Tr(DλaDλb)(τaLτ bR+τaRτ bL) . (5.4.18) SU(3) and its Quotients 118 The only nonvanishing traces in (5.4.18) are Tr(D2λ1λ1) = Tr(D 2λ2λ2) = Tr(D 2λ3λ3) = Tr(Dλ3Dλ3) = D 2 1 +D 2 2 , Tr(D2λ(3λ8)) = Tr(Dλ(3Dλ8)) = 1√ 3 (D21 −D22) , Tr(D2λ(3λ9)) = Tr(Dλ(3Dλ9)) = √ 2 3 (D21 −D22) , Tr(D2λ(8λ9)) = Tr(Dλ(8Dλ9)) = √ 2 3 (D21 +D 2 2 − 2D23) , Tr(D2λ4λ4) = Tr(D 2λ5λ5) = D 2 1 +D 2 3 , Tr(D2λ6λ6) = Tr(D 2λ7λ7) = D 2 2 +D 2 3 , Tr(D2λ8λ8) = Tr(Dλ8Dλ8) = 1 3 (D21 +D 2 2 + 4D 2 3) , Tr(D2λ9λ9) = Tr(Dλ9Dλ9) = 2 3 (D21 +D 2 2 +D 2 3) , Tr(Dλ1Dλ1) = Tr(Dλ2Dλ2) = 2D1D2 , Tr(D2λ4λ4) = Tr(D 2λ5λ5) = 2D1D3 , Tr(D2λ6λ6) = Tr(D 2λ7λ7) = 2D2D3 . (5.4.19) The metric can be written in the form ds2 = dD21 + dD 2 2 + dD 2 3 + 1 2 (D1 −D2)2(τ 1L + τ 1R)2 + 1 2 (D1 +D2) 2(τ 1L − τ 1R)2 + 1 2 [ (D1 −D2)2(τ 2L + τ 2R)2 + (D1 +D2)2(τ 2L − τ 2R)2 +(D1 −D3)2(τ 4L + τ 4R)2 + (D1 +D3)2(τ 4L − τ 4R)2 ] + 1 2 [ (D1 −D3)2(τ 5L + τ 5R)2 + (D1 +D3)2(τ 5L − τ 5R)2 +(D2 −D3)2(τ 6L + τ 6R)2 + (D2 +D3)2(τ 6L − τ 6R)2 ] + 1 2 (D2 −D3)2(τ 7L + τ 7R)2 + 1 2 (D2 +D3) 2(τ 7L − τ 7R)2 +2D21 ( 1√ 2 (τ 3L − τ 3R) + 1√ 6 (τ 8L − τ 8R) + 1√ 3 (τ 9L − τ 9R) )2 +2D22 ( − 1√ 2 (τ 3L − τ 3R) + 1√ 6 (τ 8L − τ 8R) + 1√ 3 (τ 9L − τ 9R) )2 +2D23 (√ 2 3 (τ 8L − τ 8R) + 1√ 3 (τ 9L − τ 9R) )2 . (5.4.20) The fact that the metric only depends on τ 3L− τ 3R etc., and not on τ 3L+ τ 3R etc., again reflects the U(1)3 that has to be factored out. One finds that the volume form is SU(3) and its Quotients 119 proportional to (D21−D22)2(D21−D23)2(D22−D23)2|D1D2D3| dD1∧dD2∧dD3∧τ 1L∧τ 1R∧. . .∧(τ 8L−τ 8R)∧(τ 9L−τ 9R) . (5.4.21) Note that this expression only depends on the absolute values of Di, as expected. Since the range of the Di is infinite, integration over these coordinates will give an infinity, as was the case for Hermitian matrices. Note the main difference to the measure (5.4.6) in the different powers of Di. Now assume that all quantities we are interested in are independent of the pa- rameters on UR, which will be the case in the discussion of chapter 6. Then we can integrate over these coordinates, obtaining a constant which is irrelevant in the averaging process. We are then left with integrating over the space of possible matrices UL, the coset U(1)3\U(3) = U(1)2\SU(3), and the volume form is proportional to (D21−D22)2(D21−D23)2(D22−D23)2|D1D2D3| dD1∧dD2∧dD3∧τ 1L∧τ 2L∧τ 4L∧τ 5L∧τ 6L∧τ 7L . (5.4.22) In terms of the coordinates on U(1)2\SU(3) introduced in (5.1.2), the wedge product of right-invariant forms gives the usual measure on SU(3), so that we finally get DMc = ∏ i