Вы находитесь на странице: 1из 397
THE LARGE SCALE STRUCTURE OF SPACE-TIME 8. W. HAWKING, F.RS. Lacan Profs of Mathemati inthe Unvray of Cambridge and Fellow of Gonville and Caius AND G.F.R. ELLIS Professor of Applied Mathematics, University of Cape Town CAMBRIDGE UNIVERSITY PRESS CAMBRIDGE NEW YORK PORT CHESTER MELBOURNE SYDNEY Contents Preface 1 The role of gravity 2 Differential geometry 2.1 Manifolds 2.2 Vectors and tensors 2.3 Maps of manifolds 2.4 Exterior differentiation and the Lie derivative 2.5 Covariant differentiation and the curvature tensor 2.6 The metric 2.7 Hypersurfaces 2.8 The volume element and Gauss’ theorem 2.9 Fibre bundles 3 General Relativity 3.1 The space-time manifold 3.2 The matter fields 3.3 Lagrangian formulation 3.4 The field equations 4 The physical significance of curvature 4.1 Timelike curves 4.2 Null curves x 4.3 Energy conditions 4.4 Conjugate points 4.5 Variation of arc-length 5 Exact solutions 5.1 Minkowski space-time 5.2 De Sitter and anti-de Sitter spacetimes 5.8 Robertson-Walker spaces 5.4 Spatially homogeneous cosmological models C vii) page xi 1 10 at 15 22 2 30 36 44 47 50 56 56 59 64 u 78 18 86 88 96 102 117 118 124 134 142 5.5 5.6 5.7 5.8 5.9 6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 6.9 TA a2) 13 14 15 1.6 11 8.4 8.2 8.3 8.4 8.5 9.1 9.2 9.3 CONTENTS ‘The Schwarzschild and Reissner-Nordstrém solutions page 149 The Kerr solution Gédel’s universe Taub-NUT space Further exact solutions Causal structure Orientability Causal curves Achronal boundaries Causality conditions Cauchy developments Global hyperbolicity ‘The existence of geodesics The causal boundary of space-time Asymptotically simple spaces The Cauchy problem in General Relativity ‘The nature of the problem The reduced Einstein equations ‘The initial data Second order hyperbolic equations The existence and uniqueness of developments for the empty space Einstein equations ‘The maximal development and stability The Einstein equations with matter Space-time singularities ‘The definition of singularities Singularity theorems ‘The description of singularities ‘The character of the singularities Imprisoned incompleteness Gravitational collapse and black holes Stellar collapse Black holes The final state of black holes 161 168 170 178 180 181 182 186 189 201 208 213 217 221 226 227 228 231 233 24d 249 254 256 256 261 276 284 289 299 209 308 323 CONTENTS 10 The initial singularity in the universe 10.1 The expansion of the universe 10.2 The nature and implications of singularities Appendix A: ‘Translation of an essay by P. 8, Laplace Appendix B: Spherically symmetric solutions and Birkhof’s theorem References Notation Index page 348 348 359 365 369 373 381 385 Preface The subject of this book is the structure of space-time on length- scales from 10-1 om, the radius of an elementary particle, up to 10%cm, the radius of the universe. For reasons explained in chapters 1 and 3, we base our treatment on Einstein’s General Theory of Relativity. This theory leads to two remarkable pre- dictions about the universe: first, that the final fate of massive stars is to collapse behind an event horizon to form a ‘black hole’ which will contain a singularity; and secondly, that there is a singularity in our past which constitutes, in some sense, a begin- ning to the universe. Our discussion is principally aimed at developing these two results. They depend primarily on two areas of study: first, the theory of the behaviour of families of timelike and null curves in space-time, and secondly, the study of the nature of the various causal relations in any space-time. We consider these subjects in detail. In addition we develop the theory of the time-development of solutions of Einstein's equations from given initial data. The dis- cussion is supplemented by an examination of global properties of a variety of exact solutions of Hinstein’s field equations, many of which show some rather unexpected behaviour. This book is based in part on an Adams Prize Essay by one of us (S. W. H.). Many of the ideas presented here are due to R. Penrose and R. P. Geroch, and we thank them for their help. We would refer our readers to their review articles in the Battelle Rencontres (Penrose (1968)), Midwest Relativity Conference Report (Geroch (1970c)), Varenna Summer School Proceedings (Geroch (1971), and Pittsburgh Conference Report (Penrose (19726)). We have benefited from dis- cussions and suggestions from many of our colleagues, particularly B. Carter and D. W, Sciama. Our thanks are due to them also. Cambridge 8. W. Hawking January 1973 G.F. R. Ellis Cxil 1 The role of gravity The view of physics that is most generally accepted at the moment is that one can divide the discussion of the universe into two parts. First, there is the question of the local laws satisfied by the various physical fields. These are usually expressed in the form of differential equations. Secondly, there is the problem of the boundary conditions for these equations, and the global nature of their solutions. This involves thinking about the edge of space-time in some sense. These two parts may not be independent. Indeed it has been held that the local laws are determined by the large scale structure of the universe. This view is generally connected with the name of Mach, and has more recently” been developed by Dirac (1938), Sciama (1953), Dicke (1964), Hoyle and Narlikar (1964), and others. We shall adopt a less ambitious approach: we shalll take the local physical laws that have been experi- mentally determined, and shall see what these laws imply about the large scale structure of the universe. There is of course a large extrapolation in the assumption that the physical laws one determines in the laboratory should apply at other points of space-time where conditions may be very different. If they failed to hold we should take the view that there was some other physical field which entered into the local physical laws but whose existence had not yet been detected in our experiments, because it varies very little over a region such as the solar system. In fact most of our results will be independent of the detailed nature of the physical laws, but will merely involve certain general properties such as the description of space-time by a pseudo-Riemannian geometry and the positive definiteness of enecgy density. The fundamental interactions at present known to physics can be divided into four classes: the strong and weak nuclear interactions, electromagnetism, and gravity. Of these, gravity is by far the weakest (the ratio @m*/e? of the gravitational to electric force between two electrons is about 10-*). Nevertheless it plays the dominant role in shaping the large scale structure of the universe. This is because the 2 THE ROLE OF GRAVITY strong and weak interactions have a very short range (~ 10-"em or less), and although electromagnetism is a long range interaction, the repulsion of like charges is very nearly balanced, for bodies of macro- scopic dimensions, by the attraction of opposite charges. Gravity on the other hand appears to be always attractive. Thus the gravitational fields of all the particles in a body add up to produce a field which, for sufficiently large bodies, dominates over all other forces. Not only is gravity the dominant force on a large scale, but it is a force which affects every particle in the same way. This universality was first recognized by Galileo, who found that any two bodies fell with the same velocity. This has been verified to very high precision in more recent experiments by Eotvos, and by Dicke and his collabo- rators (Dicke (1964). It has also been observed that light is deflected by gravitational fields. Since it is thought that no signals can travel faster than light, this means that gravity determines the causal structure of the universe, i.c. it determines which events of space-time can be causally related to each other. ‘These properties ‘of gravity lead to severe problems, for if a suffi- ciently large amount of matter were concentrated in some region, it could defiect light going out from the region so much that it was in fact dragged back inwards. This was recognized in 1798 by Laplace, who pointed out that a body of about the same density as the sun but 250 times its radius would exert such a strong gravitational field that no light could escape from its surface. That this should have been predicted so early is so striking that we give a translation of Laplace’s essay in an appendix. One can express the dragging back of light by a massive body more precisely using Penrose’s idea of a closed trapped surface. Consider a sphere J surrounding the body. At some instant let 7 emit a flash of light. At some later time f, the ingoing and outgoing wave fronts from J will form spheres J, and J; respectively. In a normal situa- tion, the area of 7, will be less than that of J (because it répresents ingoing light) and the area of J, will be greater than that of 7 (because it represents outgoing light; see figure 1). However if a suffi- ciently large amount of matter is enclosed within 7, the areas of 7, and J; will both be less than that of 7. The surface7 is then said to be a closed trapped surface. As ¢ increases, the area of 7, will get smaller and smaller provided that gravity remains attractive, i.e. pro- vided that the energy density of the matter does not become negative. Since the matter inside J cannot travel faster than light, it will be THE ROLE OF GRAVITY 3 trapped within a region whose boundary decreases to zero within a finite time. This suggests that something goes badly wrong. We shall in fact show that in such a situation a space-time singularity must occur, if certain reasonable conditions hold. One can think of a singularity as a place where our present laws of physics break down. Alternatively, one can think of it as representing part of the edge of space-time, but a part which is at a finite distance instead of at infinity. On this view, singularities are not so bad, but one still has the problem of the boundary conditions. In other words, one does not know what will come out of the singularity. S)) Fiune 1. At some instant, the sphere J emits a flash of light. At a later time, the light from a point p forms a sphere # around p, and the envelopes 7 and J form the ingoing and outgoing wavefronts respectively. If the areas of both ‘TF; and F,, are less than the area of , then F is a closed trapped surface. g There are two situations in which we expect there to be a sufficient concentration of matter to cause a closed trapped surface. The first is in the gravitational collapse of stars of more than twice the mass of the sun, which is predicted to occur when they have’exhausted their nuclear fuel. In this situation, we expect the star to collapse toa singu- larity which is not visible to outside observers. The second situation is that of the whole universe itself. Recent observations of the microwave background indicate that the universe contains enough matter to cause a time-reversed closed trapped surface. This implies the exist- ence of a singularity in the past, at the beginning of the present epoch of expansion of the universe. This singularity is in principle visible to us, It might be interpreted as the beginning of the universe. 4 THE ROLE OF GRAVITY In this book we shalll study the large scale structure of space-time on the basis of Einstein's General Theory of Relativity. The predic- tions of this theory are in agreement with all the experiments so far performed. However our treatment will be sufficiently general to cover modifications of Einstein's theory such as the Brans-Dicke theory. While we expect that most of our readers will have some acquain- tance with General Relativity, we have endeavoured to write this book so that it is self-contained apart from requiring a knowledge of simple calculus, algebra and point set topology. We have therefore devoted chapter 2 to differential geometry. Our treatment is reason- ably modern in that we have formulated our definitions in a manifestly coordinate independent manner. However for computational con- ‘venience we do use indices at times, and we have for the most part avoided the use of fibre bundles. The reader with some knowledge of differential geometry may wish to skip this chapter. In chapter 3 a formulation of the General Theory of Relativity is given in terms of three postulates about a mathematical model for space-time. This model is a manifold @ with a metric g of Lorentz signature. The physical significance of the metric is given by the first ‘two postulates: those of local causality and of local conservation of energy-momentum. These postulates are common to both the General and the Special Theories of Relativity, and so aro supported by the experimental evidence for the latter theory. The third postulate, the field equations for the metric g, is less well experimentally established. ‘However most of our results will depend only on the property of the field equations that gravity is attractive for positive matter densities. This property is common to General Relativity and some modifications, such as the Brans-Dicke theory. In chapter 4, we discuss the significance of curvature by considering its effects on families of timelike and null geodesics. These represent the paths of small particles and of light rays respectively. The curva- ture can be interpreted as a differential or tidal force which induces telative accelerations between neighbouring geodesics. If the energy momentum tensor satisfies certain positive definite conditions, this differential force always has a net converging effect on non-rotating families of geodesics. One can show by use of Raychaudhuri’s equation (4.26) that this then leads to focal or conjugate points where neigh- bouring geodesios intersect. To see the significance of these focal points, consider a one-dimen- sional surface Y in two-dimensional Euctidean space (figure 2). Let p THE ROLE OF GRAVITY 5 bea point not on. Then there will be some curve from to p which is shorter than, or as short as, any other curve from ¥ to p. Clearly this curve will be a geodesic, i.e. a straight line, and will intersect S orthogonally. In the situation shown in figure 2, there are in fact three geodesics orthogonal to. which pass through p. The geodesic through. the point r is clearly not the shortest curve from to p. One way of recognizing this (Milnor (1963)) is to notice that the neighbouring Ficure 2. The line pr cannot be the shortest line from p to Y, because there is @ focal point q between p and r. In fact either pr or py will be the shortest line from p to S. geodesics orthogonal to # through w and » intersect the geodesic through r at a focal point q between and p. Then joining the segment 1uq to the segment gp, one could obtain a curve from ¥ to p which had the same length as a straight line rp. However as ugp is not a straight line, one could round off the corner at q to obtain a curve from top which was shorter than rp. This shows that rp is not the shortest curve from # to p. In fact the shortest curve will be either xp oF yp. ‘One can carry these ideas over to the four-dimensional space-time manifold .@ with the Lorentz metric g. Instead of straight lines, one considers geodesics, and instead of considering the shortest curve one considers the longest timelike curve between a point p and a spacelike surface ¥ (because of the Lorentz signature of the metrie, there will be no shortest timelike curve but there may be a longest such curve). This longest curve must be a geodesic which intersects orthogonally, and there can be no focal point of geodesics orthogonal to ¥ between 6 THE ROLE OF GRAVITY SF and p. Similar results can be proved for null geodesics. These results are used in chapter 8 to establish the existence of singularities under certain conditions. In chapter 5 we describe a number of exact solutions of Einstein’s equations. These solutions are not realistic in that they all possess exact symmetries. However they provide useful examples for the suc- ceeding chapters and illustrate various possible behaviours. In particular, the highly symmetrical cosmological models nearly all possess space-time singularities. For a long time it was thought that these singularities might be simply a result of the high degree of symmetry, and would not be present in more realistic models. It will be one of our main objects to show that this is not the case. In chapter 6 we study the causal structure of space-time. In Special Relativity, the events that a given event can be causally affected by, or can causally affect, are the interiors of the past and future light cones respectively (see figure 3). However in General Relativity the metric ¢ which determines the light cones will in general vary from point to point, and the topology of the space-time manifold @ need not be that of Euclidean space 2, This allows many more possibilities. For instance one can identify corresponding points on the surfaces SF, and F, in figure 3, to produce a space-time with topology Rx. This would contain closed timelike curves. The existence of such a curve would lead to causality breakdowns in that one could travel into one’s past. We shall mostly consider only space-times which do not permit such causality violations. In such a space-time, given any spacelike surface Y, there is a maximal region of space-time (called the Cauchy development of “) which can be predicted from knowledge of data on ¥. A Cauchy development has a property (‘Global hyper- bolicity’) which implies that if two points in it can be joined by a time- like curve, then there exists a longest such curve between the points. ‘This curve will be a geodesic. ‘The causal structure of space-time can be used to define a boundary ‘or edge to space-time. This boundary represents both infinity and the part of the edge of space-time which is at a finite distance, i.e. the singular points. Tn chapter 7 we discuss the Cauchy problem for General Relativity. ‘We show that initial data on a spacelike surface determines a unique solution on the Cauchy development of the surface, and that in a certain sense this solution depends continuously on the initial data. ‘This chapter is included for completeness and because it uses a number THE ROLE OF GRAVITY 7 Time Space Past light cone Fioune 3. In Special Relativity, the light cone of an event p is the set of all light rays through p. The past of p is the interior of the past light cone, and the future of p is the interior of the future light cone. of results of the previous chapter. However it is not necessary to read. it in order to understand the following chapters. In chapter 8 we discuss the definition of space-time singularities. ‘This presents certain difficulties because one cannot regard thesingular points as being part of the space-time manifold . We then prove four theorems which establish the occurrence of space-time singularities under certain conditions. These conditions fall into three categories. First, there is the requirement that gravity shall be attractive. This can be expressed as an inequality on the energy-momentum tensor. Secondly, there is the requirement that there is enough matter present in some region to prevent anything escaping from that region. This will occur if there is a closed trapped 8 THE ROLE OF GRAVITY surface, or if the whole universe is itself spatially closed. The third requirement is that there should be no causality violations. However this requirement is not necessary in one of the theorems. The basic idea of the proofs is to use the results of chapter 6 to prove there must be longest timelike curves between certain pairs of points. One then shows that if there were no singularities, there would be focal points which would imply that there wereno longest curves between the pairs of points. Wenext describe a proceduresuggested by Schmidt for constructing a boundary to space-time which represents the singular points of space-time. This boundary may be different from that part of the causal boundary (defined in chapter 6) which represents singularities. In chapter 9, we show that the second condition of theorem 2 of chapter 8 should be satisfied near stars of more than 1} times the solar mass in the final stages of their evolution. The singularities which occur are probably hidden behind an event horizon, and so are not visible from outside. To an external observer, there appears to be a ‘black hole’ where the star once was. We discuss the properties of such black holes, and show that they probably settle down finally to one of the Kerr family of solutions. Assuming this to be the case, one can place certain upper bounds on the amount of energy which can be extracted from black holes. In chapter 10 we show that the second conditions of theorems 2 and 3 of chapter 8 should be satisfied, in a time-reversed sense, in the whole universe. In this case, the singularities are in our past and constitute a beginning for all or part of the observed universe. ‘The essential part of the introductory material is that in §3.1, §3.2 and § 3.4. A reader wishing to understand the theorems predicting the existence of singularities in the universe need read further only chap- ter 4, §6.2-§6.7, and §8.1 and §8.2. The application of these theorems to collapsing stars follows in §9.1 (which uses the results of appen- dix B); the application tothe universe as a whole is given in § 10.1, and relies on an understanding of the Robertson—Walker universe models ($5.8). Our discussion of the nature of the singularities is contained in §8.1, §8.3-§8.5, and § 10.2; the example of Taub-NUT space (§ 5.8) plays an important part in this discussion, and the Bianchi I universe model (§ 5.4) is also of some interest. A reader wishing to follow our discussion of black holes need read only chapter 4, §6.2-§6.6, §6.9, and §9.1, §9.2 and § 9.3. This discus- sion relies on an understanding of the Schwarzschild solution (§ 5.5) and of the Kerr solution (§ 5.6). THE ROLE OF GRAVITY 9 Finally a reader whose main interest is in the time evolution properties of Einstein’s equations need read only §6.2-§6.6 and chapter 7. He will find interesting examples given in §5.1, §5.2 and §5.5. We have endeavoured to make the index a useful guide to all the definitions introduced, and the relations between them. 2 Differential geometry The space-time structure discussed in the next chapter, and assumed through the rest of this book, is that of a manifold with a Lorentz metric and associated affine connection. In this chapter, we introduce in § 2.1 the concept of a manifold and in § 2.2 vectors and tensors, which are the natural geometric objects defined on the manifold. A discussion of maps of manifolds in §2.3 leads to the definitions of the induced maps of tensors, and of sub- manifolds. The derivative of the induced maps defined by a vector field gives the Lie derivative defined in §2.4; another differential operation which depends only on the manifold structure is exterior differentiation, also defined in that section. This operation occurs in the generalized form of Stokes’ theorem. An extra structure, the connection, is introduced in §2.5; this defines the covariant derivative and the curvature tensor. The connec- tion is related to the metric on the manifold in §2.6; the curvature tensor is decomposed into the Weyl tensor and Ricci tensor, which are related to each other by the Bianchi identities. In the rest of the chapter, a number of other topics in differential geometry are discussed. ‘The induced metric and connection on a hypersurface are discussed in §2.7, and the Gauss-Codacci relations are derived. The volume element defined by the metric is introduced in §2.8, and used to prove Gauss’ theorem. Finally, we give a brief discussion in § 2.9 of fibre bundles, with particular emphasis on the tangent bundle and the bundles of linear and orthonormal ‘frames. ‘These enable many of the concepts introduced earlier to be reformu- lated in an elegant geometrical way. §2.7 and §2.9 are used only at one or two points later, and are not essential to the main body of the book. [10] 2.1] MANIFOLDS it 2.1 Manifolds A manifold is essentially a space which is locally similar to Euclidean space in that it can be covered by coordinate patches. This structure permits differentiation to be defined, but does not distinguish intrin- sically between different coordinate systems. Thus the only concepts defined by the manifold structure are those which are independent of the choice of a coordinate system. We will give a precise formulation of the concept of a manifold, after some preliminary definitions. Let 2" denote the Huclidean pace of n dimensions, that is, the set of all n-tuples (21, x, ...,”") (—0o < 2 < 00) with the usual topology (open and closed sets are defined in the usual way), and let $2" denote the ‘lower half” of R*, i.e. the region of R" for which zt < 0. A map of an open set © < R” (respectively $2") to an open set 0’ < R™ (respec- tively 2”) is said to be of class CO" if the coordinates (2’?, x”?,...,’") of the image point ¢(p) in ©’ are r-times continuously differentiable functions (the rth derivatives exist and are continuous) of the co- ordinates (<1, 2%, ...,2") of p in 0. If a map is Cr for all > 0, then it is said to be C”. By a C° map, we mean a continuous map. A function f on an open set © of R is said to be locally Lipschitz if for each open set Y < @ with compact closure, there is some constant. K such that for each pair of points p,ge®, |f(p)—f(q)| < K |p—a|, where by |p| we mean {(2(p))? + (2*(p))? +... + en(p)) BF. A map ¢ will be said to be locally Lipschitz, denoted by C'-, if the coordinates of ¢(p) are locally Lipschitz functions of the coordinates of p. Similarly, we shall say that a map ¢ is Or- if it is C"* and if the (r—1)th derivatives of the coordinates of 4(p) are locally Lipschitz functions of the coordinates of p. In the following we shall usually only mention C*, but similar definitions and results hold for Or-. If #is an arbitrary set in R* (respectively $A"), a map ¢ from F to aset 2’ < R™ (respectively £R™) is said to be a Cr map if ¢ is the restriction to P and " of a Cr map from an open set © containing # to an open set 0’ containing ?”. A Cr n-dimensional manifold M is a set M together with a Cr atlas {%,, $,}, that is to say a collection of charts (%,,,) where the %, are subsets of W and the ¢, are one-one maps of the corresponding %, to open sets in R” such that (1) the %, cover M, ie. M= UUs, 12 DIFFERENTIAL GEOMETRY a (2) if &, fh % is non-empty, then the map PaO G5: Pp(Ua Up) > ba(Wa Nn Uy) is aC’ map of an open subset of R* to an open subset of R* (see figure 4). Each &, is a local coordinate neighbourhood with the local coordinates 2 (a = 1 ton) defined by the map ¢, (i.e. if pe %,, then the coordinates ofp are the coordinates of $,(p) in R*). Condition (2) is the requirement that in the overlap of two local coordinate neighbourhoods, the coordinates in one neighbourhood are Cr functions of the coordinates in the other neighbourhood, and vice versa. Bo tale 0%) iar ‘Froune 4. In the overlap of coordinate neighbourhoods &, and @%», coordinates are related by a C’ map $,0¢5-!. Another atlas is said to be compatible with a given C* atlas if their union is a Cr atlas for all @. The atlas consisting of all atlases com- patible with the given atlas is called the complete atlas of the manifold; the complete atlas is therefore the set of all possible coordinate systems covering .@. A The topology of M is defined by stating that the open sets of consist of unions of sets of the form %, belonging to the complete atlas. This topology makes each map ¢, into a homeomorphism. AO’ differentiable manifold with boundary is defined as above, on replacing ‘R*’ by ‘38’. Then the boundary of M, denoted by 0M, is defined to be the set of all points of @ whose image under a map ¢, lies on the boundary of $2” in R, 2.@isan (n— 1)-dimensional O* manifold without boundary. 2.4) MANIFOLDS 13 ‘These definitions may seem more complicated than necessary. How- ever simple examples show that one will in general need more than one coordinate neighbourhood to describe a space, The two-dimensional Euclidean plane R? is clearly a manifold. Rectangular coordinates (% y; —00<% <0, —2 0, 0 < @ < 27); one needs at least two such coordinate neighbourhoods to cover R?. The two-dimensional cylinder C? is the manifold obtained from R® by identi- fying the points (x,y) and (x+27,y). Then (z,y) are coordinates in a neighbourhood (0 < x < 2, —co 0,21 < 0, and one needs six such coordinate neighbourhoods to cover the surface. In fact, it is not possible to cover S? by @ single coordinate neighbourhood. The n-sphere S” can be similarly defined as the set of points (P+ (AEE. HME = 1 in Re, A manifold is said to be orientable if there is an atlas {%,, $,} in the complete atlas such that in every non-empty intersection %, 1 %p, the Jacobian |ax*/2x"4| is positive, where (z1,...,2") and (#,...,x'") are coordinates in %, and %, respectively. The Mbius strip is an example of a non-orientable manifold. ‘The definition of a manifold given so far is very general. For most purposes one will impose two further conditions, that .@ is Hausdorff and that is paracompact, which will ensure reasonable local behaviour. ‘A topological space -M is said to be a Hausdorff space if it satisfies the Hausdorff separation axiom: whenever p, q are two distinct points in M, there exist disjoint open sets %,¥ in M such that pe®, geV. One might think that a manifold is necessarily Hausdorff, but this is not $0. Consider, for example, the situation in figure 5. We identify the points 6, b’ on the two lines if and only if x = yy < 0. Then each point is contained in a (coordinate) neighbourhood homeomorphic to an open subset of Rt. However there are no disjoint open neighbourhoods 44 DIFFERENTIAL GEOMETRY [ea Froure 5. An example of a non-Hausdorff manifold. The two lines above are identical for x = y <0. However the two points a (z = 0) and a’(y = 0) are not identified. %,¥ satisfying the conditions ae %, a’ «¥”, where ais the point x = 0 and a’ is the point y = 0. An atlas {W,, g,} is said to be locally finite if every point pe M has an open neighbourhood which intersects only a finite number of the sets %,. M is said to be paracompact if for every atlas {%,,9,} there existsa locally finite atlas {¥7, ¥4} with each ¥% contained in some %,.. A connected Hausdorff manifold is paracompact if and only if it has a countable basis, i.e, there is a countable collection of open sets such that any open set can be expressed as the union of members of this collection (Kobayashi and Nomizu (1963), p. 271). Unless otherwise stated, all manifolds considered will be paracompact, connected C* Hausdorff manifolds without boundary. It will turn out later that when we have imposed some additional structure on (the existence of an affine connection, see § 2.4) the requirement of para- compactness will be automatically satisfied because of the other restrictions. A function f on a C* manifold @ is a map from M to R}. It is said to be of class Cr (r < k) at a point p of MH, if the expression fo $,-1 of f on any local coordinate neighbourhood %, is a Cr function of the local coordinates at p; and f is said to be a O' function on a set ¥ of M if fis a Or function at each point pey”. , A property of paracompact manifolds we will use later, is the fol- lowing: given any locally finite atlas {%,,¢,} on a paracompact C* manifold, one can always (see e.g. Kobayashi and Nomizu (1963), Pp. 272) find a set of C* functions g, such that (1) 0 for any one-forms w, and any a, fe Rt, Xe, one can regard {E*} as.a basis of one-forms since any one-form w at p can be expressed as = «, Et whore the numbers «, are defined by «, = Tes. ay DOM, a, Digg a Dg, Dy... Dyibe, (2.4) ‘The contraction of a tensor T of type (r,s), with components To-2,__4 with respect to bases {E,}, {E*}, on the first contravariant and first covariant indices is defined to be the tensor Cl(T) of type (r—1,8—1) whose components with respect to the same basis are Pebndy OUT) = Pet, Ey ®... @Ez@E@... @E’. 20 DIFFERENTIAL GEOMETRY [2.2 If (E,}, {E*} are another pair of dual bases, the contraction Ci(T) defined by them is OUT) = 128 yy, Ey @ - OBZ @E@... OE" =O, 4, THO Oo”, E,@... @E,@E’... OE = Pend, 7Ey®-. OEG@E@... OF = Ci(T), 20 the contraction C} of a tensor is independent of the basis used in its definition. Similarly, one could contract T over any pair of contra- variant and covariant indices. (If we were to contract over two contra- variant or covariant indices, the resultant tensor would depend on the basis used.) ‘The symmetric part of a tensor T of type (2,0) is the tensor S(T) defined by » Far. Pr S(T) oma) =F (Peon) +e} for all y;,72€7'*,. We shall denote the components S(T) of S(T) by 1 Te; then Ped = 5 (Tab 4 Tee}, Similarly, the components of the skew-symmetric part of T will be denoted by san = 3 (ree In general, the components of the symmetric or antisymmetric part of @ tensor on a given set of covariant or contravariant indices will be denoted by placing round or square brackets around the indices. Thus Teay..an) 2 {eum ovor all permutations of the indicos a, to @,(Ta,.:4,>~"} an! i = 4 {altemating sum over all permutations of the indices a, to 4, (Ty,...0,° Me For example, Kon = MK cat KYane+ Kea» K*oae— Kena — Kae) 2.2) VECTORS AND TENSORS 21 A tensor is symmetric in a given set of contravariant or covariant indices if it is equal to its symmetrized part on these indices, and is antisymmetric if it is equal to its antisymmetrized part. Thus, for example, a tensor T of type (0,2) is symmetric if Ty, = 4(Tay-+Tha)s (which we can also express in the form: Tas) = 0). A particularly important subset of tensors is the set of tensors of type (0,q) which are antisymmetric on all g positions (so q < nm); such @ tensor is called a q-form. If A and B are p- and q-forms respectively, ‘one can define a (p+q)-form A AB from them, where A is the skew- symmetrized tensor product ®; that is, A A B is the tensor of type (©, p+q) with components determined by (AAB)a....0...9 = Ata...0 Bo. This rule implies (A AB) = (—)"*(BA A). With this product, the space of forms (i.e. the space of all p-forms for all p, including one- forms and defining scalars as zero-forms) constitutes the Grassmann algebra of forms, If {E*} is a basis of one-forms, then the forms E% ).... \B% (a,run from 1 to) are a basis of p-forms, as any p-form Acan be written A = 4,_,E*A... \E?, where Ag..5 = Ata, .o+- So far, we have considered the set of tensors defined at a point on the manifold. A set of local coordinates {z‘} on an open set % in # defines @ basis {(2/éx*)|,} of vectors and a basis {(dz*)|,} of one-forms at each point p of , and so defines a basis of tensors of type (r, 8) at each point of &. Such a basis of tensors will be called a coordinate basis. A C* tensor field T of type (r, 8) on aset ¥ < .# is an assignment of an element of 77(p) to each point peY such that the components of T with respect to any coordinate basis defined on an open subset of ¥ are C functions. In general one need not use a coordinate basis of tensors, i.e. given any basis of vectors {E,} and dual basis of forms {E*} on ¥”, there will not necessarily exist any open set in ¥~ on which there are local coordinates {z*} such that E, = @/a2¢ and E* = dz*. However if one does use a coordinate basis, certain specializations will result; in parti- cular for any function f, the relations Z,(E,f) = E,(H,f) are satisfied, being equivalent to the relations %f/éx*dx> = d%/éx aa. If one changes from a coordinate basis E, = /éx* to a coordinate basis E, = 0/éx*, applying (2.2), (2.3) to x, 2 shows that a. ot Mn biccflm llich Clearly a general basis {E,} can be obtained from a coordinate basis 22 DIFFERENTIAL GEOMETRY [2.2 {@/éz"} by giving the functions H,¢ which are the components of the E, with respect to the basis {2/d2*}; then (2.2) takes the form E, = H,*2/dx* and (2.3) takes the form E* = E#,dz*, where the matrix E*, is dual to the matrix 'E,f. 2.3 Maps of manifolds In this section we define, via the general concept of a C* manifold map, the concepts of ‘imbedding’, ‘immersion’, and of associated tensor maps, the first two being useful later in the study of submanifolds, and the last playing an important role in studying the behaviour of families of curves as well as in studying symmetry properties of manifolds. Amap ¢ froma C* n-dimensional manifold. toa C”’ n’-dimensional manifold .#’ is said to be a C’ map (r < k, r < k’) if, for any local coordinate systems in .@ and .’, the coordinates of the image point ¢(p) in ’ are Ct functions of the coordinates of p in @. As the map will in general be many-one rather than one-one (e.g. it cannot be one-one if n > x’), it will in general not have an inverse; and if a Cr map does have an inverse, this inverse will in general not be Cr (e.g. if ¢ is the map R!—+ R? given by z->2°, then $-? is not differentiable at the point x = 0). Iff isa function on.@’, the mapping ¢ defines the function ¢*fon as the function whose value at the point p of . is the value of f at $(p), i $*Mp) = Hdl). (2.5) ‘Thus when ¢ maps points from @ to .4’, ¢* maps functions linearly from 4’ to #. Tf A() is a curve through the point pe.#, then the image curve G(A(t)) in A’ passes through the point ¢(p). If r > 1, the tangent vector to this curve at 4(p) will be denoted by $4(2/@),ly)3 one can regard it as the image, under the map ¢, of the vector (2/2),h,- Clearly . is a linear map of T,(M) into Ti’). From (2.5) and the defini- tion (2.1) of a vector as a directional derivative, the vector map $4 can be characterized by the relation: for each C* (r > 1) function f at ¢(p) and vector X at p, XP p = Fe XP) goo (2.6) Using the vector mapping $4 from to.@’, we can ifr > 1 define a linear one-form mapping ¢* from T*y,,(-4’) to 7*,(-@) by the condition: vector-one-form contractions are to be preserved under the 2.3] MAPS OF MANIFOLDS 23 maps. Then the one-form AcT'*y, is mapped into the one-form $*AeT"™, where, for arbitrary vectors XeT,,, KP*A, Xp = (A be XI yoo A consequence of this is that Pr (Af) = A(p*f). (2.7) The maps ¢y and $* can be extended to maps of contravariant tensors from - to.” and covariant tensors from .4” to -M respeo- tively, by the rules $4: Te73(p)->¢4T€T5(¢(p)) where for any ee and $*: Te TUS(p)) > *T eTXp), where for any X;¢T7,,, BAT (Ky +) Xp = Tha Xa 1 ba Xs) oto When r > 1, the Or map ¢ from # to .#’ is said to be of rank s at p if the dimension of $.4(T,(-@)) is 8. It is said to be injective at p if. (and so n < n’) at p; then no vector in Z, is mapped to zero by $y. It is said to be surjective if s = n’ (son > n’). A Cr map ¢ (7 > 0) is said to be an immersion if it and its inverse are C’ maps, ie. if for each point pe. there is a neighbourhood & of p in @ such that the inverse g- restricted to (%) is also a Cr map. This implies n < n’. By the implicit function theorem (Spivak (1965), p. 41), when r > 1, ¢ will be an immersion if and only if it is injective at every point pe.#; then ¢y is an isomorphism of 7, into the image $4(Z,) S Typ The image $(-@) is then said to be an n-dimensional immersed submanifold in €’. This submanifold may intersect itself, i.e. ¢ may not be a one-one map from .# to ¢(#) although it is one-one when restricted toa sufficiently small neighbour- hood of @. An immersion is said to be an imbedding if it js a homeo- morphism onto its image in the induced topology. Thus an imbedding is a one-one immersion; however not all one-one immersions are imbeddings, of. figure 6. A map ¢ is said to be a proper map if the inverse image (2) of any compact set # < .#” is compact. It can be shown that a proper one-one immersion is an imbedding. The image $(@) of @ under an imbedding ¢ is said to be an n-dimensional imbedded submanifold of -€’. ‘The map ¢ from to A” is said to be a O* diffeomorphism if it is a one-one OF map and the inverse ¢-! is a C’ map from .#’ to .#. In 24 DIFFERENTIAL GEOMETRY (2.3 this case, n = n', and ¢ is both injective and surjective if r > 1; con- versely, the implicit function theorem shows that if ¢, is both injective and surjective at p, then there is an open neighbourhood % of p such that 4: & + $(%) isa diffeomorphism. Thus ¢ isa local diffeomorphism near p if $y is an isomorphism from 7, to Ty,)- Ficune 6. A one-one immersion of R? in R# which is not an imbedding, obtained by joining smoothly part of the curve y = sin (1/z) to the curve {(y.0); -o 1) diffeomorphism, $4 maps T,(-) to Tyg) and (6-)* maps T*,(€) to T* y,,(#). Thus we can define amap $4 of Ty(p) to Ti(9(p)) for any r, 6, by Ten, nt Xp XNlp ae PUSAN, (GM, ba Xs be XN gor for any X,¢7,, n'€1'*,. This map of tensors of type (r, 8) gn to tensors of type (r,s) on -€’ preserves symmetries and relations in the tensor algebra; e.g. the contraction of $4 T is equal to ¢q (the con- traction of T). 2.4 Exterior differentiation and the Lie derivative We shall study three differential operators on manifolds, the first two being defined purely by the manifold structure while the third is defined (see § 2.5) by placing extra structure on the manifold. 2.4] EXTERIOR DIFFERENTIATION 25 ‘The exterior differentiation operator d maps r-form fields linearly to (r+1)-form fields. Acting on a zero-form field (i.e. a function) f, it gives the one-form field df defined by (cf. §2.2) AM’ isa Cr (r > 2) map and A is a C* (k > 2) form field on 4’, then (by (2.7)) nl (which is equivalent to the chain rule for partial derivatives). ‘The operator d occurs naturally in the general form of Stokes’ theorem on a manifold. We first define integration of n-forms: let @ be a compact, orientable n-dimensional manifold with boundary 24 and let {f,} be @ partition of unity for a finite oriented atlas {%,, $2}. ‘Then if A is an n-form field on, the integral of A over is defined as [A= Ore fede ndatdat der, 2.10) Ue 2 Jeo 2 where Ajp,., are the components of A with respect to the local co> ordinates in the coordinate neighbourhood %,, and the integrals on the right-hand side are ordinary multiple integrals over open sets $.(¥,) of R*. Thus integration of forms on -# is defined by mapping the form, by local coordinates, into R* and performing standard multiple integrals there, the existence of the partition of unity ensuring the global validity of this operation. ‘The integral (2.10) is well-defined, since if one chose another atlas {% Wp} and partition of unity {94} for this atlas, one would obtain the integral (Ef ae Aae ow Ga dat... da, Byers) where x* are the corresponding local coordinates. Comparing these two quantities in the overlap (%, 0 ¥g) of coordinate neighbourhoods belonging to two atlases, the first expression can be written yy A. OO EE Facearn tr dat dx? ,..dz", and the second can be written OBE Papasan teen a Oa oda Comparing the transformation laws for the form A and the multiple integrals in R*, these expressions are equal at each point, 80 ii ahi independent of the atlas and partition of unity chosen. 2.4] EXTERIOR DIFFERENTIATION a7 Similarly, one can show that this integral is invariant under diffeomorphisms: f aA = fi A ae a if g is a Cr diffeomorphism (r > 1) from to’. Using the operator d, the generalized Stokes’ theorem can now be written in the form: if B is an (n—1)-form field on #, then fe faam which can be verified (see e.g. Spivak (1965)) from the definitions above; it is essentially a general form of the fundamental theorem of calculus. To perform the integral on the left, one has to define an orientation on the boundary @.# of #. This is done as follows: if %, is « coordinate neighbourhood from the oriented atlas of .@ such that u, intersects @#, then from the definition of 2#, $,(%, aH) lies in the plane z1 = 0 in R* and ¢,(%, 1 #) lies in the lower half 21 < 0. The coordinates (2*, 8, ...,2") are then oriented coordinates in the neighbourhood %, n @.# of 24. It may be verified that this gives an oriented atlas on 2.4. The other type of differentiation defined naturally by the manifold structure is Lie differentiation. Consider any Cr (r > 1) vector field K on.#. By the fundamental theorem for systems: of ordinary differential equations (Burkill (1956)) there is a unique maximal curve A(t) through each point p of @ such that A(0) = p and whose tangent vector at the point A(t) is the vector X|,¢9. If {x*} are local coordinates, so that the curve A(t) has coordinates 2*(t) and the vector X has components X‘, then this curve is locally a solution of the set of differential equations dat/dt = X(xM(t), ...,20(¢)). This curve is called the integral curve of X with initial point p. For each point q of @, there is an open neighbourhood % of g and ane > Osuch that X defines a family of diffeomorphisms ¢,: @+- whenever |t| < , obtained by. taking each point p in Y a parameter distance ¢ along the integral curves of X (in fact, the ¢, form a one-parameter local group of diffeomorphisms, as ¢,,, = ¢,0¢, = $,0¢, for MH, ls], [+e] <¢, so gy =()- and $5 is the identity). ‘This diffeomorphism maps each tensor field T at p of type (r,s) into $e Th ogo. The Lie derivative LyT of a tensor field T with respect to X is 28 DIFFERENTIAL GEOMETRY [2.4 defined to be minus the derivative with respect to ¢ of this family of tensor fields, evaluated at t = 0, i.e. IxTly im F (Tp be Fly} a From the properties of $4, it follows that (1) Ly preserves tensor type, i.e. if T is a tensor field of type (r,8), then Lz'T is also a tensor field of type (7,8); (2) Lx maps tensors linearly and preserves contractions. As in ordinary calculus, one can prove Leibniz’ rule: (3) For arbitrary tensors S, T, Ix(S @ T) = LzS @T+S @LxT- Direct from the definitions: (4) Lxf = Xf, where f is any function. Under the map ¢,, the point q = ¢_(p) is mapped into p. Therefore ¢u is a map from T, to T,,. Thus, by (2.6), ae Pf le = VOM If {x} are local coordinates in a neighbourhood of p, the coordinate components of $j Y¥ at p are (ba Ply bee Vy! = Vaz OD) ~ SED) yi, Now BHD xg therefore $(SS2) a | itt nV =-S Gu Moe Be By ean One can rewrite this in the form (Lx ¥)f = X(¥f)- YXS) for all C? functions f. We shall sometimes denote Z¥ by [X,Y], i.e. LY = — LyX = (X,¥] = -(¥,X1. If the Lie derivative of two vector fields X, ¥ vanishes, the vector fields are said to commute. In this case, if one starts at a point p, goes a parameter distance ¢ along the integral curves of X and then a parameter distance ¢ along the integral curves of Y, one arrives at the 2.4) THE LIE DERIVATIVE 29 same point as if one first went a distance ¢ along the integral curves of Y and then a parameter distance ¢ along the integral curves of X (see figure 7). Thus the set of all points which can be reached along integral curves of X and ¥ from a given point p will then form an immersed two-dimensional submanifold through p. sexioextv)) xeav(P)) Frovne 7. Tho tranaformations generated by commuting veetor fields X, ¥ movea point p to points a(P), Gax(P) Feapectively. By successive opplications of these transformations, p is moved to the pointe of a two-urface. The components of the Lie derivative of a one-form w may be found by contracting the relation Lx(w @ Y) = Lyw @Y +w @LzY (Lie derivative property (3)) to obtain x6, ¥) = (Lge, ¥)+ (eo, Dx¥) (by property (2) of; Lie derivatives), where X, Y are arbitrary (1 vector fields, and then choosing Y as a basis vector E,. One finds the coordinate components (on choosing E, = @/az‘) to be (Lrxeo); = (B04]02!) X5 + (0X4) because (2.11) implies (Zex(0/2x*))) = — Xi font. Similarly, one can find the components of the Lie derivative of any Cr (r > 1) tensor field T of type (r, 8) by using Leibniz’ rule on Ix(T@E*®...@E4@E,@...@E,), 30 | DIFFERENTIAL GEOMETRY [24 and then contracting on all positions. One finds the coordinate com- ponents to be (ExT) hg g = (OPM Ay fox) Xt — Diddy OX a/aat — (all upper indices) +744, @X¢/ax* + (all lower indices). (2.12) Because of (2.7), any Lie derivative commutes with d, i.e. for any p-form field w, Se From these formulae, as well as from the geometrical interpretation, it follows that the Lie derivative ZT], of a tensor field T of type (r,8) depends not only on the direction of the vector field X at the point p, but also on the direction of X at neighbouring points. Thus the two differential operators defined by the manifold structure are too limited to serve as the generalization of the concept of partial derivative one needs in order to set up field equations for physical quantities on the manifold; d operates only on forms, while the ordinary partial derivative is a directional derivative depending only on a direction at the point in question, unlike the Lie derivative. One obtains such a generalized derivative, the covariant derivative, by introducing extra structure on the manifold. We do this in the next section. 2.5 Covariant differentiation and the curvature tensor The extra structure we introduce is a (affine) connection on . A connection V at a point p of @ is a rule which assigns to each vector field X at p a differential operator Vx which maps an arbitrary Cr (r > 1) vector field ¥ into a vector field Vz¥, where: (1) Vx¥ is a tensor in the argument X, i.e. for any functions f, 9, and Ct vector fields X, ¥, Z, VexrorZ =fVxZ+9VyZ; (this is equivalent to the requirement that the derivative Vx at p depends only on the direction of X at p); (2) Vx¥ is linear in Y, i.e. for any C1 vector fields ¥, Zanda, fe Ri, Vx(@¥ +£Z) = aVg¥ +PVxZ; (3) for any C1 function f and C1 vector field Y, Val f¥) = X()¥+fVx¥. 2.5) COVARIANT DIFFERENTIATION 31 Then VxY¥ is the covariant derivative (with respect to V) of ¥ in the direction X at p. By (1), we can define VY, the covariant derivative of Y, as that tensor field of type (1,1) which, when contracted with X, produces the vector VY. Then we have (3)<+ VY) =df@Y+fVY. A Cr connection V on a C* manifold .@ (k > r+2) is a rule which assigns a connection V to.each point such that if ¥ is a Cr+ vector field on .#, then VY is a C* tensor field, Given any C™ vector basis {E,} and dual one-form basis {E*} on neighbourhood ®%, we shall write the components of VY as ¥*,,, 50 VY = ¥*,, B@E,. ‘The connection is determined on & by n* Cr functions I,, defined by 4, = (E*, Va,E.) <> VE, = 14,,E’@E,. For any C1 vector field ¥, VY = V(YE,) = d¥°@E,+ ¥T*,.B@E,. Thus the components of VY with respect to coordinate bases {2/24}, {dz?} are ¥e,, = a¥e/au>+14,, Ye, ‘The transformation properties of the functions I, are determined by connection properties (1), (2), (3); for Tye = (E*, VayEc) = (0%, E*, Voyon,(Po"E.)> = 08, Dy(By( D9) + Doe Pye) ifE, = ©,“ E,, E€ = ©, E*. One can rewrite this as Py ¢ = 0% (By (o4) + OY OT 4,,). In particular, if the bases are coordinate bases defined by coordinates {z*}, {2}, the transformation law is Because of the term E,,(,), the I, do not transform as the compo- nents of a tensor. However if VY and VY are covariant derivatives obtained from two different connections, then vy — Oy = (14,,— f,,) YeE*@E, will be a tensor. Thus the difference terms (I4,— components of a tensor. 32 DIFFERENTIAL GEOMETRY [2.5 ‘The definition of a covariant derivative can be extended to any Cr tensor field ifr > 1 by the rules (cf. the Lie derivative rules): (1) if Tis. Cr tensor field of type (q, ), then VT is a C"— tensor field of type (g, +1); (2) Vis linear and commutes with contractions; (8) for arbitrary tensor fields S, T, Liebniz’ rule holds, i.e. V(S@T) = VS@T+S@VT; (4) Vf = df for any function f. ‘We write the components of VT as (Va,7)*~4...¢ = T@%, o:n- AS a consequence of (2) and (3), Vz, Ee = —T4,E*, where {E*} is the dual basis to {E,}, and methods similar to those used in deriving (2.12) show that the coordinate components of VT are Teekay gan = ODM A gy, gf B* + Vey Pirie + (all upper indices) —I¥,_T%-4, As o particular example, the unit tensor E,@E*, which has compo- nents 6%, has vanishing covariant derivative, and so the generalized unit tensors with components 6%,,4%,...3%,, diy, 0%, ... 3%, (p 0) connection, the standard existence theorems for ordinary differential equations applied to (2.15) show that for any point p of @ and any vector X, at p, there exists a maximal geodesic Ax(v) in -@ with starting point p and initial direction X,, i.e. such that Ax(0) = p and (0/0),|,-9 = Xp. Ifr > 1 —, this geodesic is unique and depends continuously on p and X,. If r > 1, it depends differentiably onp and X,. This means that if r > 1, one can define a C’ map exp: 1, > M, where for each X€T,, exp (X) is the point in a unit para- meter distance along the geodesic Ax from p. This map may not be defined for all Xe 7, since the geodesic Ax(v) may not be defined for all v. If » does take all values, the geodesic A(v) will be said to be a complete geodesic. The manifold . is said to be geodesically complete ifall geodesics on. are complete, that is if exp is defined on all 7,, for every point p of @. Whether .# is complete or not, the map exp, is ofrank n at p. There- fore by the implicit function theorem (Spivak (1965)) there exists an 34 DIFFERENTIAL GEOMETRY (25 open neighbourhood % of the origin in 7, and an open neighbourhood JM, of p in M such that the map exp is a Or diffeomorphism of Wg onto 4. Such a neighbourhood %, is called a normal neighbourhood ofp. Further, one can choose 4, to be conver, i.e. to be such that any point g of %, can be joined to any other point r in 7, by a unique geodesic starting at g and totally contained in #,. Within a convex normal neighbourhood W one can define coordinates (z',...,2") by choosing any point q¢/’, choosing a basis {E} of 7,, and defining the coordinates of the point rin / by the relation r xp (z*E,) (ie. one assigns to r the coordinates, with respect to the basis {E,}, of the point exp"!(r) in Z,.) Then (2/2x*)|, = E, and (by (2.15)) Myiolg = 0. Such coordinates will be called normal coordinates based on q. The existence of normal neighbourhoods has been used by Geroch (1968¢) to prove that a connected C* Hausdorff manifold .@ with a C? connection has a countable basis. Thus one may infer the property of paracompactness of a C® manifold from the existence of aC? connection on the manifold. ‘The ‘normal’ local bohaviour of geodesics in these neighbourhoods is, in contrast to the behaviour of geodesics in the large in a general space, where on the one hand two arbitrary points cannot in general be joined by any geodesic, and on the other hand some of the geodesics through one point may converge to ‘focus’ at some other point. We shall later encounter examples of both types of behaviour. Given a Ct connection V, one can define a Cr-? tensor field T of type (1, 2) by the relation T(X%,¥) = Vx¥—VyX~[X, ¥], where X, ¥ are arbitrary Cr vector fields. This tensor is called the torsion tensor. Using a coordinate basis, its components are Bye = Thy Thy ‘We shall deal only with torsion-free connections, i.e. we shall assume ‘'T = 0. In this case, the coordinate components of the connection obey, Ty, = P44, 0 such a connection is often called a symmetric connec- tion. A connection is torsion-free if and only if f,¢ = f,« for all fune- tions f. From the geodesic equation (2.15) it follows that a torsion-free connection is completely determined by a knowledge of the geodesics on. ‘When the torsion vanishes, the covariant derivatives of arbitrary O* vector fields X, ¥ are related to their Lie derivative by (X,¥] = Va¥— VX (Lx¥)* = ¥4,,X°—Xe,,¥%, (2.16) 2.5] THE CURVATURE TENSOR 35 and for any C? tensor field T of type (r,s) one finds eS — (all upper indices) +74, X4,¢+ (all lower indices). (2.17) One can also easily verify that the exterior derivative is related to the covariant derivative by dA = Ag, o,ade? A dat An... Ada? <> (dA)a...ca = (—)?Ata. where A is any p-form. Thus equations involving the exterior deriva- tive or Lie derivative can always be expressed in terms of the co- variant derivative. However, because of their definitions, the Lie derivative and exterior derivative are independent of the connection. If one starts from a given point p and parallelly transfers a vector X, along a curve y that ends at p again, one will obtain a vector X’, which is in general different from X,; if one chooses a different curve 7/, the new vector one obtains at p will in general be different from X, and X’,. This non-integrability of parallel transfer corresponds to the fact that the covariant derivatives do not generally commute. The Riemann (curvature) tensor gives a measure of this non-commutation. Given C"+1 vector fields X, Y, Z, a C’-1 vector field R(X, ¥) Z is defined by a Cr connection V by R(X, ¥)Z = Va(VyZ)— Ve(VxZ)—VizniZ- (2.18) ‘Then R(X, Y)Z is linear in X,¥,Z and it may be verified that the value of R(X, Y)Z at p depends only on the values of X, Y, Z at p, ie, it is a Cr tensor field of type (3, 1). To write (2.18) in component form, we define the second covariant derivative VVZ of the vector Z as the covariant derivative V(VZ) of VZ; it has components Zag = (2 Wier ‘Then (2.18) can be written Req X°VUDP om (Ze g ¥4), eX? — (ZA, gX4), p¥° ; = 22, o(¥4,.X°—X4,.¥°) e:ay = (2, ag— 2 ,0a) X°Y4, where the Riemann tensor components R%,.g with respect to dual bases {E,}, {E*} are defined by R%oq = (E4, R(E,, Ez) Ey). As X, ¥ are arbitrary vectors, / ga, — Ze, 4g = Re, ZP (2.19) expresses the non-commutation of second covariant derivatives of Z in terms of the Riemann tensor. 36 DIFFERENTIAL GEOMETRY 2 Since Vx @VzZ) = Van@VyZ+n@VxVzZ = (0, VzVeZ)> = X() — (Van, Ve Z> holds for any C# one-form field and vector fields X, Y, Z, (2.1 implies — (E%, Vinn Ep). Choosing the bases as coordinate bases, one finds the expression Bjoq = OT ay /2i! — OP Ag /224 + Pty May —THgy Mey (2.20 for the coordinate components of the Riemann tensor, in terms of th coordinate components of the connection. It can be verified from these definitions that in addition to th symmetry Bo jeg = — R%ygq 2 Ryo, = 0 (2.21a the curvature tensor has the symmetry Begg, = 0-<> Bpeg + Roane + Recap = 0. (2.210 Similarly the first covariant derivatives of the Riemann tensor satisfi Bianchi’s identities Reegea;e) = 0 Brea; e+ Ree, a+ racy = 0- (2.22 It now turns out that parallel transfer of an arbitrary vector alon an arbitrary closed curve is locally integrable (i.e. X’, is necessarily th: same as X, for each pe.#) only if R%,q = at all points of .; in thi case wo say that the connection is flat By contracting the curvature tensor, one can define the Ricei tenso: as the tensor of type (0, 2) with components Roa = Ryaa: 2.6 The metric A metric tensor & at a point p<. is a symmetric tensor of type (0, 2 at p, so a Cr metric on is a C’ symmetric tensor field g. The metric g at p assigns a ‘magnitude’ (|g(X,X)|)# to each vector Xe, ané defines the ‘cos angle’ (X,Y) (9X, X)-9(¥, YF between any vectors X, Ye, such that g(X, X).g(¥, ¥) + 0; vectors X,Y will be said to be orthogonal if g(X, Y) = 0. 2.6] THE METRIC 37 The components of g with respect to a basis {E,} are 9(Eq, Ey) = 9(Ex, Ea), i.e. the components are simply the scalar products of the basis vectors E,. Ifa coordinate basis {2/224} is used, then 8 = Jap Ax*@dz?. (2.23) ‘Tangent space magnitudes defined by the metric are related to magnitudes on the manifold by the definition: the path length between points p = (a) and q = y(b) along a 0°, piecewise C? curve y(t) with tangent vector 2/ét such that g(@/2t, 2/2) has the same sign at all points along y(t), is the quantity Lm J (iatefee,of00))kat (2.24) Gao We may symbolically express the relations (2.23), (2.24) in the form ds? = gy det dat used in classical textbooks to represent the length of the ‘infinitesimal’ are determined by the coordinate displacement a*—> x‘ + dz‘. The metric is said to be non-degenerate at p if there is no non-zero vector Xe7!, such that g(X, Y) = 0 for all vectors Ye7,. In terms of components, the metric is non-degenerate if the matrix (g,,) of com- ponents of g is non-singular. We shall from now on always assume the metric tensor is non-degenerate. Then we can define a unique sym- metric tensor of type (2, 0) with components g® with respect to the basis {E,} dual to the basis {E°}, by the relations 90 = Fes ie. the matrix (g%) of components is the inverse of the matrix (g4,)- It follows that the matrix (g*) is also non-singular, so the tensors 9g, Jan can be used to give an isomorphism between any covariant tensor argument and any contravariant argument, or to ‘raise and lower indices’. Thus, if X* are the components of a contravariant vector, then X, are the components of a uniquely associated covariant vector, where X, = gq)X*, X* = g%X,; similarly, to a tensor 7, of type (0, 2) wo can associnte|unique tensors 7%, = 9° Tey, Te? = Tacs Ta = gog@T.,. We shall in’ general regard such associated covariant and contravariant tensors as representations of the same geometric object (so in particular, g,,, 6,2and g® may be thought of as representa- tions (with respect to dual bases) of the same geometric object g), 38 DIFFERENTIAL GEOMETRY [2.6 although in some cases where we have more than one metric we shall have to distinguish carefully which metric is used to raise or lower indices. The signature of g at p is the number of positive eigenvalues of the matrix (9,5) at p, minus the number of negative ones. If ¢ is non- degenerate and continuous, the signature will be constant on W; by suitable choice of the basis {E,}, the metric components can at any point p be brought to the form Gon = ding (+1, +1, ..., 44, =), H(n+s)terms 4(n—s) terms where ¢ is the signature of g and n is the dimension of . In this case the basis vectors {E,} form an orthonormal set at p, i.e. each is a unit vector orthogonal to every other basis vector. A metric whose signature is n is called a positive definite metric; for such a metric, g(X,X) = 0 + X = 0, and the canonical form is Gan = diag (+1, ..., +1). Satie thie ‘terme A positive definite metric is a ‘metric’ on the space, in the topological sense of the word. A metric whose signature is (x — 2) is called a Lorentz metric; the canonical form is Gao = diag (+1,...,+1,—1). (n=1)terms With a Lorentz metric on, the non-zero vectors at p can be divided into three classes: a vector Xe 7, being said to be timelike, null, or spacelike according to whether g(X, X) is negative, zero, or positive, respectively. The null vectors form a double cone in T, which separates the timelike from the spacelike vectors (see figure 8). If X, Y are any two non-spacelike (i.e. timelike or null) vectors in the same half of the light cone at p, then g(X, Y) < 0, and equality can only hold if X and ¥ are parallel null vectors (i.e. if X = a¥, g(X, X) = 0) Any paracompact Cr manifold admits a C’— positive definite metric (that is, one defined on the whole of .#). To see this, let {f,} be a parti- tion of unity for a locally finite atts {%,,, ¢,}. Then one can define g by (X,Y) = DIG.) ($2) Y), where ¢ , ) is the natural scalar product in Euclidean space R*; thus one uses the atlas to determine the metric by mapping the 2.6] THE METRIC 39 En Null cone Null vectors lie _7 on the null cones ‘Timelike vectors lie inside the null cones: Spacelike veetors lie Fiourn 8. The null cones defined by a Lorentz metric. Euclidean metric into. This is clearly not invariant under change of atlas, so there are many such positive definite metrics on #. In contrast to this, a Cr paracompact manifold admits a Or? Lorentz, metric if and only if it admits a non-vanishing Cr line clement field; by a line element field is meant an assignment of a pair of equal and opposite vectors (X, — X) at each point p of -#, i.e. a line clement field is like a vector field but with undetermined sign. To see this, let @ be a C1 positive definite metric defined on the manifold. ‘Then one can define a Lorentz metric g by 9%, ¥)0X,Z) 9(X, X) at each point p, where is one of the pair (X, —X) at p. (Note that as X appears an even number of times, it does not matter whether X or =X is chosen.) Then g(X,X) = —9(X, X), and if Y, Z are orthogonal to X with respect to @, they are also orthogonal to X with respect to 8 and g(¥,Z) = 9(¥,Z). Thus an orthonormal basis for @ is also an orthonormal basis for g. As @ is not unique, there are in fact many HY, Z) = G¥,Z)— 40 DIFFERENTIAL GEOMETRY [2.6 Lorentz metrics on .@ if there is one. Conversely, if ¢ is a given Lorentz metric, consider the equation gq,.X® = AQq,.X° where @ is any positive definite metric. This will have one negative and (n—1) positive eigenvalues. Thus the eigenvector field X corresponding to the negative eigenvalue will locally be a vector field determined up to asign and anormalizing factor; one can normalize it by ga, X*X? = — 1, so defining a line element ficld on #. In fact, any non-compact manifold admits a line element field, while a compact manifold does so if and only if its Euler invariant is zero (e.g. the torus 7’? does, but the sphere S? does not, admit a line element field). It will later turn out that a manifold can be a reasonable model of space-time only if it is non-compact, so there will exist many Lorentz metrics on .#. So far, the metric tensor and connection have been introduced as separate structures on .@. However given a metric ¢ on ., there is a unique torsion-free connection on .@ defined by the condition: the covariant derivative of g is zero, i.e. Teo:c = 0. (2.25) With this connection, parallel transfer of vectors preserves scalar products defined by g, so in particular magnitudes of vectors are invariant. For example if @/ét is the tangent vector to a geodesic, then g(8/ét, 2/2) is constant along the geodesic. From (2.25) it follows that X(GY¥,Z)) = Vx((¥, Z)) = Va9(¥, Z) + 9(Va¥,Z) +0(¥, VnZ) = (Vx¥,Z)+9(¥, Vx Z) holds for arbitrary C" vector fields X, ¥, Z. Adding the similar expres- sion for ¥(g(Z, X)) and subtracting that for Z(g(X, Y)) shows 2, Vx¥) = H—Z9(X, ¥)) + F(Z, X)) + X(¥,Z)) +9(Z,[X, ¥) +9(¥, (ZX) —9(X, [¥, ZD}. Choosing X, Y, Z as basis vectors, one obtains the connéction on one Pane = 9(Ea: Vig Ee) = Gaal oe in terms of the derivatives of the metric components gq, = (Eq, E,), and the Lie derivatives of the basis vectors. In particular, on using a coordinate basis these Lie derivatives vanish, so one obtains the usual Christoffel relations Dave = H8Gan] 2x? + 2g qq/ 82” — Ag ye| Ex} (2.26) for the coordinate components of the connection. 2.6) THE METRIC 41 From now on we will assume that the connection on @ is the unique Or-1 torsion-free connection determined by the Cr metric g. Using this ‘connection, one can define normal coordinates (§2.5) in a neighbour- hood of a point q using an orthonormal basis of vectors at g. In these coordinates the components gq, of g at q will be + 8, and the compo- nents I¢,, of the connection will vanish at g. By ‘normal coordinates’, we shall in future mean normal coordinates defined using an ortho- normal basis. ‘The Riemann tensor of the connection defined by the metric is a Cr-* tensor with the symmetry Reesyea = 0-<> Ravoa = — Roaca (2.274) in addition to the symmetries (2.21); as a consequence of (2.21) and (2.272), the Riemann tensor is also symmetric in the pairs of indices {ab}, {cd}, i.e. Ranea = Reaan- (2.276) ‘This implies that the Ricci tensor is symmetric: Rap = Boor (2.27¢) ‘The curvature scalar R is the contraction of the Ricci tensor: R= By = R%ygag. With these symmetries, there are jyn*(n?— 1) algebraically inde- pendent components of R4s.a, Where 7 is the dimension of M; 4n(m+ 1) of them can be represented by the components of the Ricci tensor. If n= 1, Ratea = 0; if n = 2 there is one independent component of Pasea, Which is essentially the function R. If x = 3, the Ricci tensor completely determines the curvature tensor; if n > 3, the remaining ‘components of the curvature tensor can be represented by the Weyl tensor Caoeas defined by 2 Cored = Ravea + = —3 Wala Raw + Foie Rana} + RGaicIav- (n—1)(n—2) As the last two terms on the right-hand side have the curvature tensor symmetries (2.21), (2.27), it follows that Cjye also has these sym- metries. One can easily verify that in addition, Ca ie, one can think of the Weyl tensor as that part of the curvature tensor such that all contractions vanish. 42 DIFFERENTIAL GEOMETRY [2.6 An alternative characterization of the Weyl tensor is given by the fact that it is a conformal invariant. The metrics @ and @ are said to be conformal if 2-0%8 (2.28) for some non-zero suitably differentiable function 2. Then for any vectors X, Y, V, W at a point p, 9X, ¥) _ 9X, ¥) a(V,W) av, Ww)’ #0 angles and ratios of magnitudes are preserved under conformal transformations; in particular, the null cone structure in 7, is pre- served by conformal transformations, since o(X,X) > 0, = 0, < 0>9(X,X) > 0, =0, <0, respectively. As the metric components are related by Dav = Q27%Yan, 9° = Q-*g%, the coordinate components of the connections defined by the metrics (2.28) are related by fe, = P,,+24 (2 +028 — nots) - (2.29) Calculating the Riemann tensor of , one finds Reb, g = Q-AR% 4 + 85.2%, 40-1(Q7), 969% — 2071), (Q>4) "48%, the covariant derivatives in this equation are those determined by the metric g. Then (assuming n > 2) Bg = QAR + (= 2) IO), og" = (= 2) AOA) GO and O% 2a = Osea the last equation expressing the fact that the Weyl tensor is con- formally invariant. These relations imply R= Q-*R-2(n— 1) 2-9Q,,ag*t = (n— 1) (n— 4) 2-4, 2, gg". (2.30) Having split the Riemann tensor into a part represented by the Ricci tensor and a part represented by the Weyl tensor, one can use the Bianchi identities (2.22) to obtain differential relations between the Ricci tensor and the Weyl tensor: contracting (2.22) one obtains Revea;a = Rosse Re (2.31) where 9%: 2.6] THE METRIC 43 and contracting again one obtains Re, From the definition of the Weyl tensor, one can (if n > 3) rewrite (2.31) in the form Ceara = 22S (Russa 5G Gy aR) - (2.32) 4B, If nm < 4, (2.31) contain all the information in the Bianchi identities (2.22), so if n = 4, (2.32) are equivalent to these identities. A diffeomorphism ¢:.4->.# will be said to be an isometry if it carries the metric into itself, that is, if the mapped metric J, @ is equal to g at every point. Then the map $,:7,> Tig) Preserves scalar products, as HEY) p = bx Gu X, bu YY gen = hu X Se¥)] peor If the local one-parameter group of diffeomorphisms ¢, generated by a vector field K is a group of isometries (i.e. for each ¢, the trans- formation ¢, is an isometry) we call the vector field K a Killing vector field. The Lie derivative of the metric with respect to K is a! Inf = lim} (@— $148) = 0, since £ = $148 for] each ¢. But from (2.17), Dgap = 2Kia;o» 80 & Killing vector field K satisfies Killing’s equation Kajo+Ko;a = 0. (2.33) Conversely, if K is a vector field which satisfies Killing’s equation, then Inf = 0,80 Fin Bl = Bly+ [apr aB)ln at” td = B+ [igre Ben elt f'(bredeene)_ | ‘ = Bly— [Bex Cele.) At = lye Thus K is a Killing vector field if and only if it satisfies Killing’s equa- tion. Then one can locally choose coordinates x* = (2”,t)(v = 1ton—1) 44 DIFFERENTIAL GEOMETRY [26 such that Ke = takes the form x*/0t = 84,; in these coordinates Killing’s equation an] ét = 0 Jay = Ja(@). A general space will not have any symmetries, and so will not admit any Killing vector fields. However a special space may admit r linearly independent Killing vector fields K, (a = 1,...,r). It ean be shown that the set of all Killing vector fields on such a space forms a Lie algebra of dimension r over 2, with the algebra product given by the Lie bracket [ , ] (see (2.16)), where 0 0, then (2%, ...,2%) are oriented local coordinates for 0(S). Ifg is a metric on, the imbedding will induce a metric *g on F, where if X,YeT, 0*9(X,¥)|p = 9(04X,04¥)|aq). This metric is sometimes called the first fundamental form of ¥. If g is positive definite the metric 6*g will be positive definite, while if g is Lorentz, 0*g will be (a) Lorentz if g**n,, > 0 (in this case, 0(Y) will be said to be a timelike hypersurface), 2.7) HYPERSURFACES 45 (0) degenerate if gn, = 0 (in this case, 6(/) will be said to be a null hypersurface), (©) positive definite if g*°n,n, < 0 (in this case, (Sf) will be said to be a spacelike hypersurface). To see this, consider the vector N? 1ag™. This will be orthogonal to all the vectors tangent to 0(), i.e. to all vectors in the subspace H=0,(T,) in Tyg). Suppose first that N does not itself lie in this subspace. Then if (Eq, ..., E,) are a basis for 7, (N, 04(Eq), ---,Ox(En)) will be linearly independent and so will be a basis for Ti). The compo- nents of ¢ with respect to this basis will be . ea o ) - en ° a) a 0 GG4(E), O48) ~\ 0 *9(E,,E,), As the metric g is assumed to be non-degenerate, this shows that g(N,N) + 0. If g is positive definite, g(N, N) must be positive and so the induced metric 6*g must also be positive definite. If g is Lorentz and g(N,N) = g*nqn, < 0, then 0*g must be positive definite since the matrix of the components of g has only one negative eigenvalue. Similarly if g(N, N) = g%nqn, > 0, then 6*g will be a Lorentz metric. Now suppose that N is tangent to 6(). Then there is some non-zero vector Xe, such that 0,(X) = N. But g(N,0,Y) = 0 for all Ye7,, which implies 6* (X,Y) = 0. Thus 6*g is degenerate. Also, taking Y to be X, g(N, N) = g%nam, = 0. If g*nqny + 0, one can normalize the normal form n to have unit magnitude, ie. gnqn = + 1. In this case the map 0*: T*4,)>7*, will be one-one on the (n—1)-dimensional subspace H*g,) of T’* ap) consisting of all forms w at 6(p) such that g%n,w, = 0, because 6*n = 0 and n does not lie in H*. Therefore the inverse (6*)~? will be amap 0, of T*, onto H*q,), and so into T'* 4). This map can be extended in the usual way to a map of covariant. tensors on to covariant tensors on 6(/) in .@; as there already is amap 0, of contravariant tensors on S to 6(/), one can extend 0, to amap 6, of arbitrary tensors on / to 0). This map has the property that 9,T has zero contraction with n on all indices, i.e. eT Y**,.aMg =O and (G_T)*", age = 0 for any tensor TeT;(S). ‘The tensor h on 6(/) is defined by h = 6,(0*8). In terms of the normalized form n (remember g”n,7, = + 1), Tran = Gan F MoM since this implies 6*h = 6*g and h,,g’"n, = 0. 46 DIFFERENTIAL GEOMETRY [27 The tensor h*, = g*hg, is a projection operator, i.e. h%,h?, projects a vector X € Ty,) into its part lying in the subspace H = 0,(T,) of Tx,» tangent to 0), Xo = he, XP + nny X?, where the second term represents the part of X orthogonal to 0). Also ht, projects a form w¢7"y, into its part lying in the subspace Hyg: Og = Wa tty gn ery. Similarly one can project any tensor T €7';(0(p)) into its part in HOP) = Hyy®---@ Hoy @ Hip ®-.- OH, factors @ factors ie. its part which is orthogonal to n on alll indices. The map 6, is one-one from 7, to Hy). Therefore one can define a map 6* from Ty,) to T, by first projecting with h4, into Hy.) and then using the inverse (@),~*. As one already has a map 0* of forms on OY) to forms on ¥, one can extend the definition of * to a map 6* of tensors of any type on O(S) to tensors on Y. This map has the property that 6*(@,T) = T for any tensor TeT%(p) and 0,(9*T) = T for any tensor T 0. Therefore if is orientable the n-formsy defined by coordinate bases of an oriented atlas will be identical, i.e. given an orientation of .#, one can define a unique n-form field n, the canonical n-form, on #. ‘The contravariant antisymmetric tensor 99S Fey abot has components aid = (— HP! [gh 80, 8%... 34, where ¢ is the signature of g (so }(n—e) is the number of negative eigenvalues of the matrix of metric components (gq,)). Therefore these tensors satisfy the relations Aig n= (— WO 84,3)... By, (2.37) ‘The Christoffel relations imply that the covariant derivatives of ap... 2nd 7-4 with respect to the connection defined by the metric vanish, i Using the canonical n-form, one ean define the volume (with respect to the metric g) of an n-dimensional submanifold @ as Xf, a. ride ‘Thus y can be regarded as a positive definite volume measure on #. ‘We shall often use it in this sense, and shall denote it by dv. Note that dis not meant to represent the exterior differential operator here; dv is simply a measure on .@. If f is a function on #, one can define its integral over & with respect to this volume measure as Sifeens flm 2.8) VOLUME ELEMENT AND GAUSS’ THEOREM 49 With respect to local oriented coordinates {x*}, this can be expressed as the multiple integral f. flgltdxtdzt... dz, . which is invariant under a change of coordinates. If X is a vector field on .#, its contraction with y will be an (n— 1)- form field X 9, where (X.ne..a = Xa... This (n—1)-form may be integrated over any (n—1)-dimensional compact orientable submanifold ¥~. We write this integral as [xrae = aon J,=" where the canonical form y is regarded as defining a measure-valued form de, on the submanifold ¥. If the orientation of Y is given by the direction of the normal form m, then de, can be expressed as n,do where do is a positive definite volume measure on the sub- manifold ¥°. The volume measure dc is not unique unless the normal n, is normalized. If n, is normalized to unit magnitude in a metric ¢ onM, i.e. ngnyg% = + 1, then do is equal to the volume measure on ¥~ defined by the induced metric on ¥ (to see this, simply choose an orthonormal basis with n,9% as one of the basis vectors). Using the canonical form, one can derive Gauss’ formula from Stokes’ theorem: for any compact n-dimensional submanifold @ of @, 1 ir ra a . —,; |} d(X.»). Poe ¥8800= ata I ¥9 = tag face (dX .9))a...de = (— "(KN gta...€) 563 HY Bg Bal Mae Xu But mynd od ting tga aX? = Ma...aeF%e-0 HOV XP,u = Ma... aX sg on using relation (2.37) twice. Therefore f, eae, = f X¢,ydv ow * 50 DIFFERENTIAL GEOMETRY [2.8 holds for any vector field X; this is Gauss’ theorem. Note that the orientation on & for which this theorem is valid is that given by the normal form y such that k) manifold .@ is a Ct manifold & and a O* surjective map 7: & >. The manifold & is called the total space, -# is called the base space and 7 the projection. Where no confusion. can arise, we will denote the bundle simply by &. In general, the inverse image 7-1(p) of a point pe. need not be homeomorphic to 7-\(q) for another point g¢-#. The simplest example of a bundle is a product bundle ( x of, M, 7) where sf is some manifold and the projection 7 is defined by 7(p, v) = p for all pe.#, ve xf. For example, if one chooses -# as the circle St and of as the real line R}, one con. structs the cylinder C* as a product bundle over S!, A bundle which is locally a product bundle is called a fibre bundle. Thus a bundle is a fibre bundle with fibre F if there oxists a neighbour- hood & of each point g of # such that 7-1(%) is isomorphic with & x F, in the sense that for each point pe ® there is a diffeomorphism ¢, of 7-\(p) onto F such that the map y defined by y(u) = (1(t), Psi) is 8 diffeomorphism y: n-\(%) > & x F. Since # is paracompact, wo can choose a locally finite covering of @ by such open sets @,. If %, and %,s are two members of such a covering, the map ($a,0) 0 (Bp, 9") is a diffeomorphism of ¥ onto itself for each pe (%, n &,). The inverse images -1(p) of points pe.M are therefore necessarily all diffeo- morphic to ¥ (and so to each other). For example, the Mébius strip is a fibre bundle over S! with fibre R*; we need two open sets W,, UV, 2.9) FIBRE BUNDLES 51 togivea covering by sets of the form %, x Rt. This example shows that if a manifold is locally the direct product of two other manifolds, it is nevertheless not, in general, a product manifold; it is for this reason that the concept of a fibre bundle is so useful. The tangent bundle T(#) is the fibre bundle over a O* manifold obtained by giving the set & = UZ its natural manifold structure >, and its natural projection into .@. Thus the projection 7 maps each point of Z, into p. The manifold structure in & is defined by local coordinates {z4} in the following way. Let {2*} be local coordinates in an open set % of .M. Then any vector Ve, (for any pe) can be expressed as V = V‘d/éz4|,. The coordinates {24} are defined in (2) by {24} = {2*, V2}. On choosing a covering of W by coordinate neighbourhoods %,, the corresponding charts define a C— atlas on & which turn it into a C1 manifold (of dimension n*); to check this, one needs only note that in any overlap (@, n %,) the coordinates {2',} of 4 point are O* functions of the coordinates {z‘,} of the point, and the components {V4,} of a vector field are C*-1 functions of the compo- nents {Vy} of the vector field. Thus in 7-1(%, 0.%,), the coordinates {e4,} are C+ functions of the coordinates {24,}. The fibre 7-1(p) is 7, and s0 is a vector space of dimension n. This vector space structure is preserved by the map ¢,, y: T)-> R, which is given by ga, y(t) = V4(u), i.e. $,,) maps a vector at p into its com- ponents with respect to the coordinates {x,}. If {x4,} are another set of local coordinates then the map (¢z, »)0(¢,, »~) is a linear map of E* onto itself. Thus it is an element of the general linear group @L(n, R) (the group of all non-singular » x n matrices). The bundle of tensors of type (r,s) over M, denoted by TY), is defined in a very similar way. One forms the set. & = LU T%(p), defines the projection 7 as mapping each point in 7%(p) into p, and, for any coordinate neighbourhood &% in -#, assigns local coordinates {24} to 4H) by {24} = fat, TY, | 4} where {z"} are the coordinates of the point p and {7'--®,._,} are the coordinate components of T (that is, T= Te, 40/0x*®...@dx“|,). This turns & into a C*-1 manifold of dimension 2t*#1; any point u in T%(@) corresponds to a unique tensor T of type (r,s) at 7(w). The bundle of linear frames (or bases) L(.#) is a C*-1 fibre bundle defined as follows: the total space & consists of all bases at all points of @, that is all sets of non-zero linearly independent n-tuples of vectors {E,}, E,€7,, for each pe.# (aruns from 1 ton). The projection 52 DIFFERENTIAL GEOMETRY [2.9 77 is the natural one which maps a basis at a point p to the point p. If {a} are local coordinates in an open set Y < -#, then G4} = (x9, By, Bat, .. By} are local coordinates in 7—(@), where E, is the jth components of the vector E, with respect to the coordinate bases @/éz‘. The general linear group GZ(n,R) acts on L(.#) in the following way: if {E,} is a basis at pe.M, then AcGL(n, R) maps u = {p,E,} to A(™) = {P Aan Ee}. ‘When there is a metric g of signature s on -#, one can define a sub- bundle of L(-@), the bundle of orthonormal frames O(-#), which con- sists of orthonormal bases (with respect to g) at all points of @. O(.#) is acted on by the subgroup O(}(n+8), }(m—s)) of GL(n, R). This consists of the non-singular real matrices 4,, such that AarGreAae = Faas where Gp, is the matrix diag (+1, +1,..., +4, —1,-1. 1). Hats) terms 4(n—s) terms It maps (p,E,)€O(#) to (p, Ag» Ey) €O(.#). In the case of a Lorentz, metric (i.e. s = n~ 2), the group O(n —1, 1) is called the n-dimensional Lorentz group. A.C" cross-section of a bundle is a Cr map ©: .-> & such that 70 ® is the identity map on .#; thus a cross-section is a Or assignment to each point p of of an element (p) of the fibre 7p). A cross- section of the tangent bundle 7(.#) is a vector field on .@; a cross- section of T;(@) is a tensor field of type (r, 2) on @; a cross-section of L(A) is a set of n non-zero vector fields {E,} which are linearly inde- pendent at each point, and a cross-section of O(.#) is a set of ortho- normal vector fields on .@. ‘Since the zero vectors and tensors define cross-sections in 7(-#) and Ti), these fibre bundles will always admit cross-sections’ If is orientable and non-compact, or is compact with vanishing Euler number, there will exist nowhere zero vector fields, and hence cross- sections of 7(-#) which are nowhere zero. The bundles L(#) and O(.#) may or may not admit cross-sections; for example L(S*) does not, but L(R*) does. If L(-) admits a cross-section, is said to be parallelizable. R.P.Geroch has shown (1968c) that a non-compact four-dimensional Lorentz manifold .@ admits a spinor structure if and only if it is parallelizable. 2.9) FIBRE BUNDLES 53 ‘One can describe a connection on # in an elegant geometrical way in terms of the fibre bundle L(.#). A connection on-@ may be regarded asa rule for parallelly transporting vectors along any curve y(t) in .@. Thus if (E,} is a basis at a point p = y(t), ie. {p,E,} is a point u in L(#), one can obtain a unique basis at any other point y(t), i.e. a unique point 7(¢) in the fibre 7-*(y(t)), by parallelly transporting {E,} along y(t). Therefore there is a unique curve 7(¢) in L(#), called the lift of y(t), such that: (1) Flt) = 4 2) mH) = 70, (8) the basis represented by the point 7(¢) is parallelly transported along the curve y(t) in. In terms of the local coordinates {24}, the curve 7(t) is given by {*(7(t)), Byyt(0)}, where aE s(t) dé dxa(y(t)) dé + By Mas: . Consider the tangent space 7(L()) to the fibre bundle L(.@) at the point u. This has a coordinate basis {2/224|,}. The n-dimensional subspace spanned by the tangent vectors {(2/2f)xo|,) to the lifts of all curves y(t) through 7 is called the horizontal subspace H,, of T(L())- In terms of local coordinates, (8). = EM) @ Ent 2 Bi), at oa at a, det(y(O) (_2 a PO (Ge Pe woeg)> s0 a coordinate basis of H, is {2/22 — Hyd Pg, 2/0E}. Thus the con- nection in.# determines the horizontal subspaces in the tangent spaces at each point of L(-@). Conversely, a connection in .# may be defined by giving an n-dimensional subspace of T,(L(.#)) for each ue L() with the properties: (1) If AcG@L(n, RY), then the map 4g: TL) > TaiL(#)) maps the horizontal subspace H,, into Haq); (2) H, contains no non-zero vector belonging to the vertical sub- space F,, Here, the vertical subspace V,, is defined as the n*-dimensional subspace of 7,(L(#)) spanned by the vectors tangent to curves in the fibre 71(7(x)); in terms of local coordinates, V, is spanned by the B4 DIFFERENTIAL GEOMETRY (2.9 vectors {/2,,*}. Property (2) implies that 7, is the direct sum of H, and F,. The projection map 7: L(#)+>.M induces a surjective linear map ta: TAL(M))-> Tyg), such that 74(V,) = 0 and 7, restricted to H, is 1-1 onto Ty). Thus the inverse 14-1 is a linear map of Ty,(-@) onto H,. Therefore for any vector X €T,,(-@) and point we 1-\(p), there is a unique vector XeH,,, called the horizontal lift of X, such that 4(X) = X. Given a curve y(t) in, and an initial point win 7—(y(t,)), one can construct a unique curve 7(t) in L(#), where 7(t) is the curve through u whose tangent vector is the horizontal lift of the tangent vector of y(t) in . Thus knowing the horizontal subspaces at each point in L(#), one can define parallel propagation of bases along any curve y(t) in @, One can then define the covariant derivative along 7(t) of any tensor field T by taking the ordinary derivatives with respect to ¢, of the components of T with respect to a parallelly propagated basis. If there is a metric g on @ whose covariant derivative is zero, then orthonormal frames are parallelly propagated into orthonormal frames. Thus the horizontal subspaces are tangent to O(-@) in L(-#), and define a connection in O(.#). Similarly a connection on defines n-dimensional horizontal sub- spaces in the tangent spaces to the bundles 7(.) and T;(.#), by parallel propagation of vectors and tensors. These horizontal sub- spaces have coordinate bases é a {e- if Paso} aly -+ (all upper indices) Te gD%.—(all lower indices)) } respectively. As with L(), 7 maps these horizontal subspaces one-one onto Tyo(-M); thus again 74 can be inverted to give a unique horizontal lift X ¢ 7, of any vector Ke Ty). In the particular case of T(-4), u itself corresponds to a unique vector We Ty(@), and so there is an intrinsic horizontal vector field W defined on T(-@) by the connection. In terms of local coordinates {x*, V°}, = ellie W-v (ge-VPegyn)- are 2.9) FIBRE BUNDLES 56 This vector field may be interpreted as follows: the integral curve of W through u = (p,X)€7() ia the horizontal lift of the geodesic in M with tangent vector X at p. Thus the vector field W represents all geodesics on .#. In particular, the family of all geodesics through ped is the family of integral curves of W through the fibre 1p) < T(M); the curves in M have self intersections at least at p, but the curves in T(@) are non-intersecting everywhere. 3 General Relativity In order to discuss the occurrence of singularities and the possible breakdown of General Relativity, it is important to have a precise statement of the theory and to indicate to what extent it is unique. We shall therefore present the theory as a number of postulates about a mathematical model for space-time. In §3.1 we introduce the mathematical model and in § 3.2 the first two postulates, local causality and local energy conservation. These postulates are common to both Special and General Relativity, and thus may be regarded as tested by the many experiments that have been performed to check the former. In $3.3 we derive the equations of the matter fields and obtain the energy-momentum tensor from a. ‘Lagrangian. The third postulate, the field equations, is given in § 3.4. This is not 50 well established experimentally as the first two postulates, but we shall see that any alternative equations would seem to have one or more undesirable properties, or else require the existence of extra fields which have not yet been detected experimentally. 3.1. The space-time manifold ‘The mathematical model we shall use for space-time, i.e. the collection of all events, is a pair (.,8) where . is a connected four-dimensional Hausdorff C* manifold and g is a Lorentz metric (i.e. a metric of signature +2) on @. NN Two models (.,g) and (.#’,8') will be taken to be equivalent if they are isometric, that is if there is a diffeomorphism 6:.@>.4’ which carries the metric g into the metric g’, ie. 6,8 = 9’. Strictly speaking then, the model for space-time is not just one pair (M, 8) but a whole equivalence class of all pairs (.@’, ’) which are equivalent to (#,8). We shall normally work with just one representative mem- ber (.#, 8) of the equivalence class, but the fact that this pair is defined only up to equivalence is important in some situations, in particular in the discussion of the Cauchy problem in chapter 7. 156) 3.4] THE SPACE-TIME MANIFOLD 37 The manifold .# is taken to be connected since we would have no knowledge of any disconnected component. It istaken to be Hausdorff since this seems to accord with normal experience. However in chapter 5 we shall consider an example in which one might dispense with this condition. Together with the existence of a Lorentz metric, the Hausdorff condition implies that .@ is paracompact (Geroch (1968¢)). A manifold corresponds naturally to our intuitive ideas of the con- tinuity of space and time. So far this continuity has been established for distances down to about 10-5cm by experiments on pion scat- tering (Foley et al. (1967)). It may be difficult to extend this to much smaller lengths as to do so would require a particle of such high energy that several other particles might be created and confuse the experi- ment. Thus it may be that a manifold model for space-time is inap- propriate for distances less than 10-%cm and that we should use theories in which space-time has some other structure on this seale. However such breakdowns of the manifold picture would not be expected to affect General Relativity until the typical gravitational length scale became of that order. This would happen when the density became about 10°* gm om-%, which is a condition so extreme as to be completely beyond our present knowledge. Nevertheless, by adopting a manifold model for space-time, and making certain other reasonable assumptions, we shall show in chapters 8-10 that some breakdowns of General Relativity must occur. It may be the field equations that go wrong, or it may be that quantization of the metric is needed, or it may be a breakdown of the manifold structure itself that occurs. The metric g enables the non-zero vectors at a point pe.# to be divided into three classes: a non-zero vector Xe, being said to be timelike, spacelike or null according to whether g(X, X) is negative, positive or zero respectively (ef. figure 5). The order of differentiability, r, of the metric ought to be sufficient for the field equations to be defined. They can be defined in a distribu- tional sense if the metric coordinate components gy and g* are con- tinuous and have locally square integrable generalized first derivatives with respect to the local coordinates. (A set of functions f,, on R" are said to be the generalized derivatives of a function f on R* if, for any C function y- on R” with compact support, [ravrare = — [popioay are, 58 GENERAL BELATIVITY [3.1 However this condition is too weak, since it guarantees neither the existence nor the uniqueness of geodesics, for which a C?- metric is required. (A C*- metric is one for which the first coordinate derivatives of the metric coordinate components satisfy a local Lipschitz condi- tion, see § 2.1.) We shall in fact assume for most of the book that the metric is at least C®. This allows the field equations (which involve the second derivatives of the metric) to be defined at every point. In § 8.4 we shall weaken the condition on the metric to ?- and show that this does not affect the results on the occurrence of singularities. In chapter 7, we use a different kind of differentiability condition in order to show that the time development of the field equations is determined by suitable initial conditions. We require there that the metric components and their generalized first derivatives up to order m(m > 4) are locally square integrable. This would certainly be true if the metric were C4, In fact, the order of differentiability of the metric is probably not physically significant. Since one can never measure the metric exactly, but only with some margin of error, one could never determine that there was an actual discontinuity in its derivatives of any order. Thus one can always represent one’s measurements by a C* metric. If the metric is assumed to be CO’, the atlas of the manifold must be Cr+, However, one can always find an analytic subatlas in any C* atlas ( > 1) (Whitney (1936), cf. Munkres (1954). Thus it is no restriction to assume from the start that the atlas is analytic, even though one could physically determine only a Cr+? atlas if the metric were C’. We have to impose some condition on our model (4,8) to ensure that it includes all the non-singular points of space-time. We shall say that the Cr pair (.4’,8’) is a Cr-eztension of (,8) if there is an iso- metric C* imbedding 4: .@>.4’. If there were such an extension (4, 8’) we should have to regard points of .@’ as also being points of space-time. We therefore require that the model (.#,) is Or- inextendible, that is there is no C* extension (.4’, 8’) of @#, 8) where HM) does not equal 4’. Asan example of a pair (4, 8,) which is not inextendible, consider two-dimensional Euclidean space with the z-axis removed between #,=~1and x, = +1. The obvious way to extend this would simply be to replace the missing points, but one could also extend it by taking another copy (2,2) of the space, and identifying the bottom side of the «,-axis for |z,| < 1 with the top side of the 2,-axis for |z,| < 1, and also identifying the top side of the ,-axis for |z,| < 1 with the 3.4] THE SPACE-TIME MANIFOLD 59 bottom side of the z-axis for |zrq| < 1. The resultant space (M,, 23) is inextendible but not complete as we have left out the points x, = +1, = 0. We cannot put these points back in because we were perverse enough to extend the top and bottom sides of the x-axis on different sheets. If however one takes the subset % of #, defined by 1 < 2, < 2, —1 <4, <1, then one could extend the pair (%, 5/4) and put back the point x = 1, y; = 0. This motivates a rather stronger definition of inextendibility: a pair (#,) is said to be Cr-locally inextendible if there is no open set Y < .@ with non-compact closure in, such that the pair (%, 8|q) has an extension (%’, g’) in which the closure of the image of % is compact. 3.2 The matter fields ‘There will be various fields on .#, such as the electromagnetic field, the neutrino field, ete., which describe the matter content of space-time. These fields will obey equations which can be expressed as relations between tensors on .# in which all derivatives with respect to position are covariant derivatives with respect to the symmetric connection defined by the metric g. This is so because the only relations defined by @ manifold structure are tensor relations, and the only connec- tion defined so far is that given by the metric. If there were another connection on .#, the difference between the two connections would be a tensor and could be regarded as another physical field. Similarly another metric on .@ could be regarded as a further physical field. (The equations of the matter fields are sometimes expressed as relations between spinors on .&. We do not deal with such relations in this book, as they are not needed for the problems we wish to consider. In fact, all spinor equations can be replaced by rather more complicated tensor equations; see e.g. Ruse (1937).) ‘The theory one obtains depends on what matter fields one incorpo- rates in it. One should of course include all such fields which have been experimentally observed, but one might postulate the existence of as yet undetected fields. Thus for example Brans and Dicke (Dicke (1964), appendix 7) postulate the existence of a long range scalar field which is weakly coupled to the trace of the energy-momentum tensor. In the form given in Dicke (1964) appendix 2, the Brans~Dicke theory can be regarded simply as General Relativity with an extra scalar field. Whether this scalar field has been experimentally detected or not is at present under dispute. 60 GENERAL RELATIVITY (3.2 We shall denote the matter fields included in the theory by ‘Yo"*,,.. ay Where the subscript (é) numbers the fields considered. The following two postulates on the nature of the equations obeyed by the Foor are common to both the Special and the General Theories of Relativity. Postulate (a): Local causality The equations governing the matter fields must be such that if & is ‘@ convex normal neighbourhood and p and q are points in & then a signal can be sent in Y between p and q if and only if p and q can be joined by a 01 curve lying entirely in &, whose tangent vector is every- where non-zero and is either timelike or null; we shall call such a curve, non-spacelike. (Our formulation of relativity excludes the possibility of particles such as tachyons, which move on spacelike curves.) Whether the signal is sent from p to q or from q to p will depend on the direction of time in %. The problem of whether a consistent direction of time can be assigned at all points of space-time will be considered in §6.2. A more precise statement of this postulate can be given in terms of the Cauchy problem of the matter fields. Let pe be such that every non-spacelike curve through p intersects the spacelike surface «* = 0 within %. Let ¥ be the set of points in the surface 2¢ = 0 which can be reached by non-spacelike curves in Y from p. Then we require that the values of the matter fields at p must be uniquely determined by the values of the fields and their derivatives up to some finite order on F, and that they are not uniquely determined by the values on any proper subset of F to which it can be continuously retracted. (For a fuller discussion of the Cauchy problem, see chapter 7.) It is this postulate which sets the metric apart from the other fields on @ and givesit its distinctive geometrical character. If {x7} aro normal coordinates in % about p, it is intuitively fairly obvious (and is proved in chapter 4) that the points which can be reached from’p by non-spacelike curves in & are those whose coordinates satisfy (2A) + (28+ (29) (AP <0. ‘The boundary of these points is formed by the image of the null cone of p under the exponential map, that is the set of all null geodesics through p. Thus by observing which points can communicate with p, one can determine the null cone N, in 7,,. Once N, is known, the metric at p may be determined up to a conformal factor. This may be seen as 3.2] THE MATTER FIELDS 61 follows: let X,Y, be respectively timelike and spacelike vectors, The equation G(X +AY, K+AY) = g(X, X) + 2Ag(X, Y) + A*G(¥, Y) will have two real roots A, and A, as g(X,X) < 0 and g(¥,¥) > 0. If X, is known, A, and A, may be determined. But AAs = 9X, X)/9(¥, ¥). ‘Thus the ratio of the magnitudes of a timelike vector and a spacelike vector may be found from the null cone. Then if W and Z are any two non-null vectors at p, g(W, Z) = 2(9(W, W) +9(Z,Z)—g(W+Z,W+Z)). Each of the magnitudes on the right-hand side may be compared with the magnitude of either X or Y, and so g(W, Z)/9(X, X) may be found. (If W+Z is null, the corresponding expression involving W+2Z could be used.) Thus observation of local causality enables one to measure the metric up to a conformal factor. In practice this measure- ment is performed most conveniently using the experimental fact that no signal has been observed to travel faster than electromagnetic radiation. This means that light must travel on null geodesics. This however is a consequence of the particular equations the electro- magnetic field obeys, not of the theory of relativity itself. Causality will be considered further in chapter 6. Among other results, it will be shown that causal relations may be used to determine the topological structure of .. The conformal factor in the metric may be determined using postulate (b) below; thus all the elements of the theory will be physically observable. Postulate (b): Local conservation of energy and momentum ‘The equations governing the matter fields are such that there exists a symmetric tensor 7, called the energy-momentum tensor, which depends on the fields, their covariant derivatives, and the metric, and which has the properties: (i) 7? vanishes on an open set ¥ if and only if all the matter fields vanish on &, (ii) 7 obeys the equation Te, = 0. (3.1) 62 GENERAL RELATIVITY [3.2 Condition (i) expresses the principle that all fields have energy. One might possibly object to the ‘only if? on the grounds that there might be two non-zero fields, one of whose energy-momentum tensor exactly cancelled that of the other. This possibility is related to that of the existence of negative energy which will be discussed in §3.3. If the metric admits a Killing vector field K, equations (3.1) can be integrated to give a conservation law. To see this, define P# to be the vector whose components are P? = T*>K,, Then, Pa, = 7%, Ky + TK. The first term is zero by the conservation equations, and the second vanishes as 7 is symmetric and 2K,» = DxGas = 0, since K is Killing vector. Thus if 9 is a compact orientable region with boundary @@, Gauss’ theorem (§2.7) shows Prado, -f, Po.,dv= 2 (3.2) @ This may be interpreted as saying that the total flux over a closed surface of the K-component of energy-momentum is zero. ‘When the metric is flat, as it is in the Special Theory of Relativity, ‘one may choose coordinates {2} in which the components of the metric Te Jqp = €z ap (NO summation) where d,, is the Kronecker delta and eq is —1 ifa=4 and is +1 if a= 1,2,3. Then the following are Killing vectors: = a/aex (a = 1, 2, 3, 4) (these generate four translations) and @ M = e,22 ~ et & (no summation; a, = 1, 2, 3, 4) we (these generate six ‘rotations’ in space-time). These isometries form the ten-parameter Lie group of isometries of flat space-time known as the inhomogeneous Lorentz group. One may use them to define ten vectors Pe and P* which will obey (8.2). We may think of P as repre- senting the Eat energy and P, P, P as the flow of the three compo- nents of linear momentum. The P can be interpreted as the flow of oo angular momentum. If the metric is not flat there will not, in general, be any Killing vectors and so the above integral conservation laws will not hold, How- ever, in a suitable neighbourhood of a point q one may introduce 3.2] THE MATTER FIELDS 63 normal coordinates {2%}. Then at q the components gq, of the metric are 849 (no summation), and the components I™,, of the connection are zero. One may take a neighbourhood @ of q in which the gay and I", differ from their values at g by an arbitrarily small amount; then the La; and Ma,» will not exactly vanish in 9, but will in this neigh- bourhood differ from zero by an arbitrarily small amount. Thus f. Pdo, and [ Pedo, v0 vox will still be zero in the first approximation; that is to say, one still has approximate conservation of energy, momentum and angular momentum in a small region of space-time. Using this it can be shown that a small isolated body moves approximately on a timelike geodesic curve independent of its internal constitution provided that the energy density of matter in it is non-negative (for an account of the motion of asmall body in relativity, see Dixon (1970)). This may be thought of as Galileo’s principle that all bodies fall equally fast. In Newtonian terms one would say that the inertial mass (the m in F = ma) and the passive gravitational mass (the mass acted on by a gravitational field) are equal for all bodies. This has been verified to a high order of accuracy in experiments by Edtvos and by Dicke (1964). Postulate (a) enables one to measure the metric up to a conformal factor at each point. Using postulate (b) one may relate these factors at different points, for the conservation equations Te, = 0 would not in general hold for a connection derived from a metric @ = 0%. One way of doing this would be to observe the paths of small ‘test’ particles and so to determine the timelike geodesic curves. Then if y(t) is such a curve with tangent veotor K = (2/ét),, one has from (2.29) BK Re + 20710, , KK*— O-1( KK, 9040, 4. Since y(t) is a geodesic with respect to the space-time metric g, Ke(D/at) K = 0. Thus od Ka = — (KK 49,4) Kg (log Q),.. (3.3) Knowing the conformal structure, one can choose a metric @ which represents the conformal equivalence class of metrics and can evaluate the left-hand side of (3.3) for any test particle. Then the right-hand side of (3.3) determines (log Q),, up to the addition of a multiple of K%9,9. 64 GENERAL RELATIVITY [32 By considering another curve y'(t) whose tangent vector K’* is not parallel to K*, one can find (log Q),,, and so can determine Q every- where up to a constant multiplying factor. This constant factor specifies one’s units of measurement, and so can be chosen. arbitrarily. This is, of course, not the way one measures the conformal factor in practice; ono makes use of the fact that there exist a large number of similar systems (such as the electronic states of atoms) whose internal motions define a number of events along the timelike curve which represents their position in space-time. The intervals between these events seem to be independent of their past history in the sense that the intervals measured by two nearby systems correspond. If one can effectively isolate them against external matter fields (so they must move on geodesic curves) and if one assumes their internal motion is independent of the curvature of space-time, then the only thing it can depend on is the metric. Thus the are-length between two successive events on a curve must be the same for each pair of successive events onany such curve. If one takes this aro-length as one’s unit of measure- ment, one can determine the conformal factor at any point of space- time. In fact it may not be possible to isolate a system from external matter fields, Thus for example in the Brans-Dicke theory there is a scalar field which is non-zero everywhere. However the conformal factor can still be determined by the requirement that the conserva- tion equation 7%, , = 0 should hold. Thus knowledge of the energy- momentum tensor J, determines the conformal factor. 3.3 Lagrangian formulation ‘The conditions (i) and (ii) of postulate (6) do not tell one how to con- struct the energy-momentum tensor for a given set of fields, or whether it is unique. In practice one relies heavily on one’s intuitive knowledge of what energy and momentum are. However, there is a definite and unique formula for the energy-momentum tensor in the case that the equations of the ficlds can be derived from a Lagrangian. Let L be the Lagrangian which is some scalar function of the fields Yo"....@. their first covariant derivatives, and the metric. One obtains the equations of the fields by requiring that the action rf Ldv 2 3.3] LAGRANGIAN FORMULATION 65 be stationary under variations of the fields in the interior of a compact four-dimensional region Z. By a variation of the fields Vg?" gin D we mean a one-parameter family of fields Wo(u,r) where we(—e,6) and re.@, such that G@) For) = Folr), Gi) Yor) =Keolr) when red-9. We denote PLo(u,7/2ulvo by AX ‘Then al oL ete Me. a Silo BL ange a OF ore, where 'Vig*"¥,,..;¢ are the components of the covariant derivatives of Bio. But ACK (AY o%,,..a),6 thus the second term can be expressed as fal The first term in this expression can be written as fierade= fi rds, Hy 7 where Q is a vector whose components are AX%e ‘This integral is zero as condition (ii) is the statement that AY vanish at the boundary @@. Thus in order that aI/2u|,.9 should vanish for all variations on all volumes 9, it is necessary and sufficient that the Euler-Lagrange equations, eb Seca =0, (3.4) lh (ra hold for all i. These are the equations of the fields. We obtain the energy-momentum tensor from the Lagrangian by considering the change in the action induced by a change in the metric. 66 GENERAL RELATIVITY 13.3 Suppose a variation gqp(u,7) leaves the fields Vio% but alters the components g,, of the metric. Then aq aL a OL Bela Bama ac BF wt tat age ha) oH ai es feo ace. (3.5) a unchanged The last term arises because the volume measure du depends on the metric, and so will vary when the metric is varied. To evaluate this term, recall that dv is in fact the four-form (4!)-1y whose compo- nents are Tasca = (—g)#4! dig! 5,26, Sait, where g = det (gap). Therefore Fat = 4 o)-4 GE 182 80808 —R—9)- Fy! G4! Gia? 5p? 8.9 Sai* = 89 Nav0a- adv) Jar ‘The first term in (3.5) arises because A(‘Y*"®,,_.g,¢) will not neces- sarily be zero even though Ay... z is, since the variation in the metric will induce a variation in the components I4,, of the connection. As the difference between two connections transforms like a tensor, AT™,, may be regarded as the components of a tensor. They are related to the variation in the components of the metric by AT = 49%4{(Agap); 0+ (Aga); 0 — (Adre); a} (The easiest way to derive this formula is to note that since itis a tensor relation, it must be valid in any coordinate system. In particular, one could choose normal coordinates about a point p. For these coordinates the components I'4,, and the coordinate derivatives of the components a» Vanish at p. The formula given can then be verified to hold at p.) Using this relation, AVi**,,..¢,, may be expressed in terms of (49».);¢ and the usual integration by parts employed to give an inte- grand involving Ag,, only. Thus we may write aI/2u as ‘Thus = 4g" dv. 3 i (T#Ag,,) dv, where T are the components of a symmetric tensor which is taken to be the energy-momentum tensor of the fields. (See Rosenfeld (1940) 3.3) LAGRANGIAN FORMULATION 67 for the relation between this tensor and the so-called canonical energy- momentum tensor.) "This energy-momentum tensor satisfies the conservation equations as a consequence of the field equations obeyed by the Yq... For suppose one has a diffeomorphism ¢:.@—>.# which is the identity everywhere except in the interior of 2. Then, by the invariance of integrals under a differential map, 1 1 1 i 1 [te fag Thus ale (Un - $*(Lm) = If the diffeomorphism ¢ is generated by a vector field X (non-zero only in the interior of 9) it follows that Bf, ten But aft - 2 Jor (ware) x Lg Fite ad +5 [POL 0a0 a0. 23a ‘The first term vanishes as a consequence of the field equations. In the second term, Lxgqy = 2Xiq; 9 Thus fC Fa90) a0 = 2f Xe); .— Te »Xe) ao. fe 4 The first contribution may be transformed into an integral over the boundary of @ which vanishes as X is zero there. Since the second term must therefore be zero for arbitrary X, it follows that T*>,, = 0. ‘We shall now give as examples Lagrangians for some fields which will be of interest later. Example 1: A scalar field y ‘This can represent, for example, the 7®-meson. The Lagrangian is 1m LH Wats 3 where m, % are constants. The Euler-Lagrange equations (3.4) are Via Be = 68 GENERAL RELATIVITY [3.3 ‘The energy-momentum tensor is Tay = Viehie— tes (Vioieot+ 3) 8.6) Example 2: The electromagnetic field This is described by a one-form A, called the potential, which is defined up to the addition of a gradient of a scalar function. The Lagrangian is L pa, L =~ 767 Fu Keag“9™, where the electromagnetic field tensor F is defined as 2A, i.e. Fe = 2A; a Varying Ag, the Euler-Lagrange equations (3.4) are Fan:e9% = 0. This and F,»,4 = 0 (which is the equation dF = d(dA) = 0) are the Maxwell equations for the source-free electromagnetic field. The energy-momentum tensor is Tay = Ga Fas Poa 9 — We Fey Fes g*9". (3.7) Example 3: A charged scalar field This is really a combination of two real scalar fields yr, and yy. These are combined into a complex scalar field y = y,+iyr,, which could represent, for example, 7+ and 7~ mesons. The total Lagrangian of the scalar field and electromagnetic field is Lm atic oF nied — 4 where ¢ is a constant and jf is the complex conjugate of yr. Varying ¥, ¥ and A, independently, one obtains Viad E+ ied oy, y-+ieds H) tied, ,gY = 0, and its complex conjugate, and i lan iWF «ie +Vote .W) = 0. The energy-momentum tensor is Tan = aV0+ Fab: +H-V, aie P +H, pied gy + Fach tried D+ ge PacPragt + OA, AWE + Lguy: 3.3) LAGRANGIAN FORMULATION 69 Example 4: An isentropic perfect fluid The technique here is rather different. The fluid is described by a fanetion p, called the density, and a congruence of timelike curves, called the flow lines. By a congruence of curves, is meant a family of curves, one through each point of .. If 9 isa sufficiently small com- pact region, one can represent a congruence by a diffeomorphism y: [a,b] x V+ @ where [a,b] is some closed interval of Rt and / is some three-dimensional manifold with boundary. The curves are said to be timelike if their tangent vector W = (8/2), t¢ [a,b], is timelike everywhere. The tangent vector V is defined by V = (—g(W, W))# W, so g(V, V) = —1, and the fluid current vector is defined by j = pV. It is required that this is conserved, i.e. j*,, = 0. The behaviour of the fluid is determined by prescribing the elastic potential (or internal energy) € as a function of p. The Lagrangian is taken to be E=—p(ite) and the action I is required to be stationary when the flow lines are varied and p is adjusted to keep j# conserved. A variation of the flow lines is a differentiable map y: (— 6,6) x [a,b] x W-> @ such that 0, [a,b], 4) = y((a,b], 4) and y(u, [2,5], 4%) = y((a,b],4%) on @-B, (ue(—8,8)). ‘Then it follows that AW = Lx W where the vector K is K = (2/2u),- This vector may be thought of as representing the displacement, under the variation, of a point of the flow line. It follows that AVe = Va, K?— Ka, V0— VaVeK,,. Ve. Using the fact that A(j2,.) = 0 = (Aj), q, one has (Ap), a V4 + ApV*, a +P; gAV4+ PAV"); q = 0. Substituting for AV and integrating along the flow lines, one finds Sp = (pR?),, + pKy;.V°V". ‘Therefore the variation of the action integral is J, (Ox tek 79 ( a hae. Integrating by parts, a= L182) 0 (t+ vere) Ka} av, 70 GENERAL RELATIVITY [3.3 where V* = V*,, V°. If this is zero for all K, it follows that (n+p) V4 = —p, (9+ VV), where #2 = p(1-+e) is the energy density and p = pY(de/dp) is the pressure. Thus V*, the acceleration of the flow lines, is given by the pressure gradient orthogonal to the flow lines. ‘To obtain the energy-momentum tensor one varies the metric. The calculations may be simplified by noting that the conservation of the current may be expressed as Osa Togas Wo Given the flow lines, the conservation equations determine j* uniquely at each point on a flow line in terms of its initial value at some given point on the same flow line. Therefore (/—g)j* is unchanged when the metric is varied. But P= IM —95tW — 99°) Gav» bl 2pAp = (G25? — FFI) AGavs and thus Tab foc +e rere] Vay? +p co = (utp) Ve + pg. (3.8) We shall call any matter whose energy-momentum tensor is of the above form (whether or not it is derived from a Lagrangian) a perfect fluid. From the energy and momentum conservation equations (3.1) applied to (3.8) one finds HaVe+(utp) Va,, = 0, (tp) Va + (gt + V2V?)p,, = 0. (3.10) ‘These are the same as the equations derived from the Lagrangian. We shall call a perfect fiuid isentropic if the pressure p is a function of the energy density only. In this case one can introduce a conserved density p and an internal energy ¢ and derive the equations and the energy-momentum tensor from a Lagrangian. One may also give the fluid a conserved electric charge e (i.e. J*, = 0 where J = e V is the electric current), The Lagrangian for the charged fluid and the electromagnetic field is (3.9) Lm Fh Fy Rag "Opi be) Bed 3.3] LAGRANGIAN FORMULATION crt ‘The last term gives the interaction between the fluid and the field. Then varying A, the flow lines and the metric respectively, one finds Feb, , = AJ, — P(g + VeV")+F%,J?, (utp) Ve Te = utp) VOV+ py + Ft, Fo tg" Rag FP). 3.4 The field equations So far, the metric g has not been specified. In the Special Theory of Relativity, which does not include gravitational effects, it is taken to be flat. One might think that one could include gravitation by keeping the metric flat and by introducing an extra field on space-time. How- ever, experiments have shown that light rays travelling near the sun are deflected. Since light rays are null geodesics, this shows that the space-time metric cannot be flat or even conformal to a flat metric. One therefore has to give some prescription for the curvature of space-time. It turns out that this prescription can be chosen s0 as to reproduce the results of Newtonian gravitation theory in the limit of small slowly varying curvature. It is therefore not necessary to intro- duce an extra field to describe gravitation. This is not to say that there could not be an additional field that produced part of the gravitational effects. Such a scalar field has been suggested by Jordan (1955), and Brans and Dicke (see Dicke (1964)). However, as mentioned before, such an additional field could be regarded as simply another matter field and included in the total energy-momentum tensor. We therefore adopt the view that the gravitational field is represented by the space-time metric itself. The problem then becomes one of finding field equations to relate the metric to the distribution of matter. ‘These equations should be tensor equations involving the matter only through its energy-momentum tensor, i.e. should not distinguish between two different matter fields which have the saine distribution of energy and momentum. This can be regarded as a generalization of the Newtonian principle that the active grevitational mass of a body (the mass producing a gravitational field) is equal to the passive gravi- tational mass (the mass acted on by the gravitational field). This has been verified experimentally by Kreuzer (1968). ‘To determine what the field equations should be, we shall consider the Newtonian limit. Since the Newtonian gravitational field equation does not involve time, the correspondence with Newtonian theory 2 GENERAL RELATIVITY [34 should be made in a metric which is static. By a static metric is meant a metric which admits a timelike Killing vector field K which is orthogonal to a family of spacelike surfaces. These surfaces may be regarded as surfaces of constant time and may be labelled by the parameter f. We define the unit timelike vector V as f-"K, where f?=—K*K,. Then V4,,=—Veh, where V4 = V2,,V?=f-¥f,,g% represents the departure from geodesity of the integral curves of V (which are of course also integral curves of K). Note that PV, = 0. These integral curves define the static frame of reference, that is to say, the space-time metric seems to be independent of time to a particle whose history is one of these curves. A particle released from rest and following a geodesic would appear to have an initial accelera- tion of —V with respect to the static frame. If f differs only slightly from unity the initial acceleration of a freely moving particle released from rest is approximately minus the gradient of f. This suggests that one should regard f—1 as the quantity analogous to the Newtonian gravitational potential. One can derive an equation for this potential by considering the divergence of Pa: Vag = (Va Va = Vina V+ VA,y Vg Rap V2V+ (V9, g),p V+ (KB)? = Ray V2V?.. But Wa Foam Loh TPT aa and ee ee ca 80 one finds Feanlg? + V°V?) = fRay VV, The term on the left is the Laplacian of f with respect to the induced metric in the three-surface {t = constant}. If the metric is almost flat, this will correspond to the Newtonian Laplacian of the potential. One would therefore obtain agreement with Newtonian theory in the limit of a weak field (i.e. when f ~ 1) if the term on the right is equal to 47@ times the matter density plus terms which are small in the weak field limit. This will be the case if there is a relation of the form Rav = Kay (3.11) where Ky, is a tensorial function of the energy-momentum tensor and the metric, which is such that (47@)-1K,, V4V° is equal to the matter density plus terms which are small in the Newtonian limit. We shall for the moment assume a relation of this form. 3.4] THE FIBLD EQUATIONS 73 Since Rp satisfies the contracted Bianchi identities R,°, = ER, (8.11) implies Ky (8.12) ‘This shows that the apparently natural equation Ky, = 47@Ty, cannot be correct, since (3.12) and the conservation equations T,’,, = 0 would imply 7, = 0. For a perfect fluid, for example, this would mean that 4 — 3p was constant throughout space-time, which is clearly not satisfied by a general fluid. In fact in general, the only first order identities satisfied by the energy-momentum tensor are the conservation equations. From this it follows that the only tensorial function Ky, of the energy-momentum tensor and the metric which obeys the identities (3.12) for all energy— momentum tensors, is Ke Ki = &(Lan~ 3T9av) + ADav» (3.13) where « and A are constants. The values of these constants can be determined from the Newtonian limit. Consider a perfect fluid with energy density jz and pressure p whose flow lines are the integral curves of the Killing vector (i.c. the fluid is at rest in the static frame). The energy-momentum tensor is given by (3.8). Putting this in (3.13) and (3.11), one finds Fr anlge + VOV") = f(4e(u + 3p) — A). (8.14) In the Newtonian limit the pressure p is normally very small compared to the energy density . (We are using units in which the speed of light is unity. In units in which the speed of light is c, the expression + 8p should be replaced by + 3p/ct.) One would therefore obtain approximate agreement with Newtonian theory if « = 87@ and if |A| is very small. We shall use units of mass in which @ = 1. In these units, a mass of 10% gm corresponds to a length of 1m. Sandage’s (1961, 1968) observations of distant galaxies place limits on |A| of the order of 10-8 em; we shall normally take A to be zero, but shall bear in mind the possibility of other values. Onemay then integrate (3.14) over a compact region F of the three- surface {t = constant} and transform the left-hand side into an integral of the gradient of f over the bounding two-surface @F : f_fanues spn ae = f_f,anlo +040) do - - =f _filor + eV Vary re

Вам также может понравиться