Author: Sharpe R.W.  

Tags: mathematics   geometry  

ISBN: 0-387-94732-9

Year: 1997

Text
                    R.W. Sharpe
Differential Geometry
Cartan's Generalization
of Klein's Erlangen Program
Foreword by S.S. Chern
With 104 Illustrations
Springer


R.W. Sharpe Department of Mathematics University of Toronto Toronto, M5S 1A1 Canada Editorial Board S. Axler Department of Mathematics Michigan State University East Lansing, MI 48824 USA F.W. Gehring Department of Mathematics University of Michigan Ann Arbor, MI 48109 USA P.R. Halmos Department of Mathematics Santa Clara University Santa Clara, CA 95053 USA Mathematics Subject Classification (1991): 53-01, 53A40, 53A20, 53B21 Library of Congress Cataloging-in-Publication Data Sharpe, R.W. (Richard W.) Differential geometry : Cartan's generalization of Klein's Erlangen program / by R.W. Sharpe. p. cm. - (Graduate texts in mathematics ; 166) Includes bibliographic references and index. ISBN 0-387-94732-9 (hard : alk. paper 1. Geometry, Differential. I. Title. II. Series. QA641.S437 1996 516.3'6-dc20 96-13757 Printed on acid-free paper. © 1997 R.W. Sharpe All rights reserved. This work may not be translated or copied in whole or in part without the written permission of the publisher (Springer-Verlag New York, Inc., 175 Fifth Avenue, New York, NY 10010, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in connection with any form of information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed is forbidden. The use of general descriptive names, trade names, trademarks, etc., in this publication, even if the former are not especially identified, is not to be taken as a sign that such names, as understood by the Trade Marks and Merchandise Marks Act, may accordingly be used freely by anyone. Production managed by Robert Wexler; manufacturing supervised by Jacqui Ashri. Photocomposed pages prepared by Sirovich, Warren, R.I. in T^X using Springer's svsing macro. Printed and bound by R.R Donnelley and Sons, Harrisonburg, VA. Printed in the United States of America. 987654321 ISBN 0-387-94732-9 Springer-Verlag New York Berlin Heidelberg SPIN 10534001
Foreword I am honored by Professor Sharpe's request to write a forward to his beautiful book. In his preface he asks the innocent question, "Why is differential geometry the study of a connection on a principal bundle?" The answer is of course very simple; because Euclidean geometry studies a connection on a principal bundle, and all geometries are in a sense generalizations of Euclidean geometry. In fact, let En be the Euclidean space of η dimensions. We call an or- thonormal frame x, e\,..., en (n-hl vectors), where χ is the position vector and ei have the scalar products (ei,ej) = Sij, 1 < г J < n. Then the space of all orthonormal frames is a principal fiber bundle with group 0(n) and base space En, the projection being defined by mapping x, ei,..., en to x. The equations l<7<n define the Maurer-Cartan forms ω^, with uJij + Uji = 0, 1 < г, j < n. They satisfy the Maurer-Cartan equations 1<к<п
vi Foreword This is Euclidean geometry by moving frames. The ω^ define the parallelism or connection. The Maurer-Cartan equations say that the connection is flat. This formulation has a great generalization. As in all disciplines, the development of differential geometry is tortuous. The basic notion is that of a manifold. This is a space whose coordinates are defined up to some transformation and have no intrinsic meaning. The notion is original, bold, and powerful. Naturally, it took some time for the concept to be absorbed and the technology to be developed. For example, the great mathematician Jacques Hadamard "felt insuperable difficulty ... in mastering more than a rather elementary and superficial knowledge of the theory of Lie groups," a notion based on that of a manifold [1]. Also, it took Einstein seven years to pass from his special relativity in 1908 to his general relativity in 1915. He explained the long delay in the following words: "Why were another seven years required for the construction of the general theory of relativity? The main reason lies in the fact that it is not so easy to free oneself from the idea that coordinates must have an immediate metrical meaning." [2] On the technology side the breakthrough was achieved by the tensor analysis of Ricci calculus. The central theme was Riemannian geometry, which Riemann formulated in 1854. Its fundamental problem is the "form problem": To decide when two Riemannian metrics differ by a change in coordinates. This problem was solved by E. Christoffel and R. Lipschitz in 1870. Christoffel's solution introduces a covariant differentiation, which could be given an elegant geometrical setting through the parallelism of Levi-Civita. Tensor analysis is extremely effective and has dominated differential geometry for a century. Another technical tool, which has not quite received the recognition it deserves, is the exterior differential calculus of Elie Cartan. This was introduced by Cartan in 1922, following the work of Frobenius and Darboux. All the exterior differential forms on a manifold form a ring. It depends only on the differentiable structure of the manifold and not on any additional structure such as a Riemannian metric or an affine connection. Topolog- ically it leads to the de Rham theory. Less known is its effectiveness in treating local problems. A fundamental question is the equivalence problem for G-structures: Given, on an η-dimensional manifold with coordinates иг, a set of linear differential forms u/, a similar set a;*·7 with coordinates u*i, and a subgroup G С G/(n,R), determine the conditions under which there exist functions u*j =u*j(u1,...,un), 1 <i,j <n, such that after substitution the a;*·7 differ from the a;·7 by a transformation of G. The form problem in Riemannian geometry is the case G = 0(n). The solution of the form problem by Cartan's method of equivalence leads automatically to the tensor analysis. Thus, the method of equivalence is more general. In the case G = 0(n), this leads to the Levi-Civita
Foreword vii parallelism and the Riemannian geometry. In this way Euclidean geometry generalizes to Riemannian geometry. For a general G, the solution of the equivalence problem is not always easy (cf. the Preface), although it is proved that it can always be achieved in a finite number of steps. Philosophically nice problems have nice answers. Klein geometry can be developed through the Maurer-Cartan equations. The generalization of the above discussion, from 0(n) to G, gives Cartan's generalized spaces, essentially a connection in a principal bundle. A fundamental problem is the relation of the local geometry with the global properties of the spaces in question. Such a result is the so-called Chern-Weil theorem that the characteristic classes can be represented by differential forms constructed explicitly from the curvature. The simplest result is the Gauss-Bonnet formula. I wish to take this occasion to mention some recent developments on Finsler geometry [3]. This is the geometry of a very simple integral and was discussed in problem 23 of Hilbert's Paris address in 1900. By a proper interpretation of the analytical results, Finsler geometry now assumes a very simple form showing it to be a family of geometries quite analogous to the Riemannian case. Differential geometry offers an open vista of manifolds with structures, finite or infinite dimensional. There are also simple and difficult low-dimensional problems, of the garden variety. If one switches between the two, life is indeed very enjoyable. It is a great mystery that the infinitesimal calculus is a source of such depth and beauty. References [1] J. Hadamard, Psychology of Invention in the Mathematical Field, Princeton University Press, Princeton, 1945, p. 115. [2] A. Einstein, Autobiographical Notes, in Albert Einstein: Philosopher Scientists, p. 67, 2nd ed., 1949, in vol. 7 of the series The Library of Living Philosophers, Evanston, IL, edited by P.A. Schilpp. [3] D. Bao and S.S. Chern, On a notable connection in Finsler geometry, Houston J. Math. 19 (1993), 135-180. S.S. Chern Mathematical Sciences Research Institute Berkeley, CA 94720, USA
Preface This book is a study of an aspect of Elie Cartan's contribution to the question "What is geometry?" In the last century two great generalizations of Euclidean geometry appeared. The first was the discovery of the non-Euclidean geometries. These were organized into a coherent whole by Felix Klein, who recognized them as various examples of coset spaces G/H of Lie groups. In this book we refer to these latter as Klein geometries. The second generalization was Georg Riemann's discovery of what we now call Riemannian geometry. These two theories seemed largely incompatible with one other.1 In the early 1920s Elie Cartan, one of the pioneers of the theory of Lie groups, found that it was possible to obtain a common generalization of these theories, which he called espaces generalizes and we call Cartan geometries (see diagram). Euclidean generalization Klein Geometry Geometries generali- generalization zation Riemannian generalization Cartan Geometry * Geometries 1The only relationship was the "accident" that some of the non-Euclidean geometries could be regarded as special cases of Riemannian geometries.
x Preface Looking at this diagram vertically, we can say that just as a Riemannian geometry may be regarded, locally, as modeled on Euclidean space but made "lumpy" by the introduction of a curvature, so a Cartan geometry may be regarded, locally, as modeled on one of the Klein geometries but made "lumpy" by the introduction of curvature appropriate to the model in question. Looking at the same diagram horizontally, a Cartan geometry may be regarded as a non-Euclidean analog of Riemannian geometry. Cartan actually gave the first example of a Cartan geometry more than a decade earlier, in the remarkable tour de force [E. Cartan, 1910]. In that paper he considered the case of a two-dimensional distribution on a five- dimensional manifold. He showed that such a distribution determined, and was determined by, a Cartan geometry modeled on the homogeneous space G2/H, where Я is a certain nine-dimensional subgroup of the fourteen- dimensional exceptional Lie group G2· This process of associating a Cartan geometry to a raw geometric entity (the distribution) is an example of "solving the equivalence problem" for the entity in question. Although the solution of an equivalence problem is not always a Cartan geometry, in many important cases it is. When it is, the invariants of the geometry (curvature, etc.) are a priori invariants of the raw geometric entity. We recommend [R.B. Gardner, 1989] for an account of the method of equivalence. To be a little more precise, a Cartan geometry on Μ consists of a pair (Ρ,ω), where Ρ is a principal bundle Η —> Ρ —> Μ and ω, the Cartan connection, is a differential form on P. The bundle generalizes the bundle Η —> G —> G/H associated to the Klein setting, and the form ω generalizes the Maurer-Cartan form uq on the Lie group G. In fact, the curvature of the Cartan geometry, defined as dw + \{ω,ω\, is the complete local obstruction to Ρ being a Lie group. One reason for the power of Cartan's method comes from the fact that these new geometries maintain the same intimate relation with Lie groups that one sees in the case of homogeneous spaces. This means, for example, that constructions in the theory of homogeneous spaces often generalize in a simple manner to the general "curved" case of Cartan geometries. It also means that the differential forms that appear are always related to components of the Maurer-Cartan form of the Lie group, a context in which their significance remains clear. In the particular case of a Riemannian manifold M, Cartan's point of view offered a new and profound vantage point that is largely responsible for the modern insistence on "doing differential geometry on the bundle Ρ of orthonormal frames over M." The history of the study of Cartan geometries is somewhat troubled. First is the difficulty Cartan faced in trying to express notions for which there was no truly suitable language.2 Next is the widely noted difficulty in reading 2This difficulty was resolved with the introduction of the notion of a principal bundle and of vector-valued forms on such a bundle.
Preface xi Cartan.3 In his paper [C. Ehresmann, 1950] Charles Ehresmann gave for the first time a rigorous global definition of a Cartan connection as a special case of a more general notion now called an Ehresmann connection (or more simply, a connection). For various reasons4 the Ehresmann definition was taken as the definitive one, and Cartan's original notion went into a more or less total eclipse for a long time. The beautiful geometrical origin and insight connected with Cartan's view were, for many, simply lost. In short, although the Ehresmann definition gives us a good notion, it hides the real story about why it is so good. In this connection, the following quotation is interesting [S.S. Chern, 1979]: The physicist C.N. Yang wrote [C.N. Yang, 1977]: "That non- abelian gauge fields are conceptually identical to ideas in the beautiful theory of fibre bundles, developed by mathematicians without reference to the physical world, was a great marvel to me." In 1975 he mentioned to me: "This is both thrilling and puzzling, since you mathematicians dreamed up these concepts out of nowhere." Far from arising "out of nowhere," the simple and compelling geometric origin of a connection on a principal bundle is that it is a generalization of the Maurer-Cartan form. Moreover, a study of the Cartan connection itself can illuminate and unify many aspects of differential geometry. Novelties Aside from the fact that one cannot find a fully developed, modern exposition of Cartan connections elsewhere, what is new or different in this book? New Treatment This book is written at a level that can be understood by a first- or second- year graduate student. In particular, we include the relevant theory of manifolds, distributions and Lie groups. For us, a manifold is, by definition, a 3To paraphrase Robert Bryant, 'You read the introduction to a paper of Cartan and you understand nothing. Then you read the rest of the paper and still you understand nothing. Then you go back and read the introduction again and there begins to be the faint glimmer of something very interesting." 4At that stage it was easier to read Ehresmann than Cartan. There was also the attraction of a more general and global notion.
xii Preface locally Euclidean, paracompact Hausdorff space. This is the same as a locally Euclidean Hausdorff space each of whose components has a countable basis.5 In particular, Lie groups are defined to be manifolds in this sense. The result of Yamabe and Kuranishi ([H. Yamabe, 1950]) that a connected subgroup of a Lie group is a Lie subgroup implies that any subgroup of a Lie group is a Lie group in the present sense. The discussion of subman- ifolds given in Chapter 1 is broad enough to include these subgroups as submanifolds. In our coverage of bundle theory, we emphasize the abstract principal bundles rather than bundles of frames.6 Of course, these two views are really equivalent. In the case of the "first-order" geometries, the equivalence is quite simple. However, in the case of "higher-order" geometries, the choice of the higher-order frames usually seems to be decided on a rather ad hoc basis and can be complicated. Here the bundle approach gives a real advantage, and the right choice of frames becomes clear (if needed) once the bundle is understood. Another important advantage of working with the bundles themselves is that they give a common language, facilitating comparison between geometries and emphasizing the relation to the model space. In this sense, comparing Cartan geometries is like comparing Klein geometries. Chapter 3 contains a complete and economical development of the Lie group—Lie algebra correspondence based on the fundamental theorem of non-abelian calculus. One of the novelties here is the characterization of a Lie group as a manifold equipped with a Lie algebra-valued form on it satisfying certain properties. This characterization prepares the reader for the generalization to Cartan geometries in Chapter 5. Finally, in Appendix В we explain how one manifold may roll without slipping or twisting on another in Euclidean space. We also show how this notion yields a differential system that contains both the Levi-Civita connection and the Ehresmann connection on the normal bundle for a sub- manifold of Euclidean space. New Results Let us move on to some results we believe are new. In Chapter 4 we introduce the fundamental property of Klein geometries characterizing the kernel of such a geometry. This result is used in Chapter 5 in an essential way to show the equivalence of the base and bundle definitions of Cartan geometries in the effective case. In Chapter 5 we introduce and classify Cartan space forms. These geometries generalize the classical Riemannian 5The usual definition requires a manifold to have a countable basis (cf., e.g., [Boothby, W. 1986, p. 6]). 6 In much the same way, one might emphasize an abstract Lie group rather than a matrix group realizing it.
Preface xiii space forms.7 One important ingredient of this classification is the property (apparently new) of a Cartan geometry called "geometric orient ability." Another is the notion of "model mutation." Finally, in Chapter 7 we give a classification of the submanifolds of a Mobius geometry. This classification is more general than that of [A. Fialkow, 1944] in that ours allows the presence of umbilic points. Prerequisites and Conventions This book assumes very few prerequisites. The reader needs to be familiar with some basic ideas of group theory, including the notion of a group acting on a set. Results from the calculus of several variables, point set topology, and the theory of covering spaces are used in various places, and the long, exact sequence of homotopy theory is used once (at the end of Chapter 5). Aside from this, most of the material is developed ab initio. However, the reader is invited to shoulder some of the burden of the work in that essential use is made of a few of the exercises. These exercises are denoted by an asterisk to the right of the exercise number. The numbering follows a single sequence throughout the book, with all items (definitions, theorems, figures, etc.) in a single stream. Thus 4.3.2 refers to Chapter 4, Section 3, item 2. For references to items occurring in the same chapter, we omit the chapter number, so that in Chapter 4, 4.3.2 becomes 3.2. We use the following dictionary of symbols to denote the ends of various items: symbol end of Φ definition □ exercise ■ proof ♦ example Although it will often be convenient for us to write column vectors as row vectors, the reader should remember that all vectors are in fact column vectors. 7In fact, this notion is general enough to immediately allow a description of general symmetric spaces.
xiv Preface Limitations The reader will find no mention here of some basic topics in differential geometry, such as Stokes' theorem, characteristic classes, and complex geometries. Also, our approach to Lie theory is "elementary" in that we do not discuss or use the classification theory of Lie groups, with its attendant study of roots, weights, and representations. Originally, we had wished to include more than the three examples of Cart an geometries studied here; but in the end, the pressures of time, space, and energy limited this impulse. The three geometries we do study are not developed in complete analogy to each other. For example, the discussion of immersed curves in a Mobius geometry in terms of the normal forms given in Chapter 7 does of course have a Riemannian analog, but that is not studied in this book. And one may study subgeometries of projective geometries just as one studies subgeometries of Riemannian and Mobius geometries, but we do not do so here. We have also resisted the impulse to make a "dictionary" translating among the various versions of Cartan's view, Ehresmann's view,8 and the view expressed in [L.P. Eisenhart, 1964]. In the end, however, for those who are interested in it, it should be abundantly clear how Cartan's view does illuminate the others. Some Personal Remarks An author often writes a book in order to sort out his or her own understanding of the subject. This is the circumstance in the present case. When I was an undergraduate, differential geometry appeared to me to be a study of curvatures of curves and surfaces in R3. As a graduate student I learned that it is the study of a connection on a principal bundle. I wondered what had become of the curves and surfaces, and I studied topology instead. The reawakening of my interest in this subject began in 1987 when Tom Willmore very kindly wrote me a note thanking me for a preprint and mentioning his great interest in what is known as the Willmore conjecture (cf. 7.6). This led me once again to look at principal bundles and connections. In particular, I wondered whether there was an intrinsically defined Ehres- mann connection on a surface in 53 that was invariant under the group of Mobius transformations of 53. It turns out there is no such connection. However, after calculating normal forms for surfaces in the Mobius sphere 53 (cf. [G. Cairns, R. Sharpe, and L. Webb, 1994]), it became clear to me that there must be some other kind of invariantly defined structure inherited on the surface from its embedding in 53. (In Chapter 7 it is shown See, however, the discussion in Appendix A dealing with the relationship between Cart an and Ehresmann connections.
Preface x ν that a Cartan connection is defined in this situation, and, in fact, Cartan also knew this [E. Cartan, 1923].) During this time it began to seem strange to me that Ehresmann connections play such a prominent role in modern differential geometry. In some cases, such as the Levi-Civita connection, the connection is determined by the geometry. In many cases, however, one makes use of an arbitrary connection that one proves to exist by a general technique. This is the appropriate point of view for the construction of the characteristic classes of Chern and Pontryagin. There one may use any connection, since the aim is to obtain topological invariants for which the particular choice of connection does not matter. But these considerations seem to be at their base topological rather than differential geometric. My innocent question, left over from my undergraduate days, was "Why is differential geometry the study of a connection on a principal bundle?" And I began, rather impertinently, to ask this question at every opportunity, usually picking on some unsuspecting differential geometer who did not know me very well. During one of these sessions, Min Oo remarked that Elie Cartan had considered connections with values in a Lie algebra larger than that of the fiber.9 Later I read, and translated,10 Cartan's book [E. Cartan, 1935]. I browsed through Cartan's collected works and through those of his successors and interpreters. It became clear to me that Cartan had a subtle and really wonderful idea, which gives a fully satisfying explanation for the modern, and approximately true, notion that differential geometry is the study of an Ehresmann connection on a principal bundle. There seems to be no treatment of these things in the standard texts on differential geometry. In the few books where the Cartan connections are mentioned at all (e.g., [J. Dieudonne 1974], [W.A. Poor, 1981], and [M. Spivak, 1979]), they make only a brief appearance, perhaps in the exercises or toward the end of the book, and one is left with the impression that the notion is only a quaint curiosity left over from bygone days. Six years ago I began to scribble some notes about these things and to talk about them; after a number of months had passed, I realized I was writing a book on the subject. * * * I would like to thank everyone who has had an influence on this book. In addition to those mentioned above, I am grateful to Bernard Kamte, Joe Repka, Qunfeng Yang, and my wife, Mary, for their comments on portions of the manuscript. I would also like to acknowledge my gratitude to 9See [E. Ruh, 1993] for a brief recent overview of Cartan connections and some of their applications. 10A copy of my translation, which is only a rough draft, can be found in the Mathematics Library at the University of Toronto.
xvi Preface the National Science and Engineering Research Council of Canada for its support of this project through grant #OGP0004621. Finally, I would like to thank Velamir Jurdjevic for his encouragement over the years. It was Vel who suggested that, although it is perhaps impossible to catch all the errors before a book reaches print, the principal demand is that a book be interesting. As for the first part of his remark, the responsibility for any remaining errors lies with me. I will leave it to the reader to judge whether or not this principal demand has been met. Richard Sharpe Toronto, Canada October 16, 1995
Contents Foreword ν Preface ix 1 In the Ashes of the Ether: Differential Topology 1 §1. Smooth Manifolds 3 §2. Submanifolds 17 §3. Fiber Bundles 28 §4. Tangent Vectors, Bundles, and Fields 39 §5. Differential Forms 52 2 Looking for the Forest in the Leaves: Foliations 65 §1. Integral Curves 66 §2. Distributions 74 §3. Integrability Conditions 75 §4. The Frobenius Theorem 76 §5. The Frobenius Theorem in Terms of Differential Forms . . 79 §6. Foliations 81 §7. Leaf Holonomy 83 §8. Simple Foliations 90 3 The Fundamental Theorem of Calculus 95 §1. The Maurer-Cartan Form 96 §2. Lie Algebras 101
xviii Contents §3. Structural Equation 108 §4. Adjoint Action 110 §5. The Darboux Derivative 115 §6. The Fundamental Theorem: Local Version 116 §7. The Fundamental Theorem: Global Version 118 §8. Monodromy and Completeness 128 4 Shapes Fantastic: Klein Geometries 137 §1. Examples of Planar Klein Geometries 140 §2. Principal Bundles: Characterization and Reduction .... 144 §3. Klein Geometries 150 §4. A Fundamental Property 160 §5. The Tangent Bundle of a Klein Geometry 162 §6. The Meteor Tracking Problem 164 §7. The Gauge View of Klein Geometries 166 5 Shapes High Fantastical: Cartan Geometries 171 §1. The Base Definition of Cartan Geometries 173 §2. The Principal Bundle Hidden in a Cartan Geometry . . . 178 §3. The Bundle Definition of a Cartan Geometry 184 §4. Development, Geometric Orientation, and Holonomy . . . 202 §5. Flat Cartan Geometries and Uniformization 211 §6. Cartan Space Forms 217 §7. Symmetric Spaces 225 6 Riemannian Geometry 227 §1. The Model Euclidean Space 228 §2. Euclidean and Riemannian Geometry 234 §3. The Equivalence Problem for Riemannian Metrics .... 237 §4. Riemannian Space Forms 241 §5. Subgeometry of a Riemannian Geometry 244 §6. Isoparametric Submanifolds 259 7 Mobius Geometry 265 §1. The Mobius and Weyl Models 267 §2. Mobius and Weyl Geometries 277 §3. Equivalence Problems for a Conformal Metric 284 §4. Submanifolds of Mobius Geometry 294 §5. Immersed Curves 318 §6. Immersed Surfaces 323 8 Projective Geometry 331 §1. The Projective Model 332 §2. Projective Cartan Geometries 338 §3. The Geometry of Geodesies 343
Contents xix §4. The Projective Connection in a Riemannian Geometry . . 349 §5. A Brief Tour of Projective Geometry 353 A Ehresmann Connections 357 §1. The Geometric Origin of Ehresmann Connections 358 §2. The Reductive Case 362 §3. Ehresmann Connections Generalize Cartan Connections . 365 §4. Covariant Derivative 371 В Rolling Without Slipping or Twisting 375 §1. Rolling Maps 376 §2. The Existence and Uniqueness of Rolling Maps 378 §3. Relation to Levi-Civita and Normal Connections 382 §4. Transitivity of Rolling Without Slipping or Twisting . . . 388 С Classification of One-Dimensional Effective Klein Pairs 391 §1. Classification of One-Dimensional Effective Klein Pairs . . 391 D Differential Operators Obtained from Symmetry 397 §1. Real Representations of S02(R) 397 §2. Operators on Riemannian Surfaces 400 Ε Characterization of Principal Bundles 407 Bibliography 411 Index 417
1 In the Ashes of the Ether: Differential Topology It must be agreed that "hypergeometry" seems to have been devised in order to strike the imagination of people who have not enough mathematical knowledge to be aware of the true character of an algebraic construction expressed in geometric terms, for that is what "hypergeometry" really is. -R. Guenon, 1945 Is Euclidean geometry true? It has no meaning. ... One geometry cannot be more true than another; it can only be more convenient. -H. Poincare, 1902 I attach special importance to the view of geometry which I have just set forth, because without it I should have been unable to formulate the theory of relativity. -A. Einstein, 1922 Although several mathematicians, especially C.F. Gauss, studied the notion of a smooth manifold in special cases, the idea of an abstract manifold of arbitrary finite dimension seems to be due to Riemann. Mathematicians were led to these notions only slowly. As the idea of a vector space of dimension higher than three became acceptable in the last century, algebraic geometers began to study the solutions of polynomial equations in many variables. For example, they studied the algebraic curves in the complex projective plane, which in the real sense is roughly the study of ordinary surfaces in four-dimensional space. At the same time, mathematical physicists interested themselves in six-dimensional space, the state space of a single particle with three position and three momentum variables; if TV
2 1. In the Ashes of the Ether: Differential Topology particles are considered, the dimension of the state space jumps to 67V. The seamless algebraic passage between the ordinary notions of space and their higher-dimensional analogs must have prepared the way for the acceptance of Riemann's abstract manifolds. But the willingness and even the necessity of regarding these developments as truly geometrical was slow to take hold and was fiercely resisted in some quarters.1 Riemann's work was followed with papers by Christoffel, Ricci, and Levi- Civita. Klein and Lie were interested in the study of Lie groups and their homogeneous spaces, which are again examples of manifolds, albeit very special ones. A real boost to the subject came with Einstein's discovery of general relativity in 1915. At this stage it became much clearer that the fourth dimension might be scientifically regarded as more than some geometrical fantasy. Einstein himself regarded the abstract four-manifold2 as what remains of the "ether" in general relativity (cf. [A. Einstein, 1922]). In the early years these spaces were thought about only locally, as pieces of four-dimensional Euclidean space. Later the notion of "cosmology" appeared (cf. [D. Howard and J. Stachel, eds., 1989]) and began to show the influence of the global topology. Perhaps we may say that in studying smooth manifolds we are studying the possible shapes of the ether. In the hierarchy of geometry (whose "spine" rises from homotopy theory through cell complexes, through topological and smooth manifolds to analytic varieties), the category of smooth manifolds and maps lies "halfway" between the global rigidity of the analytic category and the almost total flabbiness of the topological category. We might say that a smooth manifold possesses full infinitesimal rigidity governed by Taylor's theorem while at the same time having absolutely no rigidity relating points that are not "infinitesimally near" each other, as is seen by the existence of partitions of unity (cf. [W. Boothby, 1986], pp. 193-195). Smooth manifolds are sufficiently rigid to act as a support for the structures of differential geomtery while at the same time being sufficiently flexible to act as a model for many physical and mathematical circumstances that allow independent local perturbations. Perhaps the smooth "substance" may be regarded as a mathematical model for Aristotle's materia prima or the Hindu prakriti. In this chapter we show how the differential calculus lives on after the death of a preferred coordinate system. In particular, we discuss some of the beautiful and elementary constructions surrounding the notion of a smooth ^f. [M. Monastyrski, 1987], pp. 22 and 64. 2Einstein admits [A. Einstein, 1922] his difficulty in conceiving space-time without a metric. In fact, he spent the rest of his life searching for a geometric structure on a four-manifold supplementing the metric to include electromag- netism within a unified field theory generalizing his general theory of relativity. The general question of "what geometry on a manifold supports physics?" remains vital to the present day. On the other hand, the context for all such structures seems to continue to include a smooth manifold.
§1. Smooth Manifolds 3 manifold. Care is taken to show the relation between the forms of these constructions for the concrete situation of manifolds embedded in an ambient Euclidean space and those for an abstract, coordinate-independent, manifold. §1. Smooth Manifolds The definitive modern definition of a smooth manifold seems to have been given by Hassler Whitney ([H. Whitney, 1936]), in which a smooth manifold is presented as floppy pieces of Euclidean space glued together with a sort of differentiable glue. But in order to gain perspective, we start our more detailed discussion with topological manifolds. Topological Manifolds Definition 1.1. Let Μ be a paracompact3 Hausdorff space. We call Μ an n-dimensional topological manifold (and write it Mn if we wish to denote the dimension) if for each point ρ G Μ there is an open set U in Μ containing ρ such that U is homeomorphic to an open subset of Rn by some homeomorphism φ.4 Such a pair ([/, φ) is called a local coordinate system or (in the maritime terminology) a chart on M. On the other hand, φ~ι is called a local parameterization of M. Often, however, speaking loosely, both φ and φ~ι are referred to as coordinate systems. * We note that if Μ is an η-manifold, then so is every open subset of it; in particular, the components of Μ are also manifolds. On the other hand, manifolds are quite badly behaved with respect to quotients. Taking the quotient of a manifold by almost any equivalence relation on it leads out of the category of manifolds. Some notable exceptions to this are studied later under the name of fiber bundles. Example 1.2 (the η-sphere). Let Mn = Sn = {x G Rn+1 | χ · χ = 1}· Set U — {χ G Sn | xn+i > — 1}· Then U is an open subset of 5n, as is its homeomorphic image p(U), where p: Rn+1 —» Rn+1 is the reflection in the hyperplane xn+i = 0. The open sets U and p(U) together form a covering for Sn. We can see that U (and hence p(U)) is homeomorphic to Rn by considering the stereographic projection φ from the point — en+i given by 3This is equivalent to each component of Μ having a countable basis for its topology (cf. [J. Dugundiji, 1966], p. 241). 4 We note that only the topological structure of Rn is brought into play here. In particular, neither the Euclidean structure nor the affine space structure has any role. Nevertheless, we shall continue to refer to Rn as Euclidean space.
4 1. In the Ashes of the Ether: Differential Topology φ: U —► Rn sending (χι,...,χη+ι) ·-> This situation is pictured below. 1 1 +^n+l yX\, . . . , Xn)· Xn+\ ШЛ) > (хиХ2,...,Хп) Example 1.3 (Projective space). Let Pn(R) = the set of one-dimensional subspaces of Rn+1. The surjective map p:Rn+1 — {0} —► Pn(R), sending ν ь-> (υ) (= the line spanned by υ), induces the quotient topology on Pn(R). Let us show that the topological space Pn(R) is Hausdorff. Any two distinct points of Pn(R) may be represented by vectors v\,V2 £ Sn satisfying V\ · V2 > 0. Choose ε > 0 to be less than half the distance from i?i to V2- Then no line through the origin meets both Bs{v\) and Βε{ν2), where Βε(υ) denotes the open ball in Rn+1 of radius ε about v. It follows that (vi), (v2) e Pn(R) lie in disjoint, open sets ρ(Βε(νι)),ρ(Βε(ν2)) С Pn(R). If a: Rn —> Rn+1 is any affine map whose image does not contain the origin, then the composite map pa: Rn —» Pn(R) is injective and continuous. lines = points of projective space image of a Moreover, if V С Rn is open, then p-\pa(V)) = {\v e Rn+1 | λ e K\v e a(V)}
§1. Smooth Manifolds 5 is open in Rn+1. Thus, pa is a homeomorphism onto its image. It is called the affine parameterization, and its inverse is called the affine coordinate system for Pn(R) arising from the affine map a. Since every point of Pn(R) lies in the image of some of the affine parameterization, it follows that Pn(R) is a topological manifold. A similar construction may be made for an arbitrary vector space V, yielding the projective space P(V) canonically associated to V. ♦ Smooth Manifolds The following definition comes into sharper focus if the meanings of the maritime terminology of "charts" and "atlas" are given due consideration. Definition 1.4. If Mm is a topological manifold, then a (smooth) atlas on Μ is a collection Я — {(Ε/*, φι)} of charts such that (i) the UiS form an open covering of Μ, and (ii) for each pair of charts ([/, φ) and (V, ψ) in Я, the map φ = ψφ-ι Ι φ(υ η V): φ(ϋ ПУ)-> φ(ϋ Π V) is a smooth5 (i.e., C°°) map between open sets in Euclidean space (change of coordinates; see Figure 1.6). For brevity, we may speak of the chart φ, leaving U = domain((^) nameless. * Example 1.5. Continuing Example 1.2 of the η-sphere, we have the following atlas consisting of two charts, Я = {(ί/, φ), (p(U), φ ρ)}· The composite φ(ρφ~ι): Rn+1 — {0} —► Rn+1 — {0} is easily seen to be smooth, so Л is a smooth atlas. ♦ FIGURE 1.6. Change of coordinate system. Exercise 1.7. show that the affine coordinate systems described in Example 1.3 constitute an atlas for Pn(R). □ 5Here we use not only the topological structure of Rn but also its affine structure so that the notion of differentiation makes sense. The choice of origin and the Euclidean inner product are not involved.
6 1. In the Ashes of the Ether: Differential Topology The main point of introducing smooth atlases is to be able to unambiguously differentiate the composite appearing in Definition 1.4(ii) as often as we please. For this simple purpose the particular choice of atlas is not important, so we are led to make the definition that two atlases are equivalent if their union is also an atlas. This in turn leads to the following definition. Definition 1.8. A smooth structure on a topological manifold is an equivalence class of atlases, and a smooth manifold is a topological manifold with a specified smooth structure. * For example, the atlas for the η-sphere above endows it with a smooth structure called the canonical smooth structure. Starting with a given atlas, we can enlarge it to the union of all the atlases equivalent to it. This will give a unique maximal atlas equivalent to the given one, and the smooth structure may be identified with the corresponding maximal atlas. In particular, if φ is a chart in the maximal atlas of a smooth manifold, and / is a local diffeomorphism of Rn with image(^) С domain(/), then /φ is also a chart in the maximal atlas. This kind of procedure allows a great deal of flexibility in the choice of chart. For example, we can "tidy up" the chart φ with respect to a point ρ G U by following φ with a translation in Rn so that the new chart sends ρ to 0 G Rn. We may further follow φ by some linear transformation of Rn to move the various directions into convenient positions. We shall always implicitly assume we are dealing with a maximal atlas, even if it is described by a particular choice of a nonmaximal one. Orientability Definition 1.9. Let ([/, φ) and (V, ψ) be two charts for the smooth manifold Mn. We say these charts are compatible on W С U Π V if the change of coordinate mapping Φ = ψφ~ι \ φ{ϋ Π V) has positive Jacobian determinant at each point of tp(W). We say they are compatible if they are compatible on U Π V. * Remark 1.10. Since the sign of the determinant of the Jacobian matrix (ψ ο φ~ιγ(χ) is constant on each component of φ{ϋ Π V), it follows that if ({/, φ) and (V, ψ) are compatible at χ G U Π V, then they are compatible on the component of U Π V containing x. On the other hand, if they are not compatible at a point χ e U Π V, then the charts (υ,φ) and (ν,φψ) are compatible, where φ is any linear transformation of Rn of negative determinant. Definition 1.11. Let Μ be a smooth manifold. An atlas for Μ is oriented if any two charts in it are compatible. Μ is called topologically orientable
§1. Smooth Manifolds 7 if it has an oriented atlas. A maximal such atlas is called a topological onentation for Μ. Φ Given an oriented atlas for Μ, we can always enlarge it to obtain a unique maximal, oriented atlas containing it. Thus, an oriented atlas for Μ determines an orientation of M. Exercise 1.12. Show that a connected, orientable, smooth manifold Μ has exactly two orientations. □ The question of whether or not a given manifold is orientable can be determined by studying the loops on it. In preparation for this, we are going to study paths on Μ of the form σ: (7,0,1)-* (M,p,q). Definition 1.13. Let σ be a path on Mn. Suppose we are given a partition 0 = to < ti < ... < tk = 1 and a family of charts (Ui,tpi), 1 < г < к, such that σ([ίΐ-ι,^]) С Щ, 1 < г < к. We call these charts compatible along σ if {ϋ%,ψϊ) and ([/г+ь^г+ι) are compatible at σ(^) Е^П t/i+i for 1 < i < к- 1. * Lemma 1.14. For any path σ on Mn there exists a family of compatible charts along σ. Proof. Since I is compact and {a~l(U) | U a connected open set arising from a chart} is an open cover of 7, this covering has a Lebesgue number ε > 0 (cf. [J. Dugundji, 1966], p. 234). Choose an integer к > Ι/ε, and set ti — г/к, 0 < г < к, so that 0 = ίο < ^ι < ... < £fc = 1 is a partition of 7 and σ([^-ι,^]) С Ui, where Щ is the open set of some chart {υ%,ψι)· We may assume inductively that (Щ, φι) and (£/г+ь ^г+ι) are compatible at a(ti) for г < s. If (Us,tps) and (£/s+1,(^s+i) are compatible at a(ts), we have the induction step. Otherwise we replace φ8+ι by φφ8+\, where ф: Rn —» Rn is any linear map of negative determinant so that (Us, φ8) and ({7s+i,0(^s+i) are compatible at a(ts). Renaming φφδ+ι as φ8+ι completes the inductive step. ■ Lemma 1.15. Let σ: (7,0,1) —> (M,p,q) be a path on M, and let (C/»,^), 1 < г < к, and (Vj,^j), 1 < j < I, be two compatible families of charts along σ. Then {ϋ\,φι) is compatible with (Vi,^i) at ρ =$> (Uk,<Pk) ls compatible with (νι,ψι) at q.
8 1. In the Ashes of the Ether: Differential Topology Proof. Let 0 = so < si < ... < Sk = 1 and 0 = to < t\ < ... < ti = 1 be the partitions of I corresponding to the compatible families of charts (Ui,ipi), 1 < г < к, and (Vj,i/jj), 1 < j <l, respectively. Let Si — [si_i,Si] and Tj = [tj-\,tj]. The proof is by induction, where the inductive step is Suppose that Si PiTj ^ 0 and (Ui,ipi) is compatible with (Vj,^j) on a(SinTj). Then, either (i) Si+iHTj φ 0 and (υί+ι,ψί+ι) is compatible with (Vj,^) on σ(5ΐ+ι(Ί Tj), or (ii) Si HTj+i φ 0 and (Ui,ipi) is compatible with (Vj+i^j+i) on σ(5» Π Since (I7i,(^i) is compatible with (V\,^i) at p, it will follow inductively that (C/fc,(^fc) is compatible with (V/,^/) at g. Let us verify the inductive step. If Si < tj, then Si £ SiP\ S^+i Π Tj, so Si+i C\Tj φ$. Now (Ui,tpi) and ([/i+i,<£>i+i) are compatible at σ(βί), so it follows from the inductive hypothesis that (£/i+i,^i+i) is compatible with (Vj,^j) at a(si). Hence, by Remark 1.9, ([/г+ь^г+ι) is compatible with (Vj,ipj) on σ(5ί+ι HTj). This verifies condition (i). Similarly, in the case tj < Si, we get condition (ii). ■ Now we are ready for the notion of an orientation-preserving loop, which is the key to the question of the orientability of a manifold. Definition 1.16. Let λ: (1,0,1) —» (Μ,ρ,ρ) be a loop on M. We say that λ is orientation preserving if there is a compatible family of charts (Ui, ψί), 1 < i < k, along λ such that (ΙΙι,ψι) is compatible with (Uk^k) at p. * Note that, by Lemma 1.15, if a loop λ is orientation preserving, then any family of charts compatible along λ will have the property described in Definition 1.16. Now we are ready for our characterization of the orientability of manifolds in terms of loops. Proposition 1.17. Let Μ be a smooth manifold. Then Μ is orientable <=> every loop on Μ is orientation preserving. Proof. =>: Given a loop λ: (7,0,1) —» (Μ,ρ,ρ) on M, we may select a family of charts, (ΙΙΐ,φΐ), 1 < г < к, from an oriented atlas for Μ as in the proof of Lemma 1.14. These charts will automatically be compatible since the atlas is oriented. In particular, (U\,tpi) is compatible with (Uk^k) at p, so the loop λ is orientation preserving.
§1. Smooth Manifolds 9 4=: It suffices to consider the case when Μ is connected. Fix a point ρ G M, and choose a connected chart ([/, <p) with ρ Ε U. For every point # G M, choose a path σ: (/,0,1) —» (Μ,ρ, χ) joining ρ to χ, and choose a family of connected charts ({/», v?i), 1 < ^ < &, compatible along σ, starting with (ί/ь^О = (U,ip). Set (Ux,<px) = (Uk,<pk). Then {(Ε/*,^) |xGM} is an atlas for M. We claim that it is oriented. It suffices to show that any two charts (υχ,φχ) and (ϋν,φν) are compatible at every point ζ G UxC\Uy. Suppose that (Ε/*, <£>*), 1 < г < к, is the compatible family of charts used to obtain (υχ,φχ) and that (V^,^), 1 < г < I, is the compatible family of charts used to obtain (Uy,ipy). Since Ux and C/y are connected, we can join χ to ζ in С/я and у to ζ in Uy. These paths, together with the paths from ρ to χ and y, give a loop based at z. Since this loop is orientation preserving and the family of charts along it corresponding to the open sets Ux = Uk, · · ·, E/2, U\ = U — V\, V2, · · ·, Ц — Uy is compatible, it follows that (Ux,ipx) and (Uy,ipy) are compatible at ζ eUxnuy. ш Corollary 1.18. Let Μ be a connected, smooth manifold and let ρ G M. Then Μ is orientable <=> every loop on Μ based at ρ is orientation preserving. Proof. In the proof of the proposition, the only loops we needed were ones that passed through the fixed point p. Thus it suffices to show that if, for a given orientation-preserving loop, we change the point on it that we regard as the base point, then the new loop is still orientation preserving. But this is clear. ■
10 1. In the Ashes of the Ether: Differential Topology Examples of Smooth Manifolds The fundamental example of a smooth η-manifold is of course Rn itself with its atlas consisting of the identity map alone. More generally, any finite dimensional real vector space V carries a canonical smooth structure in the following manner. If dim(Vr) = n, we take the atlas consisting of all linear isomorphisms φ: V —► Rn. The collection of such maps is an atlas since for any two, φ and ψ, the change of coordinates is a linear map ψφ~ι\~Κη —► Rn and hence smooth. Although we do not prove it here, it turns out that the canonical structure on V is, up to diffeomorphism (see Definition 1.20 ahead), the unique smooth structure except in the case with dim(V) = 4. In the latter case it turns out that there are infinitely many smooth structures on V. (Cf. [D. Freed and K. Uhlenbeck, 1984], pp. 17-19, and [R. Kirby, 1989].) We shall deal only with the canonical smooth structures on vector spaces. In particular, if V and W are finite-dimensional real vector spaces, then Hom(V, W) is a vector space whose dimension is dim(V) · dim(VF) and hence has a canonical smooth structure. The special case when V — W = Rn is of particular interest to us. It is the space of real η χ η matrices, which we denote by Mn(R). Example 1.19. Let Μ be an open subset of Rn (or more generally an open set in any finite-dimensional vector space). Then the inclusion map Μ С Rn is an atlas with one chart that provides a smooth structure on M. In particular, G/„(R) = {Ae Mn(R) | det(A) ф 0}, is an open set in the vector space Mn(R) since det: Mn(R) —» R is a continuous map. (In fact, it is a polynomial map.) Thus Gln(R) is, canonically, a smooth manifold. We can say even more. Both of the maps /i:GZn(R) x Gln(H) —» GZn(R) (matrix multiplication), l: Gln(R) —► Gln(H) (inversion) are continuous (and even rational) maps. Thus GZn(R) is an example of a topological group. ♦ Now consider two smooth manifolds Mm and Nn. We can form their Cartesian product Μ χ TV. If (p, q) G Μ χ TV, we may choose charts ([/, φ) and (V, ψ) around ρ and g, respectively. Then U χ V is open in Μ χ TV, and (U χ ν,φ χ ψ) is a chart around (p,q). It is easy to verify that the collection of all such charts is an atlas for Μ χ TV and so determines a smooth structure on it. For example, the canonical smooth structure on Rm+n is the product of the canonical structures on the factors Rm x Rn.
§1. Smooth Manifolds 11 Two great questions in the study of smooth manifolds, which were largely answered in the 1960s, were "Does a given topological manifold necessarily have a smooth structure?" and "If a topological manifold has a smooth structure, is it unique?" The answers to both of these questions is generally no. For example, it is known that the spheres of dimension <6 have unique smooth structures, but there are 28 distinct smooth structures on the 7- sphere. For η > 7, there is generally more than one smooth structure on the η-sphere (cf. [A. Kosinski, 1993]). Smooth Maps Not only does a smooth structure allow us unambiguously to differentiate the composites Φ appearing in the definition of atlases, it also allows us to differentiate certain maps between manifolds. It will take us a while to see how this may be done, and the full story will only appear in §4. Here we prepare the way by describing the functions that we will eventually differentiate. Definition 1.20. A map f:M —► N between smooth manifolds is called smooth (or C°°) if it is continuous and for each point ρ Ε Μ there is a chart (17, φ) on Μ with peU and a chart (V, ψ) on Μ with f(p) e V such that the composite Φ = ψ/φ~ι is smooth. Φ is called the coordinate expression for /. The map / is called a diffeomorphism if it is smooth and bijective with a smooth inverse. ft It is clear that the smoothness of a map is independent of the choice of charts on Μ and TV. What is more, we even have the notion of the rank of a smooth map at a point.
12 1. In the Ashes of the Ether: Differential Topology Definition 1.21. The rank of a smooth map /: Μ —» TV at a point ρ G Μ, denoted by rankp(/), is the rank of the Jacobian matrix Φ'(ψ(ρ)) = Шг'ЧШ), where φ and ^ are charts containing ρ and /(p), respectively. * Varying the choice of the charts merely left and right composes Φ with local diffeomorphisms of Euclidean space which, by the chain rule, left and right multiplies Φ'(φ(ρ)) by invertible matrices, and this will not change the rank. Differentiable topology is the study of the properties of smooth manifolds that are preserved by diffeomorphism. Here is a criterion for a smooth homeomorphism to be a diffeomorphism. Theorem 1.22. Let f: Μ —+ N be a smooth bisection between smooth manifolds on the same dimension m. Then f is a diffeomorphism if and only if its rank at each point of Μ is т. Proof. Both the condition on the rank and the smoothness are local conditions, so it suffices to consider the case when Μ and TV are open subsets of Euclidean space. Let g: N —» Μ be the inverse function for /. Now if g is smooth, then applying the chain rule to g(f(x)) = χ yields g'(f(x))f'(x) = I and hence det(g'(f (x))) det(f'(x)) = 1, so that det(//(x)) ф 0 and thus rankp/ = m. Conversely, if rankp/ = m, then det(//(p)) ф 0 and the inverse function theorem says that there is a unique local inverse h for / satisfying h(f(p)) = ρ and that h is smooth. The uniqueness tells us that g = h on their common domain, so that g = f~l is smooth. ■ Lie Groups We now introduce some of the main players in the study of differential geometry; these are the Lie groups studied by Sophus Lie, Felix Klein, and Wilhelm Killing and developed by Elie Cartan. They constitute the basic symmetry groups of differential geometry, and of nature too. Definition 1.23. A Lie group is a group G that is also a smooth manifold in a way that is compatible with the group structure in the sense that the maps (i) (multiplication) μ: G χ G —► G, (ii) (inversion) l: G —» G are both smooth. cS
§1. Smooth Manifolds 13 Example 1.24. Any finite-dimensional real vector space V is a smooth manifold in a canonical fashion and is an abelian group under vector addition. Since addition and subtraction of vectors are smooth, V is a Lie group. ♦ Example 1.25. We have already seen that Gln(R) — {x e Mn(R) | det(x) φ 0} is a topological group and that multiplication and inversion are the restriction of rational functions Mn(R) x Mn(R) —» Mn(R) and Mn(R) —* -^n(R)· Hence they are smooth and GZn(R) is a Lie group. ♦ More examples of Lie groups may be found at the end of this chapter, on page 63. Definition 1.26. A homomorphism between Lie groups G and Я is a smooth map (p:G —» H, which is also a homomorphism in the sense of group theory. * Example 1.27. The exponential mapping exp: (R, -f) —» (R+, x) is an isomorphism of Lie groups (i.e., it and its inverse are both homomorphisms). Although he did not have the larger concept of Lie groups, it was the existence of this isomorphism that made possible Napier's (1614) construction of his table of logarithms turning multiplication into addition. ♦ Smooth Maps of Constant Rank Now we study the most important species of smooth maps, those of constant rank. For clarity, we begin with three local results that will soon be rephrased in terms of smooth manifolds. Lemma 1.28. Let /:(Rr+rn,0) —» (Rr,0) be a smooth map defined on a neighborhood of 0 which satisfies /'(О) = (7,0).6 Then there is a diffeomor- phism J: (Rr+rn,0) —» (Rr+rn,0) defined on a neighborhood of 0 such that f о J is a restriction of the canonical first factor projection mapping π: Rr χ Rm -» Rr. (x,y) ■-► χ Proof. Define G: Rr χ Rm -» Rr χ Rm. (χ,υ) ·-»· (f(x,y),y) Now G is smooth and G;(0,0) = I n T ]. Thus G is a local diffeomorphism of (Rr χ Rm, 0) with inverse J, say. Since π ο G(x, у) = /(χ, г/), we have foJ = noGoJ = n. Ш So, in particular, / has rank r at the origin.
14 1. In the Ashes of the Ether: Differential Topology Now we apply this lemma to study maps of constant rank. Proposition 1.29. Let /: (Rr+m,0) -> (Rr+n,0) be a smooth map of constant rank r defined on a neighborhood of 0 that satisfies /'(O) r 0 0y Then there are diffeomorphisms, also defined on a neighborhood of 0, of the form #:(Rr+n,0) -» (Rr+n,0) and J:(Rr+m,0) -» (Rr+m,0) such that, on some neighborhood of 0, Η о / о J is a restriction of the canonical map π: Rr χ Rm -> Rr x Rn. (x,y) ·-» (x,0) Proof. Write / as / = (Д, /2): Rr+m -» Rr χ Rn, that is, fx is the first r components of / and /2 is the last η components. Then /{(0) = (/,0), so by Lemma 1.28 there is a local diffeomorphism of the form J: (Rr+rn, 0) —» (Rr+rn, 0) defined on a neighborhood of 0 such that (RrxRm,0)^(Rr,0) is a restriction of the canonical projection. Therefore, replacing / by / о J, we may assume that f(x,y) = (x, f2(x,y)). It follows that Now the condition that / has constant rank r, which is still true for our new /, means that the Jacobian matrix (9/2/Эг/) vanishes so that f2(x,y) is independent of the variable y. Thus, we may write f2(x,y) = h(x), say. Define Я: (Rr+n,0) -» (Rr+n,0) by H(x,y) = {x,y - h\x)). Then Я is a local diffeomorphism and Hf(x,y) = H(xyh(x)) = (x,0). M Exercise 1.30. Show that in general, Proposition 1.28 becomes false if we insist that either Η or J must be the identity. □ Now we put Proposition 1.29 in terms of an arbitrary smooth map of constant rank. Theorem 1.31. Let f: Мш —» Nn be a smooth map with constant rank r. For each point ρ G M, there are charts ([/, φ) and (V, ψ) around ρ and f{p), respectively, such that φ(ρ) = 0, φ{ί{ρ)) = 0, f(U) С V, U and V are connected, and ψίφ~ι is a restriction of the canonical map Rr χ Rm-r _, Rr χ Rn-r (x,y) ι-»· (ж,0)
§1. Smooth Manifolds 15 Proof. Choose arbitrary charts {ϋ,φ) and (V, ?/>) around ρ and /(p). After translation we may assume that φ(ρ) — 0 and ψ/(ρ)) = 0. Then ψ/φ~1'· (Rm, 0) —► (Rn, 0) is defined on a neighborhood of 0, and since the nxm matrix (ψ/φ~ι)'(0) has rank r, we may choose matrices A G GZn(R) and jB G GZm(R) such that (Αφ/(Βφ)-% = Α(φ/φ-1)'0Β-1 = (£ °) . Replacing ^ and φ by Л^ and jEfy?, respectively, we may assume that ψ and φ satisfy the hypotheses of Proposition 1.29. Thus there are diffeomor- phisms tf:(Rn,0)->(Rn,0) and J: (Rm,0) -> (Rm,0) defined on a neighborhood of 0, such that Η ο ψ ο f ο φ~ι ο J: (Rm, 0) —» (Rn,0) is a restriction of Rr χ Krn-r _^ Rr χ Rn-r (я,у) ·->► (ζ,Ο) Replacing ^ by Ηψ and <p by J~ ιφ finishes the proof, except for the conditions on U and V, which we leave to the reader. ■ Now let us mention the following two extreme cases of maps of constant rank. Definition 1.32. Let /: Mm —» Nn be a smooth map with constant rank. Then / is called an immersion if the rank is m, and a submersion if the rank is п. Ф We apply Theorem 1.31 to these special cases to obtain the following result. Corollary 1.33. Let /: Mm -» Nn be a smooth map. Then, (i) / is an immersion (for each point ρ G Μ there are coordinate systems (t/, <p), (V, VO about ρ and /(p), respectively, such that the composite ψ/φ~ι is a restriction of the coordinate inclusion 6:Rm —» Rm χ Rn_rn, (ii) / zs an submersion (for each point ρ Ε Μ there are coordinate systems (C/, φ), (V, ^) about ρ and /(p), respectively, such that the composite ψ/φ~ι is a restriction of the coordinate projection π: Rn χ Rm-n —» Rn
16 1. In the Ashes of the Ether: Differential Topology Proof. In each case, <= is obvious, and => comes from Theorem 1.31. ■ Exercise 1.34. Let Mm be a smooth manifold, and let ι\ Μ —» Rn be a smooth injective map. Show that ι is an immersion if and only if, for each point ρ G M, there is an m-dimensional coordinate subspace V С Rn such that the composite with the orthogonal projection onto V, πι: Μ —» V, has rank m at p. □ Exercise 1.35. Show that in Corollary 1.33 the equivalences are still true if in (i) we prescribe the coordinate system (υ,φ) and in (ii) we prescribe the coordinate system (V, <p). Q Proper Maps, Ernbeddings, and Weak Ernbeddings We conclude this section with a discussion of some variations on the theme of well-behaved mappings. We begin with a discussion of the important notion of proper mappings. Definition 1.36. Let /: X —» Υ be a continuous map with X, У Hausdorff. Then / is called proper if f~l(K) is compact for every compact К С У. (Note that if X is compact, then / is automatically proper.) ft Proposition 1.37. Let X and Υ be topological spaces that are Hausdorff and first countable. Let f:X—*Y be a continuous proper injection. Then f:X —» f(X) is a homeomorphism (where the topology on f(X) is the subspace topology), and f(X) is a closed subset ofY. Proof. Since f:X —> f(X) is a continuous bijection, we need only see that it maps open sets to open sets, or equivalently, that it maps closed sets to closed sets. Let С be closed in X. Suppose that у lies in the closure of /(C) in Y. Since Υ is first countable, there is a sequence of points χι, Χ2, · · · £ С such that f(xj) — yj —» y- Now the set К = {у,у\,У2,·· ·} is compact. Because / is proper, f~l(K) is also compact. Since X is first countable, the sequence xi,X2» · · · lying m f~l(K) Π С has a convergent subsequence, converging to χ G C, say. Since Υ is Hausdorff, the corresponding subsequence of the 2/1,2/2»··· must then converge to f(x), and hence у = f(x) G /(C). Thus /(C) is closed in Υ and / maps closed sets to closed sets. In particular, taking С = X shows f(X) is a closed subset ofY. ■ Definition 1.38. An embedding is a one-to-one immersion /: Μ —» TV such that the mapping /: Μ —» /(Μ) is a homeomorphism (where the topology on f(M) is the subspace topology inherited from TV). ft Proposition 1.39. Л proper one-to-one immersion is an embedding.
§2. Submanifolds 17 Proof. This is a simple consequence of Proposition 1.37. ■ Definition 1.40. A weak embedding is a one-to-one immersion f:M —> N such that, for every smooth map g: S —> N with g(S) С /(М), the induced map g: S —» Μ, defined by # = / ο #, is smooth. ft Lemma 1.41. ///: Μ -+ N is a weak embedding, there is a unique smooth structure on Μ for which this is true. Proof. Let M\ be Μ with a possibly distinct smooth structure such that the map /ι: Μχ —» TV (which is the same as /) is also a weak embedding. Then the induced map id: M\ —» Μ is smooth. Similarly, the induced map zd: Μ —» Mi is smooth, so the two smooth structures are identical. ■ Exercise 1.42. Let A G G/n+1(R), and consider the mapping фл: Pn(R) -» Pn(R) defined by фА(1) = Al. (a) Let an affine coordinate system on Pn(R) arise from the affine map /:Rn —► Rn+1 given by f(x\,... ,xn) = (1»#ь ... ,xn)· Show that in this coordinate system Φα(χ) = ;, where A = I , ) , with dalxl block Ύ v J cx + d \b a J and α an η χ η block. (b) Show that φ a is a diffeomorphism. □ §2. Submanifolds Originally, manifolds were regarded as subsets of Euclidean space; certainly these are the easiest ones to visualize. О <Z> In this section we study how one manifold may be situated inside another one. We give various definitions of increasing simplicity corresponding to the following picture.
18 1. In the Ashes of the Ether: Differential Topology immersed manifolds submanifolds regular submanifolds proper submanifolds Whereas all of these pictures intuitively represent one-dimensional manifolds in the plane, the first two kinds are not generally manifolds in the sub- space topology. The first, an immersed submanifold, has one or more "bad points" (the double point in this picture) or odd convergent sequences in the subspace topology; the second can also have odd convergent sequences. We study a general notion of submanifold which includes all but the immersed manifolds. The latter will be mentioned again at the end of this section. Let Mm be a smooth manifold and let TV71 be a subset. The idea is that TV will be a submanifold if there are charts ([/, φ) on Μ which, locally, straighten out TV. Definition 2.1. Let (υ,φ) be a chart on M. The components of TV Π U are called the plaques of TV in this chart. The chart ([/, φ) on Μ is said to straighten out a plaque W if φ restricts to a homeomorphism between W and an open set in some η-dimensional affine subspace7 А С Rm. In this case, the plaque is called a flat plaque of dimension η and the restriction φ | W: W —» A is called the plaque chart. If TV is covered by flat plaques, we say that TV is locally flat in Μ. Φ The picture corresponding to Definition 2.1 is the following one. 7An affine subspace of Rm is a set of the form {v + a € R771 | ν € V}, where V is a vector subspace of Rm and a 6 Rm.
§2. Submanifolds 19 U\ Note that in this picture TV meets U in two plaques, but here only one gets straightened out by this coordinate system. There is no limitation on the number of plaques in a chart; it may even be infinite. Definition 2.2. Let Mm be a smooth manifold. A subset TV of Μ is called a smooth (n-dimensional) submanifold if there is a covering {Ua} of TV by open sets of Μ such that the components of Ua Π TV are all flat plaques of dimension n.s * Of course we may always "tidy up" φ so that the affine subspace Л is a coordinate subspace, say the coordinate subspace corresponding to the first η coordinates, and so that φ(ρ) = 0; but this is not always appropriate. The coordinate charts guaranteed by the definition tell us that "locally" the pair (M, TV) looks like the pair (Rm, Rn). This eliminates some of the phenomena of strange convergent sequences and the phenomena of limit points of nontrivial topology such as appear in Figure 2.3. (a) (b) FIGURE 2.3. Odd limit points. 8This definition and the lemma and theorem that follow were inspired by a discussion in [P. Molino, 1988], pp. 11-12. We note, however, that the property that every plaque in U Π TV is flat is not easily verified unless TV has some other special property allowing us to assume that C/OTV has just one plaque. The aim is to have a definition broad enough to include nonclosed subgroups of a Lie group and, more generally, the leaves of a foliation (cf. Chapter 2) for which this "one plaque" property may fail.
20 1. In the Ashes of the Ether: Differential Topology Figure 2.3(a) is an example of what is called an immersed submanifold9 which is the image of a manifold under a one-to-one immersion. We are going to define a topology and a smooth structure on a submanifold TV. The topology is called the submanifold topology; it generally has more open sets than the induced topology. Note that each plaque of a submanifold inherits the induced topology from Μ, and in this topology the flat plaques are homeomorphic to open sets in Rn. Definition 2.4. The submanifold topology on TV is the one for which a set U С N is open if it meets every flat plaque in an open set (in the induced topology on the plaque). * Exercise 2.5. Suppose that TV is a submanifold of the smooth manifold M. Show that if {Wa} is a given covering of TV by flat plaques, then U С TV is open if and only if U Π Wa is open for every index a. □ To study the submanifold topology, we shall use the following observation. Lemma 2.6. Suppose that TV is a submanifold of the smooth manifold M. If W С TV is a flat plaque and V С Μ is an open set, then W Π V is a union of at most countably many flat plaques. Proof. Let (U,ip) be a chart on Μ that straightens out W. This means that ip(W) С Rn is an open subset of an η-dimensional affine subspace A. Write W Π V = UWa, where the Wa are the path components of W Π V. Then each Wa is an open set in W Π V and hence in W as well. Hence each ip(Wa) С Rn is open in A. Thus, Wa is a flat plaque straightened out by the chart (U Π V, φ \ U Π V) on M. Moreover, there can be at most countably many of the plaques Wa; otherwise the open set (JQ V>(Wa) С А would have uncountably many components, contradicting the fact that A, and therefore all its subspaces, have countable bases for their topologies. ■ Theorem 2.7. Let Μ be a smooth manifold and TV a submanifold equipped with the submanifold topology. Then (i) N is a topological manifold, (ii) the plaque charts provide a smooth atlas and hence a smooth structure for TV, 9 Suppose /: TV —» Μ is a one-to-one immersion. It would clearly be a difficult matter, and perhaps even impossible, to reconstruct the manifold TV and the immersion / merely from the immersed submanifold itself, that is, from the image set /(TV) С М. The definition we give for submanifold is intermediate between the notion of an immersed submanifold and that of a regular submanifold.
§2. Submanifolds 21 (iii) the inclusion map N С М is a weak embedding. Proof, (i) To see that the inclusion TV С М is continuous, it suffices to show that any open set V С М meets any plaque of TV in an open subset of that plaque. But since the topology on a plaque is just the induced topology as a subset of Μ, this is clear. Since TV С М is continuous and Μ is Hausdorff, it follows that TV is Hausdorff. The flat plaque charts show that TV is locally Euclidean. It remains to see that TV is paracompact in the submanifold topology. For this it suffices to show that each component of TV is a countable union of flat plaques. Without loss of generality, we may assume TV is path connected. Since Μ is a paracompact Hausdorff space, so is TV, and we may choose a locally finite refinement of the open cover {Ua}aei of TV (see Definition 2.2). Calling this locally finite refinement {Ua}aei again, we note that by Lemma 2.6 it still has the property that each component of Ua П TV is a flat plaque. Fix a base point χ G TV. Now if J = {αϊ,...,α*:} is a sequence of elements of 7, we let Aj denote the subset of TV which may be joined to χ by a piecewise smooth path on Μ that lies on TV and passes successively through Uai,Ua2,.-.,Uak' Of course, there may be no overlap between successive C/s, in which case Aj = ®. First note that, because TV is assumed path connected and the unit interval is compact, each point у Ε Μ may be joined to χ by a path in Aj for some J; thus \Jj Aj — TV. Next note that Aj С Uak Π TV is a union of components of Uak Π TV, each one of which is a flat plaque. If we can show that the number of components of Aj is at most countable, we will be finished since the local finiteness of the covering {Ua}aei implies that {J | φ J < к and A j φ 0} is finite and hence TV is covered by at most countably many flat plaques. We show by induction on #J that the number of components of Aj is at most countable. If # J = 1, then clearly Aj has at most one component. Suppose that J = {αϊ,... ,Qfc}, J' = {ai,...,Qfc_i},
22 1. In the Ashes of the Ether: Differential Topology and we already know that Aj> has at most countably many components. By Lemma 2.6, each plaque of Aj> meets at most countably many plaques in Uik Π TV. Thus, Aj has at most countably many components. (ii) The smooth compatibility of the plaque charts is an immediate consequence of the smooth compatibility of the charts on M. (iii) First note that the inclusion TV С М is obviously a one-to-one immersion. Now let g: S —» Μ be a smooth map, with g(S) С TV. To check that the induced map /: S —► TV is smooth, let (W, φ \ W) be a plaque chart. Then by definition, / I f~l(W) is smooth & (φ | W)(f | /_1(^)) is smooth. But (φ | W) о (/ | f~l(W)) = (φ | W) о (g \ g~\W)) = (<pg \ g-\W)) is smooth. ■ Regular Subrnanifolds Now we pass to the definition of regular subrnanifolds, which are simpler and more important than the general subrnanifolds studied above. They intersect more neatly with coordinate charts of the ambient manifold; in particular, the various components of this intersection do not pile up. However, we pay for the simplicity of regular subrnanifolds by the exclusion of important examples, including the nonclosed subgroups of Lie groups and the leaves of some foliations. Definition 2.8. An η-dimensional submanifold TV С М is regular if there is a covering {Ua} of TV by open sets of Μ such that, for each a, Ua Π TV is a single flat plaque of dimension п. Ф Proposition 2.9. For a regular submanifold, the subspace topology is the same as the submanifold topology. Proof. Let TV С М be a regular submanifold. Since this inclusion map is continuous when TV is given the submanifold topology, to show that the two topologies are the same, it suffices to show that given any subset V С TV open in the submanifold topology, there is an open set V С М with TV Π V — V so that V is also open in the subspace topology. Since V is a union of flat plaques, it is sufficient to find V in the case that V itself is a flat plaque. Choose a chart ([/, φ) that straightens this plaque and contains no other plaques. We may assume φ: (С/, V) —► (Rn x Rm_n, Rn χ 0). Then V = φ~ι(φ{ν) χ Βε(0)) will do the job, where Βε(0) is the open ε ball about the origin in Rm_n. ■ Proper Subrnanifolds Even more restrictive than regular subrnanifolds are the proper subrnanifolds.
§2. Submanifolds 23 Definition 2.10. A submanifold TV С М is called proper if it meets every compact subset of Μ in a compact subset of TV (in the submanifold topology). Note that this is equivalent to saying that the inclusion map TV С М is proper. * Theorem 2.11. Let N С Μ be a proper submanifold. Then Μ is regular. Proof. Fix ρ Ε N. Take a chart (ΖΙ,φ) for M, with U compact, that straightens out the plaque W — U Π TV containing p. Now, as in the last paragraph of the proof of Proposition 2.9, we may assume that φ{ϋ) — (p(W) x Βε(0). We may choose a sequence of open sets Uj — φ~ι{φ{νΫ) χ Be/j(0)), j — 1, 2,..., so that (Uj, φ \ Uj) is also a chart straightening out the plaque W and Π?>ι Uj — W. We want to show that some Щ meets TV in just the single plaque W, which will prove regularity. Suppose otherwise. Then we could find an infinite sequence of points χ χ, #2»... on TV — W, with Xi eUj С U each in a distinct plaque of U. Since U is compact, there is a convergent subsequence whose points, together with their limit, constitute a closed and therefore compact set К С Ό Π TV. Repeating this argument with a slightly smaller plaque W С Wf С W, we obtain К С U Π TV. But this is impossible since each point of К lies in a distinct plaque of U and these plaques constitute an open cover of К in the submanifold topology having no finite subcover, which contradicts the compactness of К. Ш Exercise 2.12. Show that neither the union nor the intersection of two proper submanifolds need be a submanifold. □ Submanifolds Described Implicitly Here is a simple criterion for the smoothness of a submanifold described implicitly. Theorem 2.13. Let Мш and Nn be smooth manifolds with m > n, and let f:M—*Nbea smooth map. For у G TV, we set My — f~l(y). Assume that f has constant rank, r say, on a neighborhood of My. Then My is a smooth proper submanifold of Μ of dimension m — r. Proof. Let ρ G My. Since the rank is constant on a neighborhood of p, by Theorem 1.31 there are coordinate charts (U,tp) and (ν,ψ) around ρ and f(p), respectively, such that φ(ρ) — 0, ψ(/(ρ)) — 0, and ι\)ίφ~χ is a restriction of the canonical map π: Rr χ Rmr -> Rr χ Rnr. (X,y) ни- (X,0) Thus φ(υί)Μυ) С 0 x Rm_r in Rr χ Rm_r. Hence My has just one plaque in U and φ straightens out this plaque. This shows that My is a regular
24 1. In the Ashes of the Ether: Differential Topology Mv Μλ λ =1/2 \ Ι A=l FIGURE 2.15. Various elliptic curves. submanifold. Since My is closed and Μ is Hausdorff, it follows that Myf)K is compact for every compact К С Μ. Thus My is proper. ■ Example 2.14. Let M\ = {(x,y) G R2 | y2 = x(x - l)(x - λ)}, which is the Weierstrass normal form of the general cubic (elliptic) curve, λ is a parameter that we will take to be real. Thus M\ is given implicitly as the zero set of the function /: R2 —* R, where /(я,2/) = y2 -x(x- l)(x-X) = y2 -x3 + (X + l)x2 -Xx. Calculating, we have ff(x,y) — (—3x2 + 2(λ -h l)x — A, 2y). The maximal possible rank for the derivative is 1, and we can find all points on M\ where this fails by solving the system x(x-l)(x-X) =y2, 3x2 - 2(A + l)x -l· 1 = 0, J/ = 0. The only solutions are (x,y,X) = (0,0,0) and (1,0,1), which shows that M\ is a smooth submanifold unless λ = 0 or 1 and that, in these exceptional cases, M0 — {(0,0)} and Μχ — {(1,0)} are also smooth submanifolds of R2. Figure 2.15 shows these singular points on Mq and M\, as well as one of the more typical cases of a nonsingular M\. ♦ Example 2.16. Let us reconsider the n-sphere Sn = {x G Rn+1 \x-x = l}. Thus, Sn = F-X({1}), where F:Rn+1 -» R given by F(x) = x-x. Calculating from first principles, Ff(x)v = 2χ·υ. Thus, F has rank 1 everywhere on Sn. It follows from Theorem 2.13 that Sn is a smooth submanifold of Rn+1. ♦
§2. Submanifolds 25 Example 2.17. Let On(R) = {A e Afn(R) | A A1 = 1} = the orthogonal group. Define F: Mn(K) -> Mn(R) by F(4) = AA*, so that On(R) = F~l(I)· Now F is smooth and we may calculate F'{A)V = lim | {F(A + ftV) - F(j4)} /i->0 ft = lim 1{(A + ftV)(i4 -l· hVf - AA1} h—>0 h = lim jfftVA* + hAVb + ftW} Л—0 ft In the particular case that A — I, we have ker F'(I) = on(R), the vector space of all η χ η skew-symmetric matrices. Note that dim on(R) = ^n(n — 1). More generally we have an isomorphism on(R) ~ ker Ff(A), which sends S ·—► SA for any A Ε On(R). In particular, the rank of F is constant along On(R), so On(R) is a smooth submanifold of Mn(R). ♦ Exercise 2.18. Note that the smooth map det: On(R) —* R has values ±1 (since 1 = аеЬ(ААг) = det(4) det(A) = det(^)2), so it follows that On(R) is the disjoint union of two open sets. Show that these are the components of On (R). □ Exercise 2.19. Assume the situation of Theorem 2.12; let W С N be a submanifold; and assume for each χ e f~l(W) that f*(TxM) + Tf^W = Tf{x)N. Show that f~l{W) is a submanifold of M. □ Submanifolds Described Parametrically Now we pass to the dual situation of a manifold described parametrically. Theorem 2.20. Let Mm and Nn be smooth manifolds, and let f:M -> N be a proper one-to-one immersion. Then f(M) is a regular submanifold of N and the map f:M —* f(M) is a diffeomorphism. Proof. We first show that f(M) is a regular manifold. Choose ρ G M, and set q — f(p)· Since / is an immersion, its rank is m everywhere. Then by Corollary 1.33 there are connected coordinate charts (U,tp) and (V, VO around ρ and q, respectively, such that φ(ρ) = 0, ψ(ς) — 0, f(U) С /(V), and ψίφ~ι is the restriction to φ{ϋ) of the canonical inclusion l: Rm -> Rm χ Rn"m. χ ■-► (z,0) Thus ψ{ί(ϋ)) lies in an open subset of Rm χ 0 С Rm χ Rn_m. Using the fact that / is proper, an argument similar to the proof of Theorem 2.11
26 1. In the Ashes of the Ether: Differential Topology allows us to shrink U and V so that f(U) Π V = f(M) Π V is connected. Since (V, ψ) straightens out the unique plaque f(M) Π V in V, it follows that f(M) is a regular submanifold. By Theorem 2.7(iii), f(M) С TV is a weak embedding. Since f:M -> N is smooth, it follows (cf. Definition 1.40) that the map /: Μ —» f(M) is also smooth. Since both f(M) С TV and f: Μ —+ N are one-to-one immersions, it follows that /: Μ —» /(Μ) is a bijective immersion. Then by Theorem 1.22, / is a diffeomorphism. ■ This theorem shows that the image of a proper embedding is a smooth submanifold. We remark that the embedding theorem of Whitney tells us that every smooth η-manifold admits a proper embedding as a smooth submanifold of Euclidean space of dimension 2n -h 1 (cf. [J. Milnor, 1965]). Immersed Manifolds and Immersed Submanifolds Definition 2.21. An immersed manifold in Μ is the image of an immersion /:7V —► M. (Note that the image is not in general a submanifold!) An immersed submanifold is the image of a one-to-one immersion /: TV —» M. For example, the elliptic curve Μχ appearing in Example 2.14 may be regarded as an immersed manifold in R2. In such a case, it is still true that each point ρ Ε Ν has a neighborhood U С N such that f(U) is a submanifold of Μ diffeomorphic to U. Exercise 2.22. Show that the curve Mx = {(x,y) G R2 | y2 = x{x - l)2} is an immersed submanifold of R2 by parametrizing it by the slope of the line through (x,y) G M\ and (1,0). □ Exercise 2.23. Show that all of the following inclusions are proper inclusions: , , ч ( weakly 1 < , .. .. > D < embedded > D {submanifolds} submanifolds I . .. .. [ J 4 J у submanifolds) J regular 1 J proper \ submanifolds J [ submanifolds (Hint: To show that the first and second inclusions are proper, consider Figure 2.3.) □ The following exercise provides a preview of the "second fundamental form" of a submanifold of Rn.
§2. Submanifolds 27 Exercise 2.24.* Suppose that Μ С R is given implicitly in a neighborhood of ρ G Μ by the zero set of a function of maximal rank F: ΈΙΝ —» ΈΙν. The tangent space to Μ at χ is defined to be the kernel of the linear surjection ^(χ):ΚΝ —► Kv and is denoted TX{M). Define the second fundamental form of Μ at χ by the formula BX:TX(M) χ ГЖ(М) -> ΚΝ/Τ(Μ), F'(x){Bx{u,v)) = F"(x)(u,v). It follows that the symmetric bilinear form Bx(u, v) G RiV/T(M) is defined. (i) Let a(t) be path on Μ С ΈΙΝ through χ and v(t) a vector field tangent to Μ along σ. Show that at x,v — —Βχ(σ,ν) mod TX(M). (ii) Use (i) to show that BX:TX(M) χ ΤΧ(Μ) -» R^/T^M) is independent of the choice of F. (iii) Show that by using the Euclidean metric on ΈΙΝ we can reinterpret Bx as a map Б^: TX(M) χ TX(M) —► ХЦМ)1- (orthogonal complement) and then define the "transpose" BX:TX{M) x T^M)1· - TX(M), {u,Bx(v,w)) = {Bx(u,v),w) for all u,υ eTx(M), weTx(M)L. Show that if a(t) is a path on Μ С ΈΙΝ through χ and w(t) a vector field normal to Μ along σ, then at x, it/ = £?£(σ,ΐ/;) mod ХЦМ)-1. (iv) Find an analog of (iii) that does not rely on using a Euclidean metric. □ In this formulation of the second fundamental form, it is clear that it has fairly strong invariance properties. The following exercises make this more precise. Exercise 2.25. Suppose that Μ С HN is given implicitly in a neighborhood of ρ G Μ as the zero set of a function of maximal rank F: ΈΙΝ —» Rv, and let φ: HN —► HN be a diffeomorphism. Set F = F о 0, and let Б and В be the second fundamental forms associated to Μ — F~l(0) and Μ = F_1(0), respectively. Show φχ(Β(ύ,ϋ)) = B(u,v) + φχ(ϋ,,ϋ) mod Τφ(χ)(Μ) for all u,v e TX(M), where w = фх(й) and ν = φχ(ν). □ Exercise 2.26. Let Μ С RN and let φ: ΚΝ -> RN be any invertible affine map. Let Bx (respectively, Βψ^) denote the second fundamental form of Μ at χ (respectively, of φ(Μ) at ф(х)). Show that φχ(Βχ(η,υ)) = Βφ(χ)(φ'χη,φ'χυ) mod TX{M) for all ti,ti6 TX(M). □ Exercise 2.27. Assume the hypotheses of Exercise 2.26, and let ρ: ΈΙΝ —» R be a smooth function. Define φ — -ф: HN —► RN, and show that φ'^η,υ) = --{р'х(у)Ф'х(и) + р'х(и)фх(у)и - φ'^η,υ)} mod φ(χ)
28 1. In the Ashes of the Ether: Differential Topology for allu^eR . (Of course, we must avoid points where ρ vanishes.) Using the results of §4, deduce the analog of Exercise 2.26 for a submanifold Μ С PN(R) and an invertible projective transformation φ:Ρη(ΈΙ) —» Pn(R). □ Exercise 2.28. Assume the hypotheses and notation of Exercises 2.25 and 2.27. Let p(x) = χ/(χ·χ) (inversion). Show that, in this case, ф'х(Вх(й, ν)) — Вф(х){и,у) - 2(u · ν/ρ{χ))φ{χ). □ Note that Exercise 2.28 shows that the "raw" second fundamental form fails to be invariant under inversion. Nevertheless, we can obtain an invariant second fundamental form by taking the "traceless" version of £?, as indicated in the following exercise. Exercise 2.29.* Let V be a Euclidean space (with inner product "·") and W a vector space. Let B: V x V —» W be a symmetric form. Define Trace * = Σ,<4 <dim ν ^(еь ег) £ W, where {e^} is an orthonormal basis for V. Define Bo(u, v) — B(u, v) — (Trace £?/dim V)u · v. (a) Show that Trace В is independent of the choice of the basis. (b) Assuming the hypotheses of Exercise 2.28, show that ф'х(Во(й,у)) — В0(щу). □ We shall revisit some of the ideas of these exercises in Chapters 6 and 7. §3. Fiber Bundles Everyone knows the usefulness in advanced calculus of studying vector- valued functions. Consider, for example, the velocity field of a particle moving along a fixed curve 7 in R2. This field is described by a function v:j —► R2 defined along the curve. From another point of view, this same field may be regarded as a function whose value at χ e 7 lies in the one- dimensional subspace Vx of R2 that is parallel to the tangent line Tx to the curve at x.
§3. Fiber Bundles 29 We emphasize that {Vx \ χ e 7} is a family of subspaces parametrized by 7· Fiber bundles formalize the notion of a "parametrized family of so-and- so's." The "so-and-so" is called the fiber and may be a vector space as in the example above (corresponding to a vector bundle) or a Lie group (corresponding to a principal G bundle). These two are the most important cases, although other fibers may also serve as necessary. Sections (or in the case of vector bundles, one often says vector fields or tensors), which generalize functions, are functions on the parameter space whose value at x lies in "the so-and-so at x." If the parametrized family is constant, then fields (sections) reduce to ordinary functions. The justification for the degree of generality employed here will appear first in the next section, where it is shown that a smooth manifold has a tangent bundle intrinsically associated with it, and again in Chapter 5, where it is shown that a Cartan geometry has a principal G bundle intrinsically associated with \t. As with manifolds, it is possible to discuss both topological and smooth bundles. Of course, our principal interest is in the smooth case. Topological Bundles and Smooth Bundles Definition 3.1. Let F be a topological space and π: Ε —> В a continuous map. We call the quadruple ξ — (E,B,n,F) a {locally trivial) fiber bundle with (abstract) fiber F if, for each point b G B, there is an open set U С В containing ρ such that π~ι(ϋ) is homeomorphic to U χ F by a homeomorphism φ such that the following diagram commutes. K~\U) —^—^UxF A /proJ! U The pair (U,tp) is called a chart (or local bundle coordinate system). Φ Note that π-1 (6), the fiber over B, denoted i% is homeomorphic to F for all b G В. В is called the base space and Ε the total space of the bundle. Given a bundle £, we sometimes denote by Β (£), Ε (ξ), etc., the corresponding parts of the bundle. If the base, fiber, and total spaces are smooth manifolds, π is a smooth map, and the charts may always be chosen to be smooth maps, then we have a smooth bundle. Two bundles ξι and £2 over the same base В are said to be isomorphic if there is a homeomorphism (diffeomorphism) φ between the total spaces such that the following
30 1. In the Ashes of the Ether: Differential Topology diagram commutes, φ is called a bundle isomorphism or, in the case when £x = ξ2 a bundle automorphism. Example 3.2. The simplest example of a bundle is the product bundle Ε — Β χ F with π = πχ (projection on the first factor). This bundle is also called the trivial bundle. If Ε is a locally compact Hausdorff space, then a necessary and sufficient condition that a bundle ξ — (Ε, £?, π, F) be (isomorphic to) the trivial bundle is the existence of a trivializatwn, which is a continuous (smooth) map t: Ε —► F that induces a homeomorphism (diffeomorphism) upon restriction to each fiber π-1 (6). In this case the map (p, t):E —» Б χ F is a homeomorphism (diffeomorphism) commuting with the canonical projection to B. ♦ Example 3.3. Perhaps the simplest example of a nontrivial bundle is the Mobius band Ε pictured in Figure 3.4. It is clear that if we remove any point from the base B, we get an open set U homeomorphic to the interval (0,1) and n~l(U) is homeomorphic to (0,1) χ [0,1]. Thus it is a bundle over В — Sl with fiber I = [0,1]. But it is not a trivial bundle; if it were, the boundary would consist of two components, whereas in fact it has only one (Figure 3.4). ♦ 'В UxF / ί projection on 1 πχ [the first factor] FIGURE 3.4. Mobius band.
§3. Fiber Bundles 31 Example 3.5. The canonical line bundle over projective space generalizes the Mobius band. This is the bundle π: Ε —► Pn(R), where Ε — {(I, v) e Pn(R) x Rn+1 | ν β 1} and π(1,ν) = I. To see this is a bundle, let a:Rn —► Rn+1 be an affine map whose image does not contain the origin (cf. Example 1.3). Then Ua = pa(Rn) is open in Pn(R) and pa:Hn —» Ua is a diffeomorphism. We define <p~x: t/a x R —* π~ι(υα) by (Ζ, λ) ι—► (/, Xa(pa)~l(l)). It is easily checked that φ~ι is a diffeomorphism that is linear on fibers, and that the pairs {(Ua,ipa)} constitute an atlas for the line bundle E. In a similar fashion we may describe the canonical line bundle Ε over the projective space P(V) associated to any vector space V. ♦ Example 3.6. By definition, every covering space π: Ε —» В is a fiber bundle with discrete fiber. ♦ G Bundles Now we consider the following, more refined notion of a bundle in which a Lie group of symmetries appears. Definition 3.7. Let ξ — (Ε, Β, π, F) be a smooth fiber bundle, and suppose that G is a Lie group that acts smoothly on F as a group of diffeomor- phisms. A G atlas for ξ is a collection Я = {(υ^ψί)} of charts for ξ such that (i) the Ui cover B, (ii) for each pair of charts ([/, φ) and (V, V7) m ·# the map Φ - ν*?"1: (t/ П У) x F -+ (t/ П 10 x F, called a coordinate change, has the form Ф(и, f) = (u,h(u)f), where h: U Π V —» G is a smooth map called a transition function. Φ Note that in the case when the homomorphism G —» Diff(F) has a kernel Я", then Я is a closed10 normal subgroup and the action factors through G/H, so the bundle may also be regarded as a G/H bundle.11 Thus there is nothing lost if we make the restriction that G acts effectively on F, namely, that Η — 1. In this case we may speak of an effective G bundle. Just as in the case of manifolds, we call two G atlases equivalent if their union is also a G atlas. 10Let Δ = {(χ, χ) G F χ F \ χ € F}. Since the action is a smooth map Gx F —* F, the graph of this map, given by Γ = {(#, χ, г/) € G χ F χ F | gx = y}, is closed and hence so is the intersection TC\GxA = HxA. 11 In Chapter 4 it will be shown that Η and G/H are again Lie groups.
32 1. In the Ashes of the Ether: Differential Topology Definition 3.8. A G structure on the smooth bundle ξ is an equivalence class of G atlases on ξ, and a G bundle is a smooth bundle ξ with a specified G structure. * Exercise 3.9 (Product bundles). Let ξj — (Ej,Bj,nj,Fj,Gj), j — 1, 2, be two "G" bundles. Show that ξιχξ2 = (Εχ χ Ε2, Β ι χ Β2, πι χ π2, Fx x F2, Gi χ G2) is a Gi χ G2 bundle in a canonical fashion. □ Exercise 3.10 (Induced, or pullback bundles). Let ξ — (E,B,n,F,G) be a smooth G bundle and let f:X —► В be a smooth map. Let /*(£) = (ΕΊ,Χ,πι,^, G), where Εχ = {(ж,е) G XxE* | /(χ) = 7r(e)} and 7Ti(:c,e) = x. Show that it is a smooth G bundle. □ Definition 3.11. A G bundle is flat if all the transition functions h:U P\V —+ G are constant. (This happens, in particular, if G is a discrete group.)12 * Exercise 3.12. Let В be a manifold with universal cover B. Show that a flat G-bundle over В with fiber F induces a trivial bundle over В with a canonical trivialization and that the action of the group of covering transformations on the fiber is a representation n\(B,b) —» G. Conversely, show that every representation πι (£?,&) —» G arises in this way. □ Given a G atlas, we can consider the union of all G atlases equivalent to it to obtain the unique maximal atlas equivalent to the given one. The 12The reader may wonder about the origin and importance of flat bundles in geometry. It often happens that one makes some locally defined geometrical construction on a manifold which depends on a choice in some discrete universe of possible choices. By the discreteness any choice can be extended smoothly over small neighborhoods, but globally one may not be able to do this. The appearance of this situation is a sure sign of the presence of a helpful flat bundle that resolves the global ambiguity. For example, at a nonumbilic point ρ on an oriented surface in 3-space, there are two principal tangent lines corresponding to the two principal curvatures. Although we may choose two unit vectors e\ and ei in these two directions at p, there is no unique choice. Of course, we may assume that the basis (ei,e2) gives the correct orientation for the surface, but that still leaves the choice between ±(ei,e2). If we work only locally, we may just choose one of the two, say (ei, ег), and be done with it. It is a fact, however, that as we push our choice (ei,e2) along a loop enclosing a single (generic) umbilic point, it will come back to ρ as the other choice, —(ei,e2). This indicates the presence of a flat bundle, in this case the flat line bundle L with discrete group ±1, with the property that the fiber turns over every time we pass around a generic umbilic. Rather than regarding ei and €2 as sections of the tangent bundle Τ of the surface, which would run into sign trouble as we pass around an umbilic, we regard them as sections of the tensor product Τ ® L. The rectification of the sign is built into L. And, moreover, rather than saying that we have now accommodated the frame by a clever construction, it is more true to say that the frame was always geometrically comprised of sections of Τ 0 L but we just didn't know it before.
§3. Fiber Bundles 33 G structure may be identified with this maximal atlas, and the remarks about tidying up the coordinate charts in the corresponding situation for smooth manifolds apply here as well. Two G bundles are isomorphic if they are isomorphic as bundles by an isomorphism identifying the G structures. Example 3.13. A finite covering space π: Μ —» Μ is a G bundle with discrete fiber whose group is the permutation group of the fiber. (If the fiber is not finite, the permutation group is not countable.) ♦ Example 3.14. We sketch a more complicated example of a G bundle, the Hopf fibration 51 -> 53 -> 52. The Lie group 51 = {AgC||A|-1} acts smoothly on S3 = {w = (wo,wi) e C2 | \w0\2 + \wi\2 = 1} by the formula λ · (wo,w\) = (\wo, \w\). We shall regard 52 as the union of two copies of С with identifications, 52 = Co υφ Ci, where φ: C$ —» CJ is given by φ(ζο) — Zq1. The map φ may be regarded as arising from two stereographic projections from the sphere of radius 1/2 in R3 as in the following figure. (The complex planes are arranged so that their positive real axes are parallel.) Cj (= copy of C0 "turned over") Co(=C) Now the map π: 53 —► 52 is given by zx =z0l =Φ(ζ0) n(w) — < wi/wo G Ci if wo ф О, wq/wi e Co if w\ ф 0. Note that if wo,w\ Φ 0, the two definitions agree. It is easy to verify that π(ύ) = π (υ) <=ϊ и = λν for some λ Ε S1. The trivializations are тг-ЧСо) » Co x S\ 7r_1(Ci) » Cx χ S\ (w0,wi) ^ (wi/w0,wo/\w0\), (w0,wi) ^ (w0/wi,wi/\wi\), and the coordinate change between these is seen to be t^Q X *b > L^j^ X »b , (z,\)»{l/z,\z/\z\). Since this is of the form (ζ,Χ) ι—► (ip(z),h(z)\), we have an 51 bundle over 52. The inverse image of a circle on 52 in 53 is a torus. Picturing
34 1. In the Ashes of the Ether: Differential Topology north pole fiber FIGURE 3.15. The Hopf fibration. 53 minus the north pole as R3 via stereographic projection, Figure 3.15 shows an accurate cutaway perspective view in R3 of the fibers on the three tori lying over the equator, and the ±45 degree parallels of 52. The fibers themselves correspond to the places where the ratio Wi/wq is constant and appear as circles in Figure 3.15. The fibers over the north and south poles, namely, the z-axis and the "central circle," are also shown. In fact, the subset of R3 consisting of R3 — z-axis - central circle is "foliated" by the tori lying over the lines of latitude on the 2-sphere. All but two of the fibers of the Hopf bundle lie on these tori in R3. On each torus the fibers run around "once in each direction." The base 52 may be regarded as "the space of fibers" in the sense that π establishes a bijection between the set of fibers and the base. Moreover, the topology on the base indicates the mutual disposition of the fibers in the sense that two fibers are "near" each other when the corresponding points in В are "near" each other. Figure 3.16 shows how the "space of fibers" of the Hopf bundle fit together topologically to form the 2-sphere. One of the tori of Figure
§3. Fiber Bundles 35 cross sectional disc FIGURE 3.16. Topology of the space of fibers. 3.16 has been singled out. A cross-sectional disc for the corresponding solid torus meets every fiber inside the torus just once. The "complementary disc" meets each fiber outside the torus once. The fibers on the torus meet each disc once in its bounding circle and provide a diffeomorphism between these two circles. Thus the "space of fibers" is made up of two 2-discs glued together by a diffeomorphism of their boundaries, which is just the 2-sphere. ♦ Exercise 3.17. Show that the Mobius band is a G bundle with G = Z/2. □ Construction of Bundles Now that we have seen an interesting example of a G bundle, let us return to the definition to see what ingredients are required for the construction of a G bundle. It is clear from the definition that if we are given an effective G bundle, then we are given (i) a Lie group G acting smoothly on a smooth manifold F, (ii) an open covering {Ua} of a manifold £?, (iii) smooth maps hpa: UanUp —> G related by the property that if Ua Π UpP\U7 φ 0, then on this intersection we have h7phpa = hlot. The latter property is a consequence of the existence of the effective G bundle and may be seen as follows. We may assume that the open sets Ua are indexed by the trivializations one of the fibers on the torus <*:K~l(Ua)->Ua xF.
36 1. In the Ashes of the Ether: Differential Topology The coordinate changes are βα l: (Ua Π Up) xF-> (Ua Π Up) χ F, where -1 IS Pa~l(u,f) = (u,hpaf). Thus, the obvious identity ^β~ιβα~χ = ηα equivalent to h1php(X = h1QL. It is quite clear that this procedure may be reversed in that if we are given the data of (i), (ii), and (Hi), we can construct a G bundle by forming the disjoint union of the products Up χ F and dividing by the equivalence relation generated by making the identifications UaxF U {Ua nUp)xF (ti,/) UpxF U (UanUp)xF (U,ft/3Q/)· Exercise 3.18. Verify that this procedure does yield a smooth G bundle over В with fiber F and that if the data (i), (ii), and (iii) arose from a G bundle, this procedure reconstructs that G bundle (up to isomorphism). □ Principal Bundles A special type of G bundle is the one for which the group G "is the same as" the fiber F in the sense that for some (and hence any) point /o G F, the map G —> F sending g —» gf0 is a diffeomorphism. (For example, the Hopf bundle fits this description.) It then follows that the G bundle is effective, so the transition maps h:U C\V —> G are determined by the bundle. Note that the diffeomorphism G —> F does not yield a canonical identification of G with the abstract fiber F, because the bijection will vary with the choice of /o G F. Nevertheless, once we have chosen such an identification, we can reconstruct the bundle using G itself as the (abstract) fiber together with the left action of G on itself, as may be seen from the following diagram. (*,gf0) I UxF 3 Τ (UnV)xF UxG 3 | (UnV)xG (μ, g) h -> (u,h(u)gf0) -> (UnV)xF VxF -> (UnV)xG | с VxG > (u,h(u)g) Moreover, it further follows from this diagram that a principal G bundle has a smooth right G action, as may be seen by comparing the coordinate changes before and after we identify the fiber with G. The right G action commutes with the coordinate changes, which themselves involve only the left G action. This leads us to the following definition.
§3. Fiber Bundles 37 Definition 3.19. A principal13 G bundle is a smooth fiber bundle ξ = (Ρ, Ρ, π, F) together with a right action Ρ xG —> Ρ that is fiber preserving and acts simply transitively on each fiber. ft Example 3.20. A regular covering space π: Μ —» Μ is a G bundle with discrete fiber whose group is πι(Μ, fe)/iV, where AT = Image π*: πι (Μ, Ь) —» πι(Μ,6). ♦ Exercise 3.21. Verify that a principal G bundle does in fact possess a canonical effective G bundle structure. [Hint: the coordinate changes must commute with the G action.] □ Exercise 3.22.* Let ξ = (Ρ, Ρ,π, Ρ) be a principal G bundle. Show that the action Ρ χ G —» Ρ is proper in the sense that if Л and Ρ are compact subsets of Ρ then {# e G | (#Л) Π Β φ 0} is compact. □ Exercise 3.23.* Let Η —» Ρ —» Μ be a principal bundle, and suppose that Μ is connected. Fix a component Pi of Ρ and an element pi Ε Pi, and set #x = {h G Я | pxh e Pi}. (i) Show that Η ι is a codimension-zero subgroup of H. (ii) Show that ΡχΗχ С Рх. (iii) Show that H\ —» Pi —» Μ is a principal #i bundle. □ The case of a principal G bundle is fundamental because of the inverse constructions that allow us to pass back and forth between effective G bundles with fiber F and principal G bundles. These passages may be described as follows. An effective G bundle with fiber F determines the transition functions h: U Π V —» G, and hence, as above, we can construct a principal G bundle called the associated G bundle. Conversely, given a principal G bundle ξ and a smooth effective action of G on a manifold P, we can use the action to construct a G bundle with fiber F denoted by ξ Xq F. These two constructions are inverse to each other. Therefore, we may say that the fiber of an effective G bundle may be regarded as a variable for which we may substitute any manifold on which G acts effectively. Vector Bundles A special case of a G bundle is a real vector bundle. In this case the fiber is a real finite-dimensional vector space V and the group is the general linear group G = Gl(V). An isomorphism of vector bundles is an isomorphism of This term apparently arises because a principal bundle generalizes the so- called principal group of a Klein geometry (cf. Chapter 4).
38 1. In the Ashes of the Ether: Differential Topology bundles that is linear on the fibers. Since the action of G on V is transitive and effective, it follows that we may pass back and forth between the vector bundle and the principal G bundle ξ. What is more, given any smooth linear representation p:G —> G(W), we can pass from the original vector bundle to the associated principal bundle ξ and then to the G bundle with fiber W denoted by ξ xp W. In fact, the fibers of the canonical map ξχ\Υ —> ξχρ\¥ are just the orbits of the left G action G χ ξ xW —> ξ xW given by g-(x,w) = (xg~\ p(g)w). Example 3.24 (Exterior power). If ξ is a vector bundle with fiber V, with group Gl(V) acting on V by the standard representation, then Gl(V) also act on AP(V), the pth exterior power. Thus, we have an associated bundle Λρ(£), called the pth exterior power of ξ. ♦ Exercise 3.25 (Whitney sum). Let ξι and £2 be two vector bundles over the same base £?, and let Δ: Β —> Β χ В be the diagonal map. Then the Whitney sum of the bundles is ξι θ£2 = Δ* (ξι х£г)· Show that the Whitney sum is associative and commutative (up to bundle isomorphism). □ Exercise 3.26 (Quotient Bundle). Let ξι and ξ2 be two vector bundles over the same base £?, and let φ: ξι —» £2 be a smooth injection covering the identity on В and linear on fibers. Show that there is a canonical quotient bundle £2/^1 over В whose fibers are the quotients of the corresponding fibers of ξι and £2· ^ Sections Now let us consider sections (also called fields or tensors, in the case of vector bundles), the generalization of functions mentioned in the introduction to this section. Definition 3.27. Let ξ = (Ε, #,π, F) be a bundle. A section over U С В is a continuous (or smooth, in the case of a smooth bundle) map a:U —> Ε such that πσ = idjg. The space of all (smooth) sections of η is denoted by Γ(η). * In general, a field in an n-dimensional real vector bundle may be regarded as a generalization of a vector-valued function on the base, which is locally, of course, exactly what a field is in this case. In fact, the same may be said about sections on a general bundle. The advantage of the generalization is that, in many geometric circumstances, fields correspond more closely to the nature of things than do functions. The close relationship between fields and functions accounts for the fact that in the presence of a trivial bundle
§4. Tangent Vectors, Bundles, and Fields 39 with a canonical trivialization t:E —> F, there is a tendency to identify the field σ:Β—>Ε with the corresponding function ta: В —> F. Let ξ = (£*,£?, π, G) be a principal G bundle over £?, and consider the bundle η = ξ XgF with fiber F, associated to ξ by a smooth effective action ρ of G on a manifold F. There is an important bijective correspondence ψ:Α°(Ε,ρ)^Γ(η) where A°(E,p) = {f:E -> F | / is smooth, and f(pg) = p(g-l)f(p)} and the section ψ(/): Β —» τ/ is induced by the map Ρ —> Ρ χ F sending p^ (pJ(p))· Exercise 3.28.* Verify that the above correspondence is well defined and bijective. □ §4. Tangent Vectors, Bundles, and Fields Tangent vectors have a schizophrenic nature. On the one hand, they have a geometric aspect in which they appear as directions in space: if I stand in a manifold, I can move in a variety of directions, which may be described as the tangent vectors at my position. On the other hand, they have an analytical aspect in which they appear as "directional derivatives," which is to say as first-order partial differential operators that, when applied to a smooth function, give its rate of increase in the given direction. Although these are "the same" notion in some sense, they nevertheless have a different development. While "directions" strike the mind as a primarily geometric notion, differential operators lie in the domain of analysis and can, for example, be composed to yield partial differential operators of arbitrary order whose geometric significance is often unclear.14 Geometric Vectors We start with the geometric view of vectors. In courses on advanced calculus, we learn how, at least in certain cases, to find the tangent plane at a point ρ of a submanifold Mm С Rn. There are two methods, depending on whether the submanifold is described parametrically or implicitly. Of 14In fact, we are simplifying matters somewhat. The schizophrenic nature of tangent vectors also includes a third personality, an algebraic one. If mx is the ideal of smooth functions on Μ which vanish at the point x, then the space of tangent vectors at χ can also be identified with the dual space (mx/m£)*. However, we do not pursue this aspect of tangent vectors here.
40 1. In the Ashes of the Ether: Differential Topology course, in cases where both methods apply, they lead to the same tangent plane which we shall denote by Tp(M)geometric· Later we shall drop the term geometric. The Parametric Method. Let ρ G Mm, where Mm is a submanifold of Rn. Assume that ρ lies in a neighborhood of Mm which is defined para- metrically by a parameterization /: (t/,0) —» (Mm,p), where U is an open set in Rm and rank0(/) = n. Then the tangent plane TP(M)geometric is also described parametrically as the image of the affine map A: Rm —» Rn given hyAv = f(0) + f'(0)v. Exercise 4.1. Let on(R) = {X e M„(R) | X1 = -X} be the \n(n - 1)- dimensional vector space of skew-symmetric matrices. Show that the Cayley parameterization /:o„(R)->Mn(R), where f(X) = (I + X)(I - X)~l is a smooth embedding whose image is an open subset of 50n(R). □ Example 4.2. Let us find the tangent plane at the identity I = /(0) of the submanifold described parametrically by / in Exercise 4.1. Calculating the derivative /;(0) from first principles, we have lim \{f(hY) - /(0)} = Km 1{(/ + hY)(I - hY)~l 1 1} = bm-{{I + hY)-{I-hY)}{I-hY)-1 = 2Y. Thus, the tangent plane at the identity is (a translated copy of) on(R) itself, parametrized by the affine map A: on(R) —» Mn(R) given by A(Y) = I + 2Y. ♦ The Implicit Method. Let ρ e Mm, where Μ is a submanifold of Rn. Assume that ρ G U С Rn, where U Π Μ is defined implicitly as the zero set of a smooth map F: U —» Rn_m such that rankp(F) = n — m. Then the
§4. Tangent Vectors, Bundles, and Fields 41 the two geometrical placements difficulty separating of tangent spaces tangent vectors FIGURE 4.4. tangent plane Tp(M)ge0metric is also described implicitly as the zero set of the affine map B: Rm -» Rn given by Bv = F'{p){v - p). Example 4.3. Let us calculate the tangent plane of the η-sphere Sn = {x G Rn+1 | χ · χ = 1} by this method. Here F:Rn+1 -» R is given by F{x) = χ · χ — 1. Calculating the derivative of F from first principles yields F'{x)v = 2x - v. Thus, rankx(F) = 1 for each χ G Sn, and the tangent plane is given by Tx(Sn)geometr[c = {x + v e Rn+1 \x -v = 0}. ♦ These constructions of the tangent planes suffer from several disadvantages. The first one, which is not very serious, is that although the tangent plane can be regarded as a vector space, it is generally not a vector subspace of the ambient Rn (Figure 4.4). This can be fixed by translating each tangent plane TP(M) geometric by — ρ so that it becomes a subspace with ρ corresponding to the origin (cf. Figure 4.4). We continue to denote this by TP(M)geometric· The more substantial second objection is that vectors in distinct tangent planes may correspond to the same point in Rn. This can be remedied by means of the following construction, which defines the geometric tangent bundle T(M)geometT-lc of a smooth submanifold Μ of Rn: T(M)geometric = {(p,v) G Г X Rn | P G M, V G Tp(M)}. It is easy to see that T(M)geometric is always a smooth submanifold of Rn χ Rn. For example, if Μ is given implicitly as Μ = {ρ € R" | F(p) = 0}, where F: R™ -» Rn_m is smooth with rankp^) = η - m for all ρ G M, then T(M) is given implicitly as T(M)geometric = {(p,v) € R" x Rn | F(p) = 0,F'(p)v = 0},
42 1. In the Ashes of the Ether: Differential Topology namely, as the zero set of the smooth function (F, B): Rn χ Rn -> Rn"m χ Rn"m, where B(p, v) = F'(p)v. Let (x, y) denote the coordinates of the two factors Rn χ Rn. Calculating the derivative, we have (^'MHffl Д ι. \дхэ) \dy3)J which has rank 2(rc — m) everywhere, since both main diagonal entries are copies of F'(p), which has rank η — га. Exercise 4.5. Show that if Μ is given implicitly by F = 0 as above, then the first factor projecton π:Τ(Μ) —> Μ has a canonical vector bundle structure. □ Analytic Vectors As interesting and geometrically explicit as the preceding results are, there still remains the major disadvantage that the definition of the tangent plane and that of tangent bundle seem to depend on the way in which Μ is embedded in Rn. It is much more satisfying and useful to ask and answer the following question: "Is there an intrinsic way of associating a tangent plane and tangent bundle to a smooth manifold so that it is not only independent of a particular embedding, but also so that if we do have an embedding Mm —» Rn, we can see that we get the same answer as above?" Using Whitney's embedding theorem it is possible, although awkward, to show directly that the above construction is what we seek. However, it is better now to abandon, temporarily, the geometric aspects of vectors and pass to the analytic description of tangent vectors as directional derivatives. To a vector ν = (vi,...,vn) £ Rn we may associate the differential operator D* = Σ *^> 1<г<п г called the directional derivative in the direction v. In the case that ν is a unit vector in Rn and /: Rn —» R is a smooth function, then the geometrical meaning of Dv(f)(x) is that it gives the rate of change of / in the direction v. We codify the notion of a differential operator in the following definition. Definition 4.6. A derivation at ρ G Mm is a real-valued operator D defined on smooth functions whose domain contains a neighborhood of ρ such that, for every pair f,gof such functions, D satisfies the following two properties:
§4. Tangent Vectors, Bundles, and Fields 43 (i) (Linearity) D(af + pg) = aD(f) + pD(g) for all α, β e R, (ii) (Leibnitz formula) D(fg) = D{f)g(p) + f(p)D(g). * The space of all derivations at a point ρ clearly constitutes a real vector space, which we call the analytic tangent space of Μ at ρ and denote by Tp(M)anaiytic. Moreover, the two properties imply that D(c) = 0 for any constant function c. It is not clear at this stage that Tp(M)anaiytic is finite dimensional, although we shall soon see it is of dimension m. Later, when we have seen that this definition of the tangent space is in full agreement with the geometric definition, we shall drop the term analytic. We need several simple results before we can compare analytic and geometric vectors. The first shows that smooth maps induce linear transformations between the analytic tangent spaces at corresponding points. In a word, it says that we can differentiate smooth maps. Theorem 4.7. (a) (Derivative) If g: (Мш,х) —» (Nn,y) is a smooth map, then g induces a linear map g*x: Tx(M)anaiytic -» Ty(iV) analytic defined by (g*xD)h = D(hg). (b) (Chain Rule) If д:(Мш,х) -» (Nn,y) and f:(Nn,y) -» (Ρ^,ζ) are smooth maps, then (fg)*x = f*yg*x. Proof. Both statements are simple exercises. ■ Corollary 4.8. If f:M—*Nisa diffeomorphism in some neighborhood of χ e M, then /*x:Tx(M)anaiytic -»l/(x)(iV)anaiytic is an isomorphism. Proof. Let χ e U С Μ with U open and f:U —> f(U) a diffeomorphism. Let g be the inverse diffeomorphism. Applying the chain rule to gf = idu and fg = id/(t/), we get id = #*y/*x and id = /*x^*y. Thus /*x is an isomorphism. ■ The next result is a special case of Taylor's theorem and the result that follows it applies it to determine the derivations at a point in Euclidean space. Lemma 4.9. Iff is a smooth function defined in a neighborhood of ρ G Rn, then /(x) = /(p)+ ς (ж*-л)(|^(р)+^)|, 1<г<п ^ °Xi > where the functions a»(x) are smooth and vanish at p.
44 1. In the Ashes of the Ether: Differential Topology Proof. Write m-m=fdf{p+l{rP))dt= Σ (**-*) f f(P+t(X-P))dt. Jo dt ^n Λ dxi Now, integrating by parts, we get df Il Гl d2 f = t-^-{p + t{x-p))\ - t Σ —1-(р + Цх-р))&-р№ oxi \0 j0 1<J-<n υχίθχ3 where the ai(x) are obviously smooth and vanish at p. Ш Proposition 4.10. If Ό is an open subset o/Rn and D is a derivation at ρ e U, then d= ς D^hd dxi l<i<n f(x) = f(p)+ Σ fc-PO^iP)+«*(*)), 1<г<п г Proof. Write a/ dxi as in the lemma. Applying D to both sides yields Df = 0+ Σ D{xi-pi)(^{p) + ai{x)\+0= £ D{Xi)^{p). ■ 1<г<п \ г / 1<г<п г Corollary 4.11. dim Tp(Rn)anaiytic = n. Proof. The formula shows that Tp(Rn)anaiytiC is spanned by the derivations * ~ 9xi \p These derivations are linearly independent since we obtain Xj = 0(1 < j < n) when a putative relation Σ XiDi — 0 is applied to the function Xj. ■ Corollary 4.12. Let f:M —> N be a smooth map. Then the rank of f at ρ e Μ is the rank of /*p:Tp(M)anaiytic -» T/(p) \1* Janalytic·
§4. Tangent Vectors, Bundles, and Fields 45 Proof. This is a local question, so we may assume that Μ is an open set in Rm (with coordinates xi,... , xm) and TV is an open set in Rn (with coordinates xi,... ,xn). We show that in this case the rank of /*p is the same as the rank of the Jacobian matrix of / at p. Set Di — d/dxi and write (f*P)Di= ^2 aijDj, 1 < i < m. l<j<n Then oij = ((/*„)A)sj = Di(xj о /) = Di(fj) = (dfj/dx^p). Thus, the Jacobian matrix is, in fact, the matrix of the linear transformation /*p in the basis Д. ■ Corollary 4.13. If g: Μ —» N is a smooth map whose derivative g*x : Tp(M)anaiytic -» T^(p)(iV)anaiytic w « /mear isomorphism for a given point ρ, then g is a local diffeomorphism (i.e., ρ has an open neighborhood U such that g(U) is open and g \ U is a diffeomorphism onto its image). Proof. The isomorphism implies that Μ and TV have the same dimension m and that / has rank m at p, so the inverse function theorem applies. ■ Identity of Geometric and Analytic Vectors At this stage it is possible to compare the analytic vectors to the geometric ones. Thus, suppose we have a smooth submanifold of Euclidean space l: Mm С Rn given as the zero set of a smooth function F = (F\,..., Fn_m): Rn _> Цп-rn of rank n _ m Then Tp(M)ge0metric = ker F'(p)l Rn -> Rn~m. Now the smooth inclusion l: Мш С Rn induces a linear map L*p'. lpylVl Janalytic * J-p\^ /analytic· If D e Tp(Mm)analytic, then upD e Tp(Rn)analytic, and by Proposition 4.10 we can write t*pD = ^2 νί~Έ~·> wnere уг = (i*pD)xi e R. л^ ^ OX% 1<г<п We claim that [v\,... ,vn) e Tp(M)geometriC. To see this, note that Fj(L(x)) = 0 for all χ e Μ and for all j. Thus, for D e Tp(M)an^ytic, we have 0 = D(Fji) = (i»*PP) = η,ι(ρ)(ι.ρ(Ό)) = fj^p) Σ υ^ I = Σ Vi^{p) \l<i<n °Xl) l<i<n °Xl => F'(P)V = 0^υ& Tp(M)geometric·
46 1. In the Ashes of the Ether: Differential Topology This yields a linear map 1 ρyNl Janalytic y ■*■ ρ\Μ· )geometric D i-»· ((l*pD)xi, ..., (upD)xn). To show that the map is injective (and hence an isomorphism since the dimensions are the same), it suffices to show that if (l*pD)xj = 0, 1 < j < η then Df — 0 for every smooth function / defined on a neighborhood of ρ in M. Now the value of Df is not altered if we restrict / to a smaller domain about p. But if the domain is small enough, then / extends to a neighborhood of ρ in Rm, that is, / = gi for g smooth on a neighborhood of ρ inRm. Thus, Df = D{gi) = g*L{p)Up(D) = (uP(D))g = Σ v< |f-(p) = 0 since Vi — (upD)xi = 0. These remarks yield a canonical identification between Tp(Mmanalytic and Тр(М)ёеоmetric in the case Μ С Rm. The Tangent Bundle Now let us return to the tangent bundle. We want to give a definition of T{M) as a smooth manifold intrinsically associated to the smooth manifold Μ and also to show that it is a G/m(R) bundle with fiber Rm under the standard action. For each manifold Μ we define T(M)anaiytic as a set to be the disjoint union of the vector spaces TP(M), ρ G M. Denote by π the canonical projection map π: T(M)anaiytic —» Μ sending the vector ν e TP(M) to the point ρ G M. If /: Μ —» TV is a smooth map, then we define a map analytic —» T(7V)anaiytic by sending the vector ν G Tp(M) to /*pv G Tf(p)(N) such that the diagram commutes. /* ^M)a55Jtfc > ^analytic ^ / Μ > N Now we put a topology on the sets T(M)anaiytiC by modeling pieces of it on a special case of the geometric tangent bundle. We do this first for the case Μ — Ό — open set in Rm. As sets we have the identification πΝ
§4. Tangent Vectors, Bundles, and Fields 47 J- (^)analytic — ·*■ \U /geometric — сУ X rv , Now the right-hand side comes equipped with a topology, a smooth structure, and even the structure of a (trivial) smooth Gin(R) bundle. The idea is to pass this to T(M)anaiytic. Let Я = {(t/, φ)} be a smooth atlas for M. Then φ: U -» φ(ϋ) С Rm is a homeomorphism onto its image and φ*\ T({7)anaiytic —* ^(^(^OJanaiytic — <p(U) x Rm. The Topology. A set W С T(M)analytic is open ^^(1УП T(£/)anaiytic) is open in φ(ϋ) χ Rm for all (17, φ) G Λ. The Atlas for the Smooth Structure. Take Τ = {(T(£/)anaiytic,^*) I (и,<р)ел}. The Bundle. Again take <B = {(T(£/)anaiytic,^*) I (t/,<p) G Л}. Exercise 4.14.* Show that T(M)anaiytic is a smooth manifold of dimension 2ra and that (T(M)anaiytic» M, π, Rm) is a smooth GZm(R) bundle that can be canonically identified with the geometric tangent bundle T(M)geometric whenever Μ is a submanifold of Rm. □ From now on we drop the decorations analytic and geometric in reference to tangent vectors and bundles. Since the action of Gin(R) on Rn is faithful, we have available the bundle <-» G bundle correspondence, described on page 36, and so we may speak of the bundle associated to any representation of Gin(R). Exercise 4.15.* Let Μ and TV be smooth manifolds and let πΜ\ Μ χ TV —> Μ and π ν' Μ χ TV —» TV be the canonical projections. They induce maps πΜ*:Τ(Μ χ TV) -» T(M) and πΝ*:Τ(Μ χ TV) -» T(TV), which we can combine into a map 7Гм* х πΝ*:Τ(Μ χ TV) -> T(M) χ T(TV). Show that this map is a diffeomorphism. [Hint: Use the atlas for Μ χ TV to reduce the question to the case where Μ and TV are open sets in Euclidean spaces.] □ Exercise 4.16.* Continuing the notation of the last exercise, we fix ρ G Μ and q e TV and define iM: M->MxTV, iN: TV-> Μ χ TV, UM: TP(M) χ Tq(N) -> TP(M), ΠΝ: ΤΡ(Μ) χ T,'(W) -> T9(JV), (ω, υ) ι—> ω, (ω, υ) ι—> υ.
48 1. In the Ashes of the Ether: Differential Topology Show that the following maps are inverse to each other: πΜ*(ρ,ς) χ πΝ*(ρ,ς) T{p,q)(MxN) 7=± Tp(M)xTq(N). Exercise 4.17. Show that a smooth manifold Μ of dimension η is ori- entable if and only if the nth exterior power \n(T(M)) is a trivial line bundle over M. [Hint: Choose a never-zero section e G Γ(λη(Τ(Κη))). If Xn(T(M)) is trivial, we may choose a never-zero section σ G Γ(λη(Τ(Μ))). Consider the charts (U,tp) such that, for each χ G Ε/, σ(χ) and β(φ(χ)) correspond up to a positive scalar factor under the isomorphism φ*χ:λη(Τχ(Μ)) ^ \η(Τφ(χ)(ΈΙη)). Show these charts yield an orientation for Μ.] □ It sometimes happens that the tangent bundle T(M) of a smooth manifold Μ may be described by a G atlas, where G is some closed subgroup of Gln(H). In this case T(M) acquires a G structure, so that it becomes a G bundle. In this case we say that the manifold Μ has a G structure. For example, every manifold has an On(R) structure (cf. [W. Boothby, 1986], p. 195). In general, if Η is a closed Lie subgroup of G and a G bundle has an atlas involving only Я, we say that the Η structure defined by such an atlas is a reduction of the structure group from G to Η. Here is a useful result generalizing part of the mean-value theorem. Proposition 4.18. Let f: Μ —> N be a smooth map of smooth manifolds. Assume f*:T(M) —» T(N) is the zero map on each tangent space. Then f is constant on each component of M. Proof. It suffices to show that for any chart (ν,ψ) on TV, / is constant in each component of the open set f~l(V)· Let p, q lie in one component of f-\V). We show that f(p) = f(q). Let σ: (7,0,1) -> (f-l(V),p,q) be a smooth path joining ρ and q. Now consider F = tyfo\ I —» Rm. By the chain rule we have Ff(t) = 0 for all i, so by the ordinary mean-value theorem, each of the m components of F is constant. Thus, F is constant, and so f{p) = ii~lF{0) = ^~lF(l) = f(q). Ш Next we generalize part of the fundamental theorem of calculus. Proposition 4.19. Let Μ and N be smooth manifolds with Μ connected and N having trivial tangent bundle, with trivialization t:T(N) —» V. Let f,g:M—>N be smooth maps satisfying
§4. Tangent Vectors, Bundles, and Fields 49 (i) f(p) = q(p) for some point ρ e M, (ii) i/* = tg* as maps T(M) -» V. Then f = g. Proof. It suffices to show that f(q) = g(q) for all q G M. But Μ is connected, so we may choose a smooth path σ:(7,0,1) —» (Μ,ρ, g), and we are reduced to demonstrating the result for the case Μ = R. Moreover, by covering σ(7) with coordinate charts and using the compactness of 7, we can write σ as a composite of finitely many paths, each one lying in a single coordinate chart. Thus we are reduced to considering the case Μ = R and N = Rm. Now, if £/* = tg* for one trivialization, then the same is true for any other trivialization, so writing fa and gi for the coordinates of / and #, we see that condition (ii) becomes dfi dgi . ~m = ~m for a11 '· Thus, fi(t) = gi(t) + Ci for some constant C*. But by (i), all the d must vanish, and so / = g. ■ Vector Fields A vector field on a smooth manifold Μ is a smooth section of the tangent bundle T(M). In particular, it is a map X associating to each point ρ Ε Μ a vector Xp G TP(M). Let us see what smoothness means in terms of local coordinates. Let (T(U),(p*) be a bundle coordinate chart corresponding to the chart (U,ip) on M. Setting V — ip(U), we have the commutative diagram of smooth maps: T(U) -^ T(V) = VxKm I I и -^ ν The horizontal arrows are diffeomorphisms that identify the left and right sides. Thus, we may write a vector field X on U in these coordinates as a map Y:V ->V x Rm, Y{x) = (χ,αι(χ),... ,am(x)). Clearly, Υ (and hence X) is smooth if and only if each am: V —» R is smooth. We may say this is an invariant way by noting that Υ smooth => Y{f) = У^аг(х)^—(x) 1S smooth for all smooth /: V —» R. Conversely,
50 1. In the Ashes of the Ether: Differential Topology Y(f) — /_\ai^x)~^—(x) smooth for all smooth /: V —» R • OXi Σ9χ' <α{χ)ητ^-{χ) = dj{x) is smooth for all j OXi => Υ smooth. Proposition 4.20. A section X ofT(M) is smooth <=$> X(f) is smooth for allfeC°°(M). Proof. => To check if X(f) is smooth, we need only check that its restriction to each chart U is smooth. But by the remarks above, X smooth implies that X(g) is smooth for all g G C°°{U). Now if / G C°°{M), then / | U G C°°(U), and so X(f) | U = X(f \ U) is smooth. Thus X(f) is smooth. <= Now assume that X(f) is smooth for all / G C°°(M). If we can show that X(g) is smooth for all g G C°°(U), then we will be done. But for each point ρ G U there is a function h G C°°(M) that is identically 1 on a neighborhood V of ρ and identically zero on a neighborhood of M-U (cf. [W. Boothby, 1986], pp. 193-195, on partitions of unity). Thus, hg G C°°(M) so X{hg) is smooth on M. But X(g) = X(hg) on V, and so X(g) is smooth on V. Since ρ was arbitrary, X(g) is smooth on all of U. Ш The mapping properties of vector fields are not very good. If φ: Μ —» TV is a smooth map and X is a vector field on Μ, then φ*(Χ) is not in general a vector field on TV. It can generally be regarded only as a section of the pullback bundle φ*(Τ(Ν)). If φ is injective, then φ*{Χ) may be regarded as a vector field on φ{Μ). The following definition is the best we can do by way of a naturality property. Definition 4.21. Let φ: Μ —» TV be a smooth injection and let X and F be vector fields on Μ and TV, respectively. We say that X and Υ are <p related if X(/<p) = (Г (/))<? for all / G C°°(TV). * We may now assemble the properties of smooth vector fields X on Μ which say that (1) X:C°°{M) -» C°°{M) is an R-linear map, (2) X(fg) = X(f)g + fX(g). Property 2 merely restates the definition of a derivation at ρ for each ρ G M. The only thing new here is that, for X smooth, X(f) is also smooth. In fact,
§4. Tangent Vectors, Bundles, and Fields 51 these properties say that the smooth vector fields on Μ are the derivations Der(M) on C°°(M). The derivations Der(M) constitute a vector space over R (of infinite dimension). But there is an additional structure, a multiplication called bracket and denoted [X, Y], which turns Der(M) into a Lie algebra.15 This operation is defined by [X, Y] — XY - YX and satisfies (1) (Bilinearity) [, ]: Der(M) χ Der(M) -» Der(M) is R-bilinear, (2) (Skew symmetry) [X,Y] = -[Г,Х], (3) (Jacobi identity) [[X,Г], Z] + [[У, Z], X] + [[Z,X], Г] = 0. These properties are all easy to show. In addition, there is the following naturality property. Lemma 4.22. Let φ: Μ —> N be a smooth injection. Let X\ and X2 be vector fields on M, and let Y\ and Y2 be vector fields on N. If Xj is φ related to Yj for j = 1,2, respectively, then [Xi,X2] and [Yi,Y2] are φ related. Proof. Since {Υ^})φ — Xj(/^), we have ([Yi,Y2]f)<p = {ΥιΥ2Ϊ)φ - (Υ2Υιί)φ = Χι{{Υ2ί)φ)-Χ2{{Υιί)φ) = Χι(Χ2(/φ))-Χ2(Χι(/φ)) = [XuX2](ftp). m We end the section by showing the local coordinate expression for the bracket of two vector fields. We have *= Σ «(*)£. y= Σ ых)± 1<г<п 1<7<тг J л Jrf. \ l dXi l dXi j dXj ' Exercise 4.23. (i) Let D e Der(M). Assume that / G C°°(M) vanishes on a neighborhood V οϊ ρ e M. Show that Df also vanishes on V. (ii) Show that every derivation D G Der(M) arises from a smooth vector field. □ See Chapter 3.2.6 for the formal definition.
52 1. In the Ashes of the Ether: Differential Topology §5. Differential Forms Let us begin with a rough and ready description of p-forms for ρ < 2. The 0-forms (with values in a finite-dimensional vector space V) on a manifold Μ are just the V-valued functions on M. The 1-forms generalize the derivatives of functions on M. The 2-forms are used as a way of formalizing the necessary conditions on a 1-form for it to be the derivative of a function. Now let us give the formal definition of a 1-form. Definition 5.1. Let Μ be a smooth manifold and V be a finite-dimensional vector space. A smooth V-valued 1-form on Μ is simply a smooth map ω:Τ(Μ) -» V that is linear on each fiber TX(M). When V = R (or dim(Vr) = 1), we speak of a smooth 1-form on M. More generally, if Ε is a flat vector bundle over M, a smooth E-valued 1-form on Μ is simply a smooth map ω:Τ(Μ) —» 2£ restricting to a linear map ωχ:Τχ(Μ) —» 2£x on each fiber of T(M). Exterior Derivative of a Smooth Function If /: Μ —» R is a smooth function defined on a smooth manifold, then a 1-form arises from differentiation as follows. The derivative of / is f*:T(M) —» T(R), and since T(R) has a canonical trivialization T(R) и R x R, (t,a&|t)~(i,a) we may write Д in the form /*χν = (f(x),f'(x)v). Then define the 1- form df:T(M) -» R by the formula d/(v) = /'(x)v for ν G TX(M). In exactly the same way, if /: Μ —» TV and Τ (TV) has a canonical trivialization T(7V) « TV χ V (for example, when TV = V is a vector space), we can define a 1-form d/ with values in the vector space V, which is well defined (relative to the given trivialization). In every case, the form df is a smooth 1-form on M, which we call the exterior derivative of /. Exercise 5.2. If we are given a trivialization Θ:Τ(Ν) —» V, then any smooth map τ: Ν —* Gl(V) determines a new trivialization by twisting the original according to ΘΤ:Τ(Ν) —» V, where θτ{ρ,ν) = (ρ,τ(ρ)θ(υ)). Show that the exterior derivative computed relative to the trivialization θ is the same as that computed relative to the trivialization θτ if and only if r is constant.16 □ This exercise shows that, in general, it is impossible to have a canonically defined exterior derivative for forms with values in a vector bundle Ε over M, since the differentiation is not independent of the coordinate changes. There is, however, an important exception to this. In the case of a flat bundle E, the coordinate changes are constant; in this case, we do have exterior derivatives.
§5. Differential Forms 53 Proposition 5.3. If ν Ε Tp(M) and f: Μ —» R is smooth, then df(v) — v(f)- Proof. Write f*pv = α JU Applying both sides to the identity function h.K -> R, Λ(ί) = t yields v(/) = v(hf) = (f*pv)(h) = a. Thus, f*pv = v(f)lk\ f( ν ^°' ^У *^е definition of d, we get the result. ■ Definition 5.4. If / is a smooth F-valued function on Μ and X is a smooth vector field on M, we define X(f)\p — /*(XP). * Exercise 5.5. If e^ is any basis of V, and we write / = Σ Λ^, show that *(/)Ip = Σ/<*(**>)*· Q Interpretation of dx The notion of the exterior derivative of a mapping allows us to throw some light on Leibnitz's differential notation for the calculus. For example, everyone is familiar with the expression for the derivative of у — f(x) in the form ^ = f\x) or, written more symbolically as "differentials" by dy = f'{x)dx. The latter expression can be given the following interpretation. The smooth map /: R —» R given by у — f(x) has derivative Д: T(R) —» T(R), which, if written in the canonical trivialization for T(R) with coordinates (χ,ν), has the form f*xv = (/(x), f'(x)v). Thus, the exterior derivative d/ is given by dfx(v) = f'{x)v. In particular, if we regard x: R -> R as the identity function, then dx(v) — v. Hence dfx(v) — ff{x)dx{v), that is, dy = f\x)dx. This identifies the Leibnitz dy with the exterior derivative of the mapping у = f(x). The form dx has the interpretation as "the general infinitesimal change of x" in the sense that if we apply it to any "specific infinitesimal change of x" in the amount Ax 6 TX(R), we get dx(Ax) — Ax. Similarly, dy is "the general allowable infinitesimal change in у — /(χ)" in the sense that if we apply it to any "specific infinitesimal change of x" in the amount Ax G TX(R), we get dy(Ax) = ff(x)Ax for the resulting specific change in V- Interpretation of Total Derivatives Now let us look at the connection between the exterior derivative and the Leibnitz differential notation for smooth maps /:Rn —» R. Let (x,v) denote the coordinates for points in T(Rn) и Rn χ Rn. We have the ith coordinate function a^: Rn —» R, Xi(x) — X{. As in the one-variable case, there is the convenient confusion between the function
54 1. In the Ashes of the Ether: Differential Topology and the coordinate. The exterior derivative dxi'. Rn χ Rn —» R is given by dxi(v) — V{. Now write f*xv — (f(x), f'(x)v) so we see that the exterior derivative of / is given by dfx(v) = f'(x)v= Σ д^щ = Σ dx~dXi^' 1<г<п 1<г<п Suppressing the vector υ yields dfx = Σι<ΐ<η№f/dxi)dxi, which is the Leibnitz total derivative. For another discussion of the relationship between the old and new notation, see [M. Spivak, 1970], pp. 153-154. Exterior Differential Forms Now we come to the question of higher derivatives. As usual in modern differential geometry, we shall be concerned only with the skew-symmetric part of the higher derivatives. In essence, what we shall be doing is taking the partial derivatives with respect to the base (i.e., manifold) variables and skew-symmetrizing the result, thus forgetting about the part of the higher derivatives that vanish under this procedure. However, this will not be made explicit in our treatment. The part of the higher derivative that disappears has not been studied much in differential geometry since Elie Cartan showed how useful it is to consider only the skew-symmetric part, that is, the exterior derivative. The old masters did use the symmetric part, and more recently it seems to have found an application in probability theory (cf. [P.A. Meyer, 1989], [J.E. White, 1982], and [B.L. Foster, 1986]). What are we going to differentiate? We are going to differentiate V- valued p-forms on Μ to obtain V-valued (pH-l)-forms on M. These p-forms consist of the following generalizations of 1-forms. Definition 5.6. Let Μ be a smooth manifold and V be a finite-dimensional vector space. A smooth V-valuedp-form17 on Μ is a smooth map ω: Τ(Μ)Θ ... Θ T(M) —► V whose restriction to any fiber TX(M) Θ ... Θ ΤΧ(Μ) -» V is multilinear and totally skew-symmetric in its vector arguments for all χ G M. When V = R (or dim(V) = 1), we speak of a smooth p-form on M. # Although the definition is for general p, for the most part we shall need to work only with ρ < 2 or 3. Note that, by definition, Α°(Μ, V) is declared to be the V-valued functions on M. Also note that AP(M, V) = 0 for ρ > dim M. 17Just as in the case of 1-forms, we could consider forms with values in a flat vector bundle over M. Once again exterior differentiation is canonically defined (cf. Footnotes 10 and 11).
§5. Differential Forms 55 The p-forms make up an infinite-dimensional vector space denoted by AP(M,V). We set A(M,V)= 0 Ap(M,V). 0<p<m Note that a smooth map f:M —► TV induces a backward map /*: Ap(N, V) -+ Ap(M, V) defined by (f*u)x{Xi,..., Xp) = w/(l)(/Ji,.. ·, f*Xp), where Xb ..., Xp e ГЯ(М). The exterior (or wedge) product of a Vi-valued p-form ω\ with a V2- valued g-form ω2 is a Vi ® Wvalued (p -f g)-form cji Л CJ2. It is defined by18 ωι Λ £j2(vi,...,vp+g) Σ (-^VfMi)' · · · >Mp)) ® ^2K(p+i), · · · ,Mg))· σ runs over all (p,g) shuffles Exercise 5.7. Show that ωχ Λ α;2 is a (p -f g)-form. □ Exercise 5.8. Show that if /: TV —» Μ is a smooth map, then /*(ωιΑω2) = Г(и1)АГ(и2). □ Now we are going to describe the totally skew maps at a point. Let U and V be finite-dimensional vector spaces of dimension и and v, respectively. Let ei,...,en be a basis of [/, ej,...,e* be the dual basis of {/*, and /i,...,/u be a basis of V. Let the symbol /, J,... denote the sequence {гьг2,...,гр}, {ii,i2, · · · JP}, · · ·, where 1 < h < i2 < ... < ip < u, etc. We use the abbreviations e/ = {ei1,..., ег-р), e7 = (e^,..., efp), etc. Exercise 5.9. Show that (i) / Π J φ 0 =* ej Л е} = 0, (ii) JnJ = 0=>ejAe} = (-1)σε^, where If is the list of indices J, J written in increasing order, and σ is the shuffle permutation mapping {/, J} to K. □ σ is a shuffle permutation Ίϊ r < s => a(r) < a(s) whenever both r and s lie in {1,2,... ,p}, or both lie in {p + 1,... ,p + g}.
56 1. In the Ashes of the Ether: Differential Topology Lemma 5.10. The vector space of all totally skew-symmetric maps U xU χ ... χ t/ -» V ρ copies /ms dimension I j ν and has a basis given by ejfj, where I runs through all the possibilities described above and 1 < j < v. Proof. First note that e*j(ej) = Sjj (= 1 if / = J and vanishes otherwise). Now given an arbitrary totally skew map F:U χ U... x U —> V, we can express it in terms of the basis for V as F = Σ Fifi, where Fi'.U xUx ... χ U —> R are the various components of i*1 in the basis of V. Clearly, the Fi are themselves totally skew maps. Thus it suffices to consider the case V = R. The independence of the e) is easily seen, for if ^ a/ej = 0, then evaluating on ej shows aj = 0 for all J. Conversely, evaluating F — ^ F(e/)ej on any ej always gives 0, so it follows from the multilinearity of the expression that the ej span the space of totally skew maps. ■ Next we consider p-forms in the case Μ = Rm. Since T(M) = Rm χ Rm, we have the canonical identification T(M) ©...© T(M) «Rwx(Rm ©...© Rm). (p copies) (p copies) Let us consider the coordinate function X{: Rn —>· R, with dxi:T(Rn)-*R, (x,v)-+Vi. Now the dxi\ form a basis for Tx(Rn)*, and thus dx/ = dx{x A dxi2 Λ ... Λ dx{p (where / is the sequence 1 < i\ < %2 < ... < ip < m), restricts to dxj\ to give a basis for the totally skew-symmetric R-valued functions Tx(Rn)®p —» R. It follows that an R-valued p-form on Rm has the form ω = ^ ai{x)dxi, where the summation takes place over the sequences / described above and the coefficient functions ai(x) are smooth. Exercise 5.11. Show that the p-form ω = Y^ai{x)dxi on Rm is smooth <&> the coefficient functions ai(x) are all smooth. □ Exterior Differentiation of Forms Now we define the exterior derivative d:Ap{M,V) —» ЛР+1(М, V) as follows. Choose a coordinate chart ([/, (/?) on Μ and a basis Д,..., fv of F. In terms of this coordinate system, we define
§5. Differential Forms 57 d [У2aijdxifj) = ^{dajj) A dxjfj. To see that this is well defined, (i.e., independent of the choice of coordinate chart (£/, φ) and of the basis of V), we could, by brute force, try to compute the effect of changing these choices. However, it makes better sense to follow the more subtle route of showing that d, defined with respect to a coordinate system as above, satisfies several properties and that these properties characterize it. Lemma 5.12. (i) d{u\ + ω2) = du\ + du2. (ii) ά(ωι Α ω2) = άω\ Λω2 + (-l)fcu;i Λ άω2, where к = deg(cji). (iii) d(dw) = 0, or, more briefly, d2 = 0. Proof, (i) is obvious. (ii) By virtue of (i), we may assume that ωχ = f dxjep and ω2 = 9 dxjhq, where {ep} is a basis of V\ and {hq} is a basis of V2. Then cji Λ ω2 = fgdxj A dxjev 0 /i9, so d(u\ Α ω2) = d{fg)dxj A dxjep 0 hq — (9df H- / dg)dxj A dxjep 0 hq = gdf A dxj A dxjep 0 hq + / dg A dxj A dxjep 0 hq — gdf A dxi A dxjep 0 hq + ( —l)fc/ dxj A dg A dxjep 0 hq — dwi Αω2 + ( — l)fco;i Λ dw2. (iii) Again by virtue of (i) we may assume that ω = f dxjep. Then (It ■ A rlnr tp„ ят OXi άω = d(f dxiep) — df A dxjep — £_. ъ—dxi A dxjepi and so (df \ d2 f 2^ -r—dxi A dxjep 1 = 2_\ я—я—^хэ Л ^Xi Л dxjep = 0, C/Xi J OX^OX j since the terms Θχ £χ, dxj A dxi are skew symmetric in the indices г and j. ' ■ Lemma 5.13. // two vector-valued operators on forms satisfy the three properties of the previous lemma and agree on functions, then they are identical. Proof. Let d! be another operator on forms defined in an open set of Rn which satisfies the three properties of the previous lemma. To verify that d = df, it suffices to show that dw = d'u for ω = f dxjep. We have
58 1. In the Ashes of the Ether: Differential Topology d'(/ dxrep) = (d''f)dxjep + / d'{dxjep) (by ii) = {df)dxjep -f / d'{dxjep) since d' = d on functions = du + fd'(dxjep) (by i). Thus it suffices to show that d'{dxj) = 0, where G^C/ = dx^ Л б?Хг2 Л ... Л dXip = d'x^ Л d'Xi2 Л ... Л G^Xip by equality on functions. Thus, d'{dxj) — df(dfX[) = 0 inductively by (ii) and (iii). ■ Corollary 5.14. There is a unique operator d: AP(M,V) -» Av+l(M, V), defined for all ρ and for all V, that satisfies (i), (ii), and (iii) and reduces to df = Σ -Q^dxi on functions. Proof. By the lemma, the ds defined in terms of the various coordinate charts must all agree on the intersections of these coordinate charts. ■ By a procedure like that above, we could also show that our d is the unique operator d:Av(M) -> Av+l(M) satisfying (i), (ii), and (iii) and reducing to df = Σ §£-dxi on functions. By Proposition 5.3, df(X) — X(f) for F-valued functions on M. There are similar expressions for the derivative du> of a p-form ω evaluated on ρ Η-1 vector fields Xq> · · ·, Xp- The most important case for us is the case ρ = 1. We demonstrate this case and leave the rest as an exercise. Lemma 5.15. Let ω be a V-valued 1-form on the smooth manifold M, and let X and Υ be two vector fields on M. Then άω(Χ, Υ) = Χ(ω(Υ)) — Υ(ω(Χ))-ω([Χ,Υ\). Proof. It suffices to verify this formula for 1-forms of the form ω = f dg since every 1-form is locally expressible as a sum of such terms. {LHS}: άω = df A dg so du(X, Y) = df{X)dg{Y) - df(Y)dg(X) = X(f)Y(g)-Y(f)X(g); {RHS}: Χ(ω(Υ))-Υ(ω(Χ))-ω([Χ,Υ}) = X(f dg(Y)) - Y(f dg(X)) - f dg([X, Y]) = X(fY(g)) - Y(fX(g)) - f[X, Y](g) = X(f)Y(g) + fXY(g) - Y(f)X(g) - fYX(g) - f[X, Y](g) = X(f)Y(g) + f[X, Y](g) - Y(f)X(g) - f[X, Y](g) = X(f)Y(g) - Y(f)X(g). Hence, {LHS} = {RHS}. ■
§5. Differential Forms 59 Exercise 5.16.* Show that if u; is a p-form, then dw(*o,...,*p) = Σ (-1)%И^о,..-Дг,...,^р) 0<г<р 0<i<j<p where a "hat" means "omit this entry." □ Proposition 5.17 (Naturality of d). If f: Μ —> TV is a smooth map, then the following diagram commutes for all p. AP(N,V) — >AP+\N,V) f AP(M,V) f ->AP+\M,V) Proof. It suffices to check the result for maps between Euclidean spaces. The proof is by induction on p. By linearity we may assume the p-form is ω = g dxj. For ρ = 0 we have f*(dg)(X) = dg{UX) = (f,X)g = X(f*(g)) = X(gf) = d(gf)(X), and so f*(dg) = d(f*(g)).19 Now assume the result holds in dimension <p. Writing dxi = dxj Α άχκ, with |7| = ρ and \J\, \K\ < p, we have d(f*w) = d(r(gdxjAdxK)) = d(f*(gdxj)Af*(dxK)) = d(f*(gdxj)) Л Г(ахк) ± f*(gdxj) Λ d(f*(dxK)) = (f*d{gdxj)) Λ f*{dxK) ± r(gdxj) Λ f*(d(dxK)) (by induction) = f*(d(gdxj)AdxK)±0 = f*(d(gdxj AdxK)) = Γ{άω). ■ We remark that the naturality of the exterior derivative is perhaps its most important property. It says that d is a version of differentiation that is independent of the coordinate system used. Thus, it constitutes a basic ingredient in the search for "a theory of differential equations that is independent of the coordinate system used to describe them" (cf. the Introduction to Chapter 8). In fact, every differential equation may be expressed in terms of the operator d (cf., e.g., [S.S. Chern and C. Chevalley, 1952]). 9Note that f*{g) = gf (composition).
60 1. In the Ashes of the Ether: Differential Topology Change of Coefficients If φ: V\ —> V2 is a linear map of vector spaces,20 then it induces a linear map (P*:Ap(M,Vl)-+Ap(M,V2) defined by φ*(ω)(Χι,... ,XP) = φ*(ω(Χ\,... ,XP))· It is clear that such a coefficient homomorphism always commutes with d. Coefficient Multiplication In the case that the coefficient vector space V has a multiplication given by m:V <S>V —» V (i.e., a bilinear but not necessarily commutative or associative map), then we obtain a corresponding map m*: A(M,V ®V) -> A(M,V) ω ь-> га* (ω), where (ra*(u;))(Xi,... ,XP) = πι,(ω(Χι,... ,XP)). This map turns A(M, V) into an algebra with multiplication given by AP(M, V) χ Aq(M, V) -^ Ap+q(M, V®V)^ Ap+q(M, V). Exercise 5.18. d(m*(ui А ω2)) = τη*(άωι Λ ω2) Η- (—l)pra*(u;i Λ du;2). ^ Let us now consider the case where the coefficients consist of a graded associative algebra <B — 0r>o #r· As above we denote the multiplication in *B by ra: *B <g) *B —» #. Then the multiplication on forms is AP(M, <Br) χ Л9(М, %>) -^ Ap+q{M, % <g> <BS)) ^ Ap+q(M, 2r+s) and yields a bigraded associative differential algebra, as the following exercise will show. Exercise 5.19. Let Ъ be as above. (i) Show that (dm*(ui A u;2)) = m^{du\ Α ω2) -f ( —l)pra*(u;i Λ du^). (ii) Show that ra*(ra*(u;i Λ ω2) Λ α;3) = ra*(u;i Λ ra*(u;2 Λ о;з))· (iii) If <B is graded commutative (e.g., $ is an exterior algebra or a Clifford algebra), show that m*(u)p,rAuq,s) — {—l)pr+qsm^{uq^Aup^r), where ир,г е Ap(M,%), etc. □ This may also be a morphism of flat vector bundles.
§5. Differential Forms 61 A particularly interesting case of coefficient multiplication occurs when V = £J, a Lie algebra.21 We denote the product here by (ω\,ω2) ~* [ωι,α^]· In this case we obtain a differential graded Lie algebra defined by the properties given in the following exercise. Exercise 5.20.* Let ир,ид,иг G A(M,q) be forms of dimension p, q, and r, respectively. Show that the multiplication [ , ] on A(M, g) satisfies the formulas (i) d[up,uq} = [duup,uq] + (-1)ρ[ωρ,άω4], (ii) [ω4,ωρ] = (-1)">+1[ωρ,ω4], (iii) {-ΐηωρ,ω4],ωτ] + (-1)™[^ □ Note that if ω G Al(M,g), then the mapping T(M) Θ Γ(Μ) -> g sending (X,У) —» [а;(Х),а;(У)] is skew symmetric and is therefore an element of A2(M,g). The relation between [ω(Χ),ω(Υ)] and [ω,ω](Χ,Υ) is given by the following lemma. Lemma 5.21. [ωλ,ω2\{Χ,Υ) = [ω1(Χ),ω2(Υ)} + [ω2(Χ),ωι(Υ)]. In particular, [ω(Χ)ΜΥ)] = \[»Μ(Χ,Υ)· Proof. [ωλ,ω2](Χ,Υ) = (τη,(ωιΛω2))(Χ,Υ) = πι((ωιΛω2)(Χ,Υ)) = τη(ωι(Χ) ® ω2(Υ) - ωλ(Υ) ® ω2(Χ)) = [ωι(Χ),ω2(Υ)]-[ωι(Υ),ω2(Χ)\ = [ωι(Χ),ω2(Υ)] + ЫХ),иг(У)] ■ The proof of the following result has already been shown, but we state it explicitly for future reference. Proposition 5.22. Let f:M —> N be a smooth map and let φ: ί) —> g be a Lie algebra homomorphism. Then the map φ*/*: Л*(TV, ί)) —» Л*(М, д) zs α homomorphism of Lie algebras. It is the study of the deep meaning associated with the g-valued 1-forms and 2-forms on Μ—which may be regarded as elements of the graded Lie algebra A*(M,g)—which is the subject of this book. 21 See Chapter 3 for the formal definition.
62 1. In the Ashes of the Ether: Differential Topology Basic and Sernibasic Forms The terms basic and sernibasic have to do with forms on the total space of a bundle. They refer to how these forms are related to the fiber and base directions. Let F —» Ε Д Μ be a fiber bundle over the manifold Μ. Then ж*:АР(М)сАР(Е). Definition 5.23. Let ω G А*>(Е). (i) a; is basic if it lies in the image of π*. (ii) ω is sernibasic if ω(υι,... ,vn) = 0 whenever v\ is tangent to a fiber. Lemma 5.24. (i) Basic forms are sernibasic. (ii) ω is sernibasic <Φ each ρ Ε Ε has a neighborhood U on which there are basic forms ωι such that ω = Σ αιωι for some functions ai'.U —> R. Proof, (i) If ω is basic, then ω = π* τ/ for some form η on Μ. If v\ is tangent to a fiber, then π*νι = 0, so ω(υι,... ,vn) — π*η(υι,... ,vn) = η(π*υι,...,π*υη) = 0. (ii) <=: Since the forms a;» are basic, they are also sernibasic and their combination ω = Y^aiUi is also sernibasic. =>: Fix ρ G E. Choose a local trivialization φ of Ε over a coordinate neighborhood (V,x) around π(ρ) G Μ so that ^:π-1(ν) ^FxF. Write Ρ = (Pm,Pf) in this local trivialization. Then choose a coordinate neighborhood (W,y) around pf G F. Then (χ ο ψ, у о ψ) is a local coordinate system on a neighborhood ψ~ι(ν χ W) of p. Abbreviate (χ ο ψ, у о ψ) by (x,у). Write the sernibasic p-form ω as cj = ^ ajjdxj Λ cfa/j. Let 6/ and ej be the bases dual to cfo/ and dyj. Since cj is sernibasic, акь — и(рк Л еь) = 0 if L ^ 0. Thus, α; = Y^a^dxi. Ш Lemma 5.25. Lei Η -^ Ρ —> Μ be a principal Η bundle with right action Rh'. Ρ —* Ρ, /г G Я. Lei η be a p-form on P. Then ω is basic <=ϊ ω is sernibasic and right Η invariant. Proof. =>: By Lemma 5.24(i), we need only show that ω is basic =ϊ ω right Η invariant. But ω is basic =ϊ ω = π*η for some form η on Μ => R*hu = Д*тг*77 = (тгД/0*/? = 7г*77 = ω (cf. Definition 3.19). -Ф=: By Lemma 5.24(ii), we may write, locally, ω = Σαΐωί f°r some basic forms Ui and functions ai: C/ —» R. We may assume that the Ui are independent at each point. Now the Η invariance implies that
§5. Differential Forms 63 Y^diUi = R*h {^aiUiJ = ^,RhaiRhui = ^2(Rhai)^i => Rhai = di. Thus, di(p) = di(ph), so the coefficients ai are constant along the fibers and so are basic functions. Thus, ω itself is basic. ■ Some Lie Groups Here is a description of some of the most common Lie groups. These examples are linear groups, which means that each occurs as a subgroup of Gln(R) which we studied in Exercise 1.19. We also list their tangent spaces at the identity expressed as subspaces of the vector space Mn(R). This means that these tangent spaces have been translated so as to pass through the origin. They are named by the corresponding Gothic letters. The Positive General Linear Group (the identity component of Gin(R) G/+(R) = {A e G/„(R) | det A > 0}, flI+(R) = Mn(R). The Projective General Linear Group PGln(R) = Gln(R)/{\I € Gln(R) \ λ e R}, pg[„(R) = {A e M„(R) | Trace A = 0}. The Special Linear Group Sln(R) = {Ag Gln(R) | det A = 1}, sl„(R) = {A G Af„(R) | Trace A = 0}. The Projective Special Linear Group PSln(R) = Sln(R)/{XI e Sln(R) | λ e R}, ps(„(R) = {A 6 Mn(R) | Traced = 0}. The Orthogonal Group On(R) = {Ae Gln(R) | AAl = 1}, o„(R) = {Ae Mn(R) \А + А* = 0}. The Orthogonal Group of signature p, q 0P;9(R) = {A 6 Glp+q(R) | ΑΣΡ,4Α* = Σρ>9} , oPl,(R) = {A 6 Mp+g(R) | ΑΣΡί4 + Σ,ΡΛΑι = 0} ,
64 1. In the Ashes of the Ether: Differential Topology where Σρ,,= (£ Д The Special Orthogonal Group SOn(R) = {A 6 Gln(R) | AA* = l,detA= 1}, so„(R) = {A G M„(R) | A + A* = 0} = o„(R). T/ie Special Orthogonal Group of signature p, q 50Pj,(R) = {A e G/P+9(R) | ΑΣρ,,Α* = Ep,„deti4 = 1}, soP;9(R) = {A 6 MP+9(R) | ΛΣΡ;9 + ЕР;9Л' = 0}, where ΣΜ=(£ Д T/ie Lorentz Group Lnil(R) = {Л G Gin+i(R) | Л'ЕА = Σ}, l„,i(R) = {Л G M„+i(R) | Α'Σ + ΣΛ = 0}, /0 0-1 where Σ = 0 J„_i 0 \-l 0 0 The Unitary Group Un(C) = {Ae Gln(C) | AA* = /„}, _ ,t ип(с) = {ЛбМп(с)и + А* = о}, whe*eA -Α· The Special Unitary Group SUn(C) = {A<E Gln(C) | AA* = In, det A = 1}, su„(C) = {Л G Af„(C) | Л + А* =0, ТгасеЛ = 0}. ГЛе Euclidean Group Eucn(R) = |Q °) € G/„+i(R) | υ G R",Ag 50n(R)|, euc„(R) = |Y* д\ 6M„+1(R)|t;GRn,^eson(R)|. The Positive Affine Group Afft(R) = { (* ° ) G G/++1(R) 11; 6 R", Л G G/+(R)J , off+(R) = |Q ° ) G M„+1(R) | υ G R", Л G fl[„(R)J . Exercise 5.26. (i) Verify that each of these Lie groups is in fact a group and that each has the corresponding tangent space at the identity. (ii) Show that Ln,i(R) и On>i(R). □
2 Looking for the Forest in the Leaves: Foliations ... leaves, leaves overhead and underfoot and in your face and in your eyes, endless leaves on endless trees. -Ursula K. Le Guin, 1977 A foliation is, roughly speaking, a decomposition of a manifold Μ into p-dimensional submanifolds (the leaves of the foliation) that lie neatly side by side.1 The simplest case (p = 1) consists of the integral curves of a nonzero vector field, or a line field. This case is studied in §1. The higher- dimensional analog of a line field is a p-dimensional distribution, that is, a p-dimensional subspace in each tangent space fitting together smoothly with each other. These are studied in §2. Here, however, in contrast to the vector field case, when ρ > 1 we cannot always find leaves tangent to the distribution at every point. There are integrability conditions that must be satisfied in order for this to be possible. These are studied in §3, and in §4 we show that they are sufficient for the existence of the leaves. In §5 we reformulate both the notion of distribution and the integrability conditions in terms of vector-valued differential forms. In §6 we give the modern definition of foliations, and in §7 we show that the leaves near a point on a given leaf are distributed in the same fashion no matter which point on the leaf one chooses. This leads us to introduce leaf holonomy in §8, which we use to study the space of leaves of a foliation, showing in particular that if all the leaves are compact and have trivial holonomy, then г¥от the formal definition, see §6.
66 2. Looking for the Forest in the Leaves: Foliations they are all diffeomorphic and form the fibers of a smooth fiber bundle. We remark that we do not make use of §7 and §8 in the rest of the book. Although the theory of foliations may appear at first sight to be a rather abstract subject, as we shall see it is a natural and useful generalization of the ideas of elementary calculus. In this chapter we present only a brief introduction to this subject which has undergone a vigorous development in the last two decades. For more information, we recommend the excellent book by Pierre Molino ([P. Molino, 1988]). §1. Integral Curves In this section we study a special case of how a submanifold may be obtained by specifying its tangent planes. The original example of this sort of thing is the fundamental theorem of calculus, which says that given the slope of a graph in R2 over each point of an interval I = (a, b) on the x-axis, there is a unique graph, or integral curve, passing through each point over 7. Here is a picture of some of these graphs for a given slope function. On a general smooth manifold Μ, we may again obtain integral curves by arbitrarily prescribing smoothly varying tangent lines. But, in fact, in this section we will deal rather with vector fields on Μ (which do determine line fields at points of Μ where they do not vanish) and these yield parameterized integral curves. Parameterization of a leaf really only makes sense when the leaf has dimension one, so the generalization (in §2) of the work in this section to foliations consisting of higher-dimensional leaves will of necessity be only partial (but still very interesting). Given a smooth vector field X on a smooth manifold Μ, an integral curve for X through ρ is a parametrized curve σ: ((—ε, ε), 0) —» (Μ,ρ) whose tangent at the point a(t) is the vector Xa(t), that is, X, σ(ί) . ,ч def σ{τ) = a*t In the following figure we picture some of the integral curves of a vector field in R2. In this picture there is one point Ρ (a critical point) where the vector field vanishes.
§1. Integral Curves 67 In this particular case, this point is the end of two integral curves and the beginning of two others. The parametrized integral curves "slow down" as they approach such a point and "speed up" as they leave it. Of course, a particle moving under the influence of such a velocity field will never pass through a critical point even if it were to reach such a point in finite time (which cannot happen for a smooth vector field; cf. Exercise 1.18. Its velocity would be zero there, and so it would sit there forever. Our first aim in this section is to show that, locally, a smooth vector field always has an integral curve through any point ρ where it doesn't vanish. In fact, Theorem 1.2 on the linearization of vector fields given below shows even more: it shows that around any point ρ £ M, where Xp φ Ο, there is a local coordinate system ([/, ψ) such that the integral curves near ρ are given by (ψ2,...,φη) = constant. This means that locally, up to diffeomorphism, the integral curves are arranged exactly as the family of lines parallel to the χ ι axis in Rn. To show this we need the following local existence theorem from the ordinary differential equations (cf., e.g., [L. Loomis and S. Sternberg, 1968], pp. 266-275). Theorem 1.1. Let f(t,x) = (fi(t, x),..., fn(t,x)) be a smooth, Rn-valued function defined on some open set J xU o/Rx Rn? where 0 £ J. Consider the system of equations for an unknown function g = (#i,..., gn): R —» U given by 9[(t) = fi(t,gi(t),...,gn(t)), 92(t) = f2(t,gi(t),...,gn(t)), 9n(t) = fn{t,gi(t),...,gn(t)), with initial conditions g\(0) = c\, #2(0) = C2, · · ·, <7n(0) = cn. (i) (Existence and uniqueness). For any с = (ci,...,cn) £ U, there is an ε > 0 such that this system has a unique smooth solution g:(-e,e) -* U.
68 2. Looking for the Forest in the Leaves: Foliations (ii) (Smoothness in initial conditions). For any ρ £ U, there is an open set V with ρ £ V С U, an ε > 0, and a smooth map g: (—ε, ε) χ V —» U such that g(-,c): (—ε, ε) —> U is the unique smooth solution of the system with initial condition с for any с £ V. Theorem 1.2 (Linearization of Vector Fields). If X is a vector field defined on a smooth η-manifold M, then for each ρ £ Μ where Xp φ 0 it is possible to find a local coordinate system ([/, ψ) around ρ such that U is an open set of the form [—ε, ε] χ [—ε, ε] χ ... χ [—ε, ε] (α "cube") in these coordinates, ρ is at the center of the cube, and φ*{Χ) = д/дх\.2 Proof. Since this is a local result, we may assume that Μ is an open set U С Rn, ρ = 0, X is never zero on [/, and, after a change of basis in Rn, Xq — e\. Write Χ = Σ fj(x)eji where fy. U —» R for each j. Now let g: R χ Rn —» Rn (defined on some neighborhood of (0, c)) be the solution of the system (dg/dt)(t,c) = f(g(t,c)), with initial condition #(0,c) = с whose existence is guaranteed by Theorem 1.1. φ is going to be the (suitably restricted) inverse of h(xu. ..,xn) = #(#i,0, x2, · · ·, in)- In terms of /г, the conditions on g become dh/dxi = f(h) and h(0, x2, · · ·, xn) = (0, ^2, · · ·, xn), namely, /i*(ei) = X at every point and h*(ej) — ej (2 < j < n) at the origin. Thus, the Jacobian matrix for h is the identity at the origin, and so h is a local diffeomorphism on some cube centered at the origin. The inverse φ of h has the required properties. ■ Exercise 1.3 (Application to PDE's). Consider the partial differential equation , J/ , ч df , J/ _, ч ai(xW+a2(xfe +' ·'+a-(xfe =6(ж)' where χ £ С/, an open set in Rn, a(x) = (αι(χ),... ,an(x)) and b(x) are smooth functions, and a{x) is never zero on U. Show that about each point of U there is a coordinate system у — (г/χ,... ,?/n) such that in the ^-coordinates the partial differential equation assumes the form It =c^· дуг Deduce the general solution of this equation. □ 2These coordinates are sometimes called "flow box" coordinates.
§1. Integral Curves 69 Complete Vector Fields If crj:((aj,bj),Cj) —► (M,p) (j = 1,2) are two integral curves through p, then after a translation of the parameter for σ-z (say), we may assume that C\ = C2. Then by the uniqueness part of the theorem on differential equations, we see that σ\ = σ2 on (αι,&ι) Π (α2,62). Thus, we can define σ on (ab &i) U (a2, b2) to be σ^· on (a^, fy). Applying Zorn's lemma, we see that a maximal integral curve through a point ρ always exists. There are four possible cases for a maximal integral curve σ. Up to a translation of parameters, it can only have domain of one of four types: (Ο,α), (0, oo), (—oo,0), or (—00,00). Definition 1.4. An integral curve is complete if its domain is of the last type. A vector field is complete if all of its integral curves are complete, ft Exercise 1.5. Find four never-zero vector fields on the interval (0,1) whose integral curves exhibit the four possible types of domains. □ Proposition 1.6. A vector field with compact support? is complete. Proof. Suppose that X is a vector field on Μ with compact support. Let a(t) be a maximal integral curve of X, with domain (a, b). Let us assume that b < 00 and deduce a contradiction. (Showing that a = —00 is handled in a similar manner.) If X vanishes at a point σ(ίο), then clearly a(t) = σ(ίο) for all t > to, and so b = 00. Thus, we may assume that Xa(t) IS never zero for t G (a, b). It follows that σ is a curve on the compact support of X, and hence there is a sequence tk —► b such that a(tk) converges to a cluster point, у say, of the set {σ(ί) | t G (a, 6)}. Now we pass to the product manifold Μ χ R, and consider the vector field Υ = (X,d/dt) on it. We easily check that r(t) = (σ(ί), t), t e (a, 6), is an integral curve for Υ and that the projection on Μ of any integral curve for Υ is an integral curve for X. Now Υ is a never-zero field, so the linearization of vector fields (Theorem 1.1) applies to it, yielding • a neighborhood V of the point у = (?/, 6), • flowbox coordinates (x,u) G Rn x R for V in which - V is a cube of side 2ε, - ?/ corresponds to the point (0,b) (in flowbox coordinates), 3The support of a vector field on Μ is the closure of the subset of points of Μ where X does not vanish.
70 2. Looking for the Forest in the Leaves: Foliations Υ = д/дщ the integral curves of Υ have the form χ χ (—ε, ε) (in flowbox coordinates). 2ε (Ι 11 τ rweR ι/ τ У xeRn Since Нт^-*сю T(tk) — У the integral curve r must meet V, and in fact it must meet the integral curve 0 χ (—ε, ε) (given in flowbox coordinates). It follows that the integral curve r (and hence also the integral curve σ) is defined in the interval (a, b -f ε) which shows that b is not maximal, a contradiction to the finiteness of b. Ш Proposition 1.7. Let Μ be a proper submanifold of N. Let X be a complete vector field on TV such that for every χ G M, Xx G TX(M). Then X\ also complete. \m Proof. Let σ: ((—ε, ε),0) —► (Μ,ρ) be an integral curve on Μ for X\M through p. Then it is also an integral curve on TV for X through ρ Ε N. But since X is complete on TV, σ extends to an integral curve σ: (R, 0) —► (N,p) on TV for X through ρ e N. If a(t) still lies on Μ for all t e R, then we're done. Therefore we may suppose that this fails for some t > 0 (reversing the sign of X if necessary). Set t0 = sup{t e R | σ([0, t]) С Μ}. If σ(ί0) Ξ Μ, then we have a contradiction, since then, by Theorem l.l(i), for some δ > 0 we must have σ((ίο — ε, to + ε)) С Μ which conflicts with the maximality of to. If a(to) $. M, then the sequence Xk — cr(£o(l — (Vn))) *s a sequence of points on Μ which converges on TV but not on M. But Μ is proper so this is also a contradiction. ■ Exercise 1.8. Show by example that the condition that the submanifold be proper is necessary for the conclusion to hold. □ Exercise 1.9. Show that the vector fields on R2 given by X = у2{д/дх), Υ = х2(д/ду) are complete, but that X + Υ is not complete. □ Exercise 1.10. Show that any smooth vector field on a smooth manifold Μ is the sum of two complete vector fields. [Hint: you may assume that the constant function 1 on Μ may be written as the sum of two nonnegative smooth functions 1 = ρχ + p2 with the property that the support of each Pi (г = 1,2) is a disjoint union of compact sets.] □
§1. Integral Curves 71 One-Parameter Groups of Diffeornorphisrns Let us consider a smooth map φ: R χ Μ —► Μ. For convenience we shall often abbreviate φ(ί,χ) as ^(x). Definition 1.11. φ is a one-parameter group of diffeornorphisrns (or &flow) if (1) <po(ff) = я f°r aU x € M, (2) <^s(^i(x)) = (^s+i(x) for all s, ί e R, and all χ e M. * We remark that (i) <£>s(<£_s(x)) = <£>o(#) = x so that each <ps is in fact a diffeomorphism. (ii) The map R —> Diff(M) which sends ί —► φί is a group homomor- phism. Lemma 1.12. Suppose that {Ua} is an open cover of Μ and εα > 0. (i) A one-parameter group is determined by its restriction to the sets (ii) Suppose that we are given smooth maps φα: (—εα, εα) xUa —> Μ such that 1.11(1) holds and 1.11(2) holds where it makes sense, and that these maps agree along the overlaps of the domains. Then if either the cover {Ua} is finite or inf εα > 0 these maps are obtained as the restrictions of a unique one-parameter group of diffeornorphisrns. Proof, (i) Suppose that φ and θ are two one-parameter families with the same restrictions to each (—εα,εα) χ Ua- We claim that for each χ G Μ we have tpt(x) = 0t(x) for all t. Clearly this is true for t G (—εα,εα), where χ G Ua- Moreover the set 14 = {t G R | <ρ*(χ) = 0t(x)} is closed. But if t G Vx, then we have tpt(x) = 0t{x) G £//3 for some β. We then have, for all s G (-εβ,εβ), φ3+ι(χ) = φ3(ψι(χ)) = θ3(θι(χ)) = 03+t{x). Thus Vx is open. Since 0 G 14, Vx is not empty and hence 14 = R. (ii) Since a finite cover implies ε = ίηίεα > 0 we assume the latter. This implies that we have a well-defined smooth map φ:{—ε, ε) χ Μ —> Μ defined by φ\(—ε,ε)χυα = ψα\(—ε, e)xUa- Then we define ^:RxM —» Μ by φ3{χ) = Ψ™/η(χ) (the Ti-fold composition) for any η > \s/e\. Claim 1. φ is well defined. If η > \s/e\ and A; is a positive integer then Р^кЛх) = (^/kn)n(x) = (v>8/n)n(x)- In particular, if m and η are both > \s/e\, then φ"Β/η{χ) = φ^η(χ) = ^/m(4
72 2. Looking for the Forest in the Leaves: Foliations Claim 2. φ is smooth. Since φ3{χ) = φ™/η(χ) it suffices to note that φ3/η(χ) is smooth. Claim 3. φ is a one-parameter group of diffeomorphisms, φ3+ι{χ) = {φ{3+ι)/η)η{χ) = {φ3/ηΨι/η)η{^) — (iPs/n)n(iPt/n)n(x) (Since we can reorder for small parameters) = φ3{ψι{χ))· Claim 4- φ restricts to φα on (—εα,εα) χ Ua. By definition φ restricts to φα on (—ε,ε) χ Ua. But then the argument of (i) above shows φ restricts to φα on (—εα,εα) χ Ua. ■ The Generator of a One-Parameter Group of Diffeomorphisms Let us fix χ e Μ and put a(t) = ipt(x) so that σ: (R, 0) —» (M,x) is a curve on M. Now the tangent vector to this curve at χ is σ*(Α)), where D3 = §-t\t=s. We set Xx = σ*φ0). Then X is a smooth vector field on Μ, called the infinitesimal generator of φ. It tells how each point begins to move under the action of the motion ipt for "infinitesimal" t. Exercise 1.13. Verify that the infinitesimal generator is a smooth vector field. [Hint: Show that Xx(f) = (d/dt)f(tp(t,x))\t=0.] □ By Lemma 1.12 we see that for any ε > 0, knowledge of φι for \t\ < ε suffices to determine φί for all t G R. This may make it plausible that a knowledge of (pt for t infinitesimal (i.e., knowledge of the infinitesimal generator) is enough to determine the flow ipt. We develop this theme in the following proposition and its corollary. Proposition 1.14. Let <pt be a flow with the vector field X as its infinitesimal generator. The curves σ(ί) = (ft(x) are integral curves for the vector field X. Proof. We must show that Χσ(3) = a*(D3) for all s. Fix t and set a(s) = <*p{s,x) and σι (s) = ip(s,tp(t,x)), which is φ(β + t,x) by property (2) of Definition 1.11. Hence ai(s) = a(s + t). Thus,
§1. Integral Curves 73 γ д<71 s=0 ds σ*(Α)· s=t Corollary 1.15. Two one-parameter families of diffeomorphism with the same infinitesimal generator are equal. Proof. The vector field determines the integral curves, which determine the one-parameter families of diffeomorphisms. ■ Proposition 1.14 and its corollary imply that the following definition makes sense. Definition 1.16. Let φ be a flow with infinitesimal generator X. Then we say that φ is the flow generated by Χ. Φ There is a partial converse to Proposition 1.14: Proposition 1.17. Let X be a vector field on the manifold M. Assume any one of the following: (a) M is compact; or (b) X has compact support; or (c) X is complete. Then there is a unique one-parameter group of diffeomorphisms with X as generator. Proof. If such a one-parameter group exists then it is unique by the last corollary. Since (a) => (b) => (c), it is enough to prove existence assuming (c). Since X is complete, we have an integral curve σ: R, 0 —» Μ, χ through any point χ G M. We define ipt(x) = a(t). This defines ψί(χ) for all χ and £, and it remains to show that φι is a one-parameter group of diffeomorphisms. Once again, we use the trick of passing to the vector field Υ — (д/ds, X) on R χ M, and note that r{t) = F(t,s,x) = (s + t,ipt{x)) is an integral curve for Υ through the point (5, x). Since У is a never-zero field, we can linearize it on a neighborhood U x V of any point of R χ Μ as in the following figure — > zo
74 2. Looking for the Forest in the Leaves: Foliations and then the integral curves are given locally in this coordinate system by translations F(t, z0, · · ·, Zm) = (z0 + £, zi,..., zm) (where, of course, each Zj is a function of (s, x)). Prom this local description it is clear that (i) F is smooth, and (ii) F(u, F(t, s, x)) = F(t -f u, s, x) for s and £ small. Thus <£>t(x) is smooth in (χ, £). Moreover, (u + s + £, <Pu(<pt(ff))) = F(u, s + t, 4>t(x)) = F(u,F(t,s,x)) = F(u + t,s,x) = (u + s + t, φη+ι{χ)) (for s and t small), and so φ satisfies property (ii) for s and t small, and hence for all s and t. Property (i) is automatic. ■ Exercise 1.18. Let X be a smooth vector field on R with Xq = 0. Show that no integral curve can reach 0 in finite time. (Note that if X is complete, this follows from Proposition 1.17, since the one-parameter family of diffeomorphisms φι generated by X clearly satisfies ^(0) = 0 for all t. Also, since φί is a diffeomorphism, we can therefore never have ipt{y) — 0 for у ф 0.) Show that the same is true for a smooth vector field on a smooth manifold. □ §2. Distributions Now we wish to generalize part of the study of integral curves to include leaves of larger dimension. We shall formalize the idea of "prescribing the tangent planes" in the notion of a distribution. Definition 2.1. An r-dimensional distribution on Μ is a collection iZ) = {(Dp} of r-dimensional subspaces (Dp С TpM, one for each ρ G Μ, that are smooth in the sense that they may be described on any sufficiently small open set U С Μ as the span of r smooth vector fields {Χι, Χ2» --·,ΧΓ}· These vector fields themselves are called a local basis for (D on U. A distribution is said to be involutive (or in involution, or integrable) if for any point ρ Ε Μ there is a chart ([/, φ) around ρ such that the vector fields
§3. Integrability Conditions 75 form a basis of <D on U. * Definition 2.2. A connected r-dimensional submanifold TV of Μ is called an integral submanifold for the r-dimensional distribution iZ) if (Dp = TP(N) for each ρ e TV.4 * In the case of an involutive distribution, it is clear that the equations φτ+ι{ν) = Cr+i, · · ·, φη(ρ) = Cn describe an (n—r)-parameter family of r-dimensional integral submanifolds for iZ) (one for each choice of с = (cr+i,..., cn)) and that any point in U has such an integral submanifold passing through it. §3. Integrability Conditions In contrast to the case of a single vector field, not every distribution has an integral submanifold through every point. Example 3.1. Consider the two-dimensional distribution on R3 consisting of planes normal to the vector field n(x,y,z) — (г/, —χ, 1) as in the figure below. We can see directly from this figure that there are no integral surfaces through 0 for this distribution. For if there were such a surface through the 4More generally, if s < r, an s-dimensional connected submanifold AT of Μ is called an integral submanifold for *D if TpN С ίΖ)ρ for each ρ 6 N. However, we won't use this generalization in this book.
76 2. Looking for the Forest in the Leaves: Foliations origin, it would be tangent to the (x, г/)-р1апе there, but a small loop on the surface about the z-axis could never actually close up since its z-coordinate would always be increasing as we pass counterclockwise around it. ♦ Suppose N is an integral submanifold for a distribution on M. Suppose that a local basis for the distribution is given on a neighborhood of ρ Ε Ν is by Xi,..., Xn. Let i: N —> Μ be the inclusion map. Since Xj\n is г related to Xj for each j = Ι,.,.,η, we have %^[Χι\ν,Χ^\ν] = [Xi,Xj]\n- It follows that [X2-,Xj] is also tangent to N at p, that is, it lies in span{XbX2,--.,Xn} at p. Definition 3.2. The system of vector fields {Χχ, X2, ...Дп}опС/сМ is called (algebraically)5 involutive (or mtegrable) if the bracket [Xi,Xj] lies in span{Xi, X2, · · ·, Xn} for each г and j. $b As we have just seen, algebraic involutivity is a necessary condition that a system of η < m vector fields {Xi,X2,... ,Xn} on an m-manifold Μ have an integral submanifold through each point p. The existence of an integral manifold through each point is in turn a necessary condition for the distribution to be involutive. In the case of Example 3.1, a local basis for the distribution is given by X - — - — X -— x — dx dz' dy dz' so that [XbX2] = 2— £span{XbX2}. It follows that the distribution has no integral manifolds. Exercise 3.3. Show that if {Xb X2,..., Xr} and {Y\, Y2,..., Yr} are two local bases on an open set U С М for the distribution CD, then one is algebraically involutive if and only if the other is. □ The exercise implies that the algebraic involutivity is a property of the distribution itself. Hence we see that a necessary condition that a distribution (D be involutive is that it be algebraically involutive. §4. The Probenius Theorem The content of the following theorem is that the condition of algebraic involutivity, which is necessary for the involutivity of the distribution, is 5This word is bracketed because we shall omit it as unnecessary once the Frobenius theorem (Theorem 4.1), is proved.
§3. Integrability Conditions 77 also a sufficient condition. This fact allows us to dispense henceforth with the adjective algebraic in the phrase algebraically mvolutive. Theorem 4.1 (Frobenius). Let Μ be an m-dimensional manifold and let CD be an r-dimensional distribution on M. Then CD algebraically involutive <5 CD involutive. Proof. (We follow [S.S. Chern and J. Wolfson, 1981]. For another interesting proof, cf. [W. Boothby, 1986], p. 161.) <= This direction was shown above. => For r = 1 the theorem is implied by the linearization of vector fields. We proceed by induction on r > 2. The method is to use the induction hypothesis and the linearization of vector fields to successively improve a local basis of CD. The theorem is local, so we may assume Μ = U = an open subset of Rm, ρ = 0 G U, and CD is given on U by a local basis {Xi,X2,.-.,Xr}· At each step the vector fields and the open set U will change, but we will keep the same notation for them. Our goal is to produce a coordinate system χ = (χχ,..., xm) and a local basis for CD on U of the form {Χι = д/dXi, X2 = д/дх2, ·.., Xr = d/dxr}. Step 1. There is a local coordinate system χ = (χχ,..., хш) and a local basis for CD on U of the form {Χχ, X2, · · ·, Xr-i, Xr — d/dxr}. This is just the linearization of the vector field Xr. Step 2. There is a local coordinate system χ = (χι,... ,xm) and a local basis for CD on U of the form {Χχ,Χ2, ·. ·, Xr-i, Xr — d/dxr} such that {Xi,X2, · · · ,Xr-i} is in algebraic involution and Xj(xr) = 0 for 1 < j < r- 1. Set Χ'ά = Xj - Xj(xr)Xr for 1 < j < r - 1, Xr — Xr. Note that X'j(xr) = 0 for 1 < j <r- 1, X'Axt) = 1- Writing6 [X'i,Xj] = ai:J{x)Xr mod{Xi,X2,... ,X'r_x} for 1 < i, j < r - 1, 6An equation of the form A = В mod{Ci, · · ·, Cs} means that A — В lies in the span of C\,..., Cs.
78 2. Looking for the Forest in the Leaves: Foliations we see that α^·(χ) = 0 for 1 < г, j < r — 1 by evaluating both sides on xr. Thus {X[, Xf2, · · ·, Xr-i) is m algebraic involution. Step 3. There is a local coordinate system χ = (χχ,..., хш) and a local basis for CD on C/ of the form {Χχ = д/дх\, X2 = д/дх2,. · · ,Xr-i — д/дхг-ι, Xr = Х^кл<тСл9/9хл}, where the final coefficients £a, r < A < m are functions of xr, xr+x,..., xm alone. By the induction hypothesis applied to the results of step 2, we can find a coordinate system у = (г/χ,..., ym) such that span{Xx,X2,---,Xr-i} = span{d/dyi,d/dy2,... id/dyr-i}. It follows that {д/ду\,д/ду2, · · ·, д/дуг-\,д/дхг} is a local basis for the distribution and, moreover, that dxr/dyj = 0 for 1 < j < r — 1. Writing [d/dyj,d/dxr] = bj(x)d/dxr mod{d/dy\,..., d/dyr-i} for 1 < j < r — 1, we see that bj(x) = 0 for 1 < j < r — 1 by evaluating both sides on xr. Thus we may write [_ dyj' dxr = Σ c^ forl<j<r-l. Now, expressing the d/dxr in terms of the г/s, we may write — = Τ α— Putting this in the previous formula yields 1<A< W,Wa = г£ *к1& forl^^-L \<А<т J \<k<i—1 Comparing the two sides, we see that OCa/Oxj — 0 for r < A < m and 1 < j < r — 1· It follows that £r,..., £m are functions of xr,..., xm alone. 5£ep ^. There is a local coordinate system χ = (χχ,.. ·, xm) and a local basis for <D on C/ of the form {Χχ = д/дх\, X2 — д/дх2, · · · , Xr-i = д/дхг-ι, Xr = ^г<л<тСл(^г,---,^т)(9/9хл)}, where the coefficients Сл, г < Л < m, are functions of xr, xr+x,..., хш alone. Set Xj=Xj for 1 <j<r-l, Xfr = Xr- 22 Сл^л· KA<r-l
§5. The Frobenius Theorem in Terms of Differential Forms 79 Then {X[,..., Xfr} is still a local basis for Ί) and X'r has the form Xr — / J Сл(хг? · · · » xm)"o · r<A<m Step 5. There is a local coordinate system χ — (χι,... ,xm) and a local basis for £> on U of the form {Xi = д/дх\, Х2 = д/дх2, ...,ХГ = д/дхг}. The vector field Xr is a vector field on Rm_r, and by the linearization of vector fields we can find a change of the variables that makes X'r a coordinate field without affecting the other variables. This completes the proof. ■ The proof of the Frobenius theorem not only tells us that the leaves through any point exist, but also tells us something about their mutual disposition. It says that, locally, an integrable r-dimensional distribution on an m-dimensional manifold looks like the affine r-planes in Rr χ Rm_r with "second coordinate" constant. More formally, we have the next corollary. Corollary 4.2. Let Μ he an m-dimensional manifold and let <D be an in- volutive r-dimensional distribution on M. Then for each point ρ G Μ there is a local coordinate system (£/, x) about ρ such that the integral manifold through q meets U in a set containing У (я) = W Ξ U I Xj(q') = Xj(q) form-r<j< m}. Moreover, if (U,x) and (V,y) are two such coordinate systems, then the coordinate changes Φ = xy~l have the form V^l \%1 > · · · > %m)·) · · · > ^m—r\%l > · · · > %τη)·) ™m—r-f-1 \^m—r-f-1 >···>*^τη/?···> ^m\^m—r-f-1 > · · · > ^m) )· Proof. This is a simple consequence of step 5 in the proof of the theorem. §5. The Frobenius Theorem in Terms of Differential Forms For us, the most useful formulation of the Frobenius theorem is in terms of differential forms. This is the context in which a distribution is given, not in local terms as the span of smooth vector fields, but in global terms as the kernel of a V-valued 1-form.
80 2. Looking for the Forest in the Leaves: Foliations Proposition 5.1. Let ω be a smooth 1-form on Mm with values in the vector space V. Assume that η — dim ker ωχ is constant for χ G M. Then ker ωχ is a distribution. Proof. We must show that, locally, ker ωχ is spanned by η linearly independent, smooth vector fields. Choose a basis {ei,..., eq} for V, and write ω = Y^Uiei, where the Ui are smooth 1-forms. Fix a point ρ Ε Μ around which we want to find a local basis of ker ωχ. Write ωι = Σ Q>ij(x)dxj in a local coordinate system on U about ρ so that ω = Y^aij{x)dxjei, and set A(x) = (aij(x)), a q χ ra matrix. Now ω(υ) = 0 <=> 2_]aij(x)vj = 0 for all г (where Vj = dxj(v)), so dim ker ωχ = η «Φ rank(a^(x)) —m —п. Thus we may assume, possibly after changing the basis of V and reordering the Xj that the first (ra —η) χ (m — n) block of A(p) — (α^-(ρ)) is invertible. Thus, the same is true on some neighborhood of p, which we again call C/, and because of the rank condition the last (q — (m — n)) rows are linearly dependent on the first (m — n) rows. Thus, on U we clearly have ker ω = Пкег uj, with the intersection taken over the range 1 < j < m — n. Now set _ J Ui, 1 < г < m — n, \ dxi, m — η < i <m. The 77i, 1 < г < m, form a basis for the 1-forms on U. We write т^ = Y^bij(x)dxj, 1 < г < т. The vector fields Xj = Y^Cjk{x){d/dxk) dual to the щ are smooth since the coefficients are given by (bij(x)) (cjk(x)) = /, and (bij(x)) is invertible. Moreover, the smooth vector fields Xj, for ra, η < j < m, form a local basis for the distribution ker ωχ. Ш Exercise 5.2. Let Μ be a smooth manifold and ω: Τ(Μ) —► V a (smooth) trivialization of the tangent bundle. Show that for all ν G V, ω~ι{ν) is a smooth vector field on M. If ρ G M, let c(t,p,v) (defined for all t on some neighborhood of zero) denote the integral curve of ω~ι{ν) with c(0,p, v) — p. Show that c(at,p, v) = c(t,p,av). Show that the map ехрш:Т(М) —► Μ χ Μ, which sends (x,г*/) —► (х,с(1,х,о;(г*/))), is defined on some neighborhood of the zero section and is a diffeomorphism of some neighborhood U of the zero section onto a neighborhood of the diagonal in Μ χ Μ. □ Now we are in a position to ask what form the integrability conditions for a distribution assume when the distribution is given as the kernel of a V-valued 1-form ω. We have the next result which is the differential form version of Frobenius' theorem.
§6. Foliations 81 Proposition 5.3. Let ω be a smooth 1-form on Μ with values in the vector space V. Assume that η = dim ker ωχ is constant for χ G Μ, and let *D — {ker ωχ \ χ G M} be the distribution determined by ω. Then CD is integrable & du(X, Y) = 0 whenever ω(Χ) = ω(Υ) — 0. Proof. Let us choose a local basis X\,..., Xr for CD in a neighborhood U of a point of M. Then CD integrable on U <3> [Xj,Xk] G span{Xb ..., Xr} for 1 < j, к < г ^ ш([Х,,Хк]) = 0 for 1 < j, к < г. But since u(Xj) = 0, 1 < j < r, we have (cf. Lemma 1.5.15) for 1 < j,k < r MXj'Xk) = *>(**)) - Хк(и(ха))-ш([х,,хк\) = -L>([Xj,xk]), and so <D integrable on U & du(Xj,Xk) = 0 for 1 < j, к < г, which implies the result. ■ Corollary 5.4. Let ω be a smooth 1-form on Μ with values in the vector space V. Let W С V be a subspace, and assume that η = dim uj~1(W) is constant for χ G M. Then CD — {^^{W) \ χ G M} is a distribution and CD integrable <&> du(X, Y) G W whenever ω(Χ) and ω(Υ) lie in W. Proof. The distribution CD is the kernel of the smooth 1-form ω mod W taking values in V/W. The equivalence of the integrability for condition for ω mod W and the one given above for CD is clear. ■ Exercise 5.5. Let ω = (cji, ... ,ων) be a smooth 1-form on Μ with values in ΈΙν. Assume that η = dim ker ωχ is constant for χ G M, and let CD = {ker ωχ | χ G M} be the distribution determined by ω. Show that CD integrable <=$> duii G Ι (ωχ, ...,ων), where Ι (ωχ,. ..,ωυ) is the ideal in A(M) generated by cji, ..., ωυ. □ §6. Foliations The modern study of integrable distributions is called the theory of foliations. The idea is to abstract the property contained in the corollary to
82 2. Looking for the Forest in the Leaves: Foliations the Frobenius theorem. A foliation on a manifold consists of the special coordinate systems guaranteed by this corollary, which clearly determine the integral manifolds. The integral manifolds themselves are called the leaves of the foliation. Definition 6.1. Let Mm be a smooth manifold. A q-codimensional foliated atlas on Μ is an atlas Я such that, if (t/, φ), (V, ψ) G Я, then the coordinate changes Φ = ψφ~ι have the form Φ: Rm-q χ R9 -+ Rm"9 x R9, (χ,?/)^(Φ1(χ,2/),Φ2(2/)), namely, the last ^-coordinates depend only on the last q variables. ft Definition 6.2. Two foliated atlases are equivalent if their union is a foliated atlas. A foliation on Μ is an equivalence class of foliated atlases. Note that every foliated atlas is equivalent to a maximal foliated atlas, which may be identified with the foliation. As we did for the definition of a smooth structure, we shall always implicitly assume that our atlases are maximal. Given two charts ([/, φ), (V, ψ) in a foliated atlas, the coordinate change Φ = ψφ~λ sending (χ, у) ь-> (Ф1(х,у),Ф2(у)) determines the diffeomor- phism Ф2·^9 —► R9· We can use this diffeomorphism to replace the coordinate system φ on U by θ = (id χ Фг)^, also on U. The coordinate change between the charts ([/, Θ) and (V, ψ) is then the simpler diffeomorphism (x,y) -+ (Φι(ζ,2/),2/). Let Mm be a smooth manifold with a foliation. If (υ,φ) is a foliated chart and ^fe,V2):i/-Rm-«xR', then for xei/, set ©^^(OxT^fR')). Exercise 6.3. Show that Ί)χ is independent of the choice of foliated chart used to define it and determines an integrable distribution of codimension q. □ Definition 6.4. The integrable distribution guaranteed by Exercise 6.3 is called the distribution associated to the foliation. ft Definition 6.5. The leaf L through a point ρ of a foliation on Μ is defined to be the set of points on Μ which can be joined to ρ by a piecewise smooth path everywhere tangent to the distribution associated to the foliation, ft
§7. Leaf Holonomy 83 Proposition 6.6. Every leaf L of a foliation of codimenswn q on Mm is a smooth submanifold of dimension πι — q and is an integral submanifold for the associated distribution 2). Proof. Let χ G L and ([/, φ) be a chart of a foliated atlas with χ G U. Now χ e (p~1(Rm~q x y) for some у е Kq. Let W be the path component of ip~1(Rrn~q x y) containing x. Since W is connected and is an integral manifold of the distribution 2), we have χ G W С L. On the other hand, W is a flat plaque of L in the chart (ϋ,φ). Thus, L is covered by flat plaques of dimension m — q, and so L satisfies the condition of Definition 1.2.1 for a submanifold. Moreover, since L is a union of open plaques, each of which is an integral submanifold for the distribution 2), it follows that L is also an integral submanifold for 2). ■ Corollary 6.7. Every leaf of a foliation is a maximal connected integral submanifold of the associated distribution CD. Proof. It suffices to show that every connected integral manifold of Ί) containing χ lies in the leaf containing x. Suppose that TV С Μ is such a submanifold. Let L be the leaf through x. Since TV is connected, any point у £ N may be joined to χ by a smooth path on TV. Since TV is an integral submanifold for 2), this smooth path is everywhere tangent to 2), and hence y£L. Thus, TV С L. Ш Definition 6.8. A product (or trivial) foliation is one on a manifold Μ = Nq χ L for which the atlas is just the product of the atlases of the factors. * Example 6.9. If F —► Ε —► В is a smooth bundle, then the local product structure determines a foliation on E, called the vertical foliation, for which the leaves are the fibers. ♦ §7. Leaf Holonomy Now we show that for a fixed leaf X, the way that the various nearby leaves arrange themselves about L near a given point ρ G L is in fact independent of the choice of ρ G L. Of course, this arrangement does depend very much on the choice of L itself. This will lead to the notion of holonomy, which describes how the leaves near L "wind around" L. Sliding Along Leaves The basic construction is the following. Fix a leaf L and two points on it xo,ffi £ £· Next choose a continuous path σ: (7,0,1) —► (Χ,χ0»^ι)· Now at
84 2. Looking for the Forest in the Leaves: Foliations any point of the path σ, say at σ(£), we pick a chart ([/, φ), φ = (χ, у): U —► цгп-q χ R9? of the foliation so that the plaque of L in U containing σ(ί) corresponds, under <p, to Rm_9 χ 0. We call Tt = φ~λ(0 χ R9) a transversal at σ(ί). Each point of Tt determines the leaf through it, although distinct points may lie on the same leaf, as shown in the following diagram. Now the coordinate changes of the last q coordinates do not depend on the first m — q coordinates. It follows that on the overlap between two charts at σ(α) and σ(6), the coordinate change induces a diffeomorphism φ^α'·Τα —► Tb.7 Moreover, it is clear that <рс,ъЧ>ъ,а = ψο,α? Since σ(Ι) is compact, finitely many such coordinate systems suffice to cover it. By taking the composite of the corresponding finitely many diffeomorphisms, we can pass from one end of the path to the other to obtain a diffeomorphism ha:To —> T\ called the slide map as shown in the following diagram. Exercise 7.1. Show that the germ of ha:To —► T\ at σ(0) is independent of the particular coordinate systems used to define it and is even independent of the choice of σ: (/, 0,1) —► (X, xq, x\) within its homotopy class. □ Thus, we have the following theorem. 7The diffeomorphism may not be defined on all of Ta nor have all of Тъ as its image. Rather, it will be a diffeomorphism between neighborhoods of σ(α) and a(b) in Ta and Тъ, respectively. This means that it is not a diffeomorphism but the germ of a diffeomorphism. 8At least this is true on the (nonempty) domain where it makes sense.
§7. Leaf Holonomy 85 Theorem 7.2. Let σ: (1,0,1) —► (Χ,χο,χι) be a continuous path on a leaf of a foliation of Μ. Let To and T\ be two transversals of the foliation at х0 andx\, respectively. Then there exist open subsets T$ andT[ of To and T\, which are still transversals of the foliation at xq and x\, respectively, and a diffeomorphism φ: Tq —► T[ such that (i) £χ = Ap(x) for all x ^ Tq, where £χ denotes the leaf through x, (ii) φ depends only on the homotopy class of σ: (/,0,1) —► (£,xo»#i)» (iii) if σj corresponds to φ^ for j = 1,2 and σι(1) = аг(0), then the composite path σ\* σ2 (i.e., first follow σι, then follow аг) corresponds to the composite mapping ψ2°ψ\- Corollary 7.3. Let L be a closed leaf, and let Τ be a transversal. Then ΤΊΊ L is a discrete subset ofT. Proof. It suffices to show that Κ Π L is discrete for every compact set К in T. Since L is closed in Μ, it follows that A — Κ Π L is closed. If A is not discrete, then there is a point ρ Ε Κ which is an accumulation point of A. Since A is closed, ρ G A. Now by the theorem, if q G A, there is a diffeomorphism between some neighborhood of ρ in Τ and some neighborhood of q in Τ mapping points of A to points of A. Thus, q e A is also an accumulation point of A. This means that every point of A is an accumulation point. Thus, Л is a perfect set. But every perfect set is uncountable, so we see that L contains the uncountable set A, which is obviously discrete in the leaf topology. But L is paracompact, so any discrete set is at most countable,9 which is a contradiction. ■ Corollary 7.4. Every closed leaf is a regular submanifold of M. Proof. Let L be a closed leaf, pGX, and let Τ be a transversal at ρ arising from the chart ([/, φ). Since Τ Π L is a discrete subset of Γ, we can shrink the size of the chart ([/, φ) so that the new reduced transversal meets the leaf in the single point p. This chart shows that L is a regular submanifold near p. Since ρ is arbitrary, L is a regular submanifold. ■ Exercise 7.5. Let А С Rn be a closed set with every point an accumulation point. Show that A is uncountable. [Hint: Since each point ρ G A is an accumulation point, for every η > 0 there is a point q Φ ρ with \q—p\ < 1/n. Fix ρ G A and take two distinct such qs for η = 1. Then repeat the argument for each of the qs in place of ρ for η — 2. Repeat this process to construct A discrete implies each point ρ G A has an open neighborhood U with A — ρ mUc.
86 2. Looking for the Forest in the Leaves: Foliations inductively an uncountable number of convergent sequences with distinct limits.] □ Definition 7.6. Let us fix a leaf £, a point ρ G L and a transversal Τ to the foliation at p. Let G(T,p) be the group of germs diffeomorphisms of (T,p) at p. Then our construction associates to each loop λ: (I,dl) —► (L,p) a germ of a diffeomorphism <ρ(λ) G G(T,p). In fact, because of properties (ii) and (iii) of Theorem 7.3, φ induces a group homomorphism φ*: щ(£,р) —► G(T,p) called "the leaf (or transverse) holonomy." ft Exercise 7.7. Show that if Το,ΧΊ are two transversals through the end points x$,x\ G £ of a path σ: (7,0,1) —► (Χ, χο,χι), then the canonical (germ of a) diffeomorphism /:(To,xo) ~► C^b^i) induces a commutative diagram: Or □ Of course, the reason for using germs of diffeomorphisms is to have well- defined maps, since there is no guarantee how the "Poincare first return map" φ(Χ) will behave with respect to the transversal Τ on two counts. First, a loop may either increase or decrease the size of Τ (or even do something more complicated). Second, different loops may do different things. However, in the special case that the leaf has finitely generated fundamental group (in particular, if it is compact),10 there is a kind of bound on the behavior of the holonomy in the sense that we may choose transversals T\ С T2 at ρ such that for some set of generators {pi,..., Qk} of m(L,p) we have ip*(gi)T\ С T2 for 1 < г < к. If, in this special case, the holonomy is trivial, it follows easily that ip*{gi) is defined on T\ for all % and, moreover, is the identity there. It then follows that φ* (g) is defined and is the identity on T\ for all g G πι(Χ,ρ). This suggests the following. Theorem 7.8. Let Μ be a foliated manifold and let L be a compact leaf with trivial holonomy. Then there is a neighborhood U of L in Μ such that there exists a leaf-preserving diffeomorphism L χ Τ —> С/, where Τ is a transversal at ρ Ε L. Here we use the fact that the fundamental group of a compact manifold is finitely generated. Cf. [J. Dugundji, 1966].
§7. Leaf Holonomy 87 Proof. Since L is a compact leaf in M, each point of L lies in a coordinate neighborhood of the foliated atlas meeting L in just one plaque. Using the compactness of L again, we may refine this cover to a finite one consisting of foliated coordinate neighborhoods (ϋι,ψι), 1 < г < η, meeting L in the plaques {/», 1 < г < п. Since the holonomy is trivial, we may assume once and for all that, for each у G Κς, the leaf containing the plaque (^r^R771-9 χ y) depends only on y, not on i. This means we have matched up all the plaques, as in the following figure. coordinate system φ{ on Ui coordinate system PjOnUj The remaining problem is to match up all the transversals. Let us now choose smaller open sets W% С Ui that still form a cover of L and satisfy Wi С Ui. To prove the theorem it is enough to construct inductively a sequence of open sets Vk of L and smooth maps Д: Vfc x R9 —► Μ, 1 < к < η, such that (i) Ui<i<fcWicVfccUi<i<fc^, (ii) Д is a leaf-preserving diffeomorphism onto its image which gives the canonical inclusion Vk С L upon restriction to 14 x 0, and such that the leaf containing fk{Vk x y) is independent of к. Set V\ — R\V\ L, and let /i be the composite Vi χ W φ^Χά ^(Vi) xR^ftcM. Next we describe the inductive step. Thus, we suppose we are given the open set Vi in L and the smooth map fi'.Vi χ Kq —► Μ satisfying the hypotheses (i) and (ii) above for г — к and we wish to obtain them for i = A;-hi. We will do this by altering the definition of Д on (VkC\Uk+i) xΚ-ς to make it match up with ipk+i- As we shall see, in order to do this we will also have to shrink the size of the Hq factor in Vk x Κς to an open disc, but we shall continue to denote the reduced factor by R/7 since they are diffeomorphic. Set t/fc+i = C/, Wk+i = W, fk — /, and Vk — V. Choose К compact with W С К С ί/, and restrict the size of Kq so that f((Vf)K) χ Rq) С U as in the following picture.
88 2. Looking for the Forest in the Leaves: Foliations transverse directions U· m fflE XV xR*) To alter /, we require a simple technical lemma. Sublemma 7.9. Let А с В be smooth manifolds and let f:A χ Kq —► BxHq be a diffeomorphism onto its image of the form f{x,y) = (Ф(х1у)1у) which satisfies Φ (ж, 0) = χ for all χ e A. Let ρ: Α χ Hq —► R be any smooth function. Set fp(x,y) = (Ф(х,ур(х,у)),у). Then for any compact set К С A there is a neighborhood TofO in Hq such that fp: 'mt(K) χ Τ —► В xRq is a diffeomorphism onto its image. In particular, if p(x, у) = О at some point, then fp{x,y) — (x,y) at that point. Proof. Since / and fp agree along the coordinate plane у = 0 and / is an immersion there, fp is also an immersion. It follows by continuity of the Jacobian matrix that fp is also an immersion on some neighborhood of the coordinate plane у = 0 in A x Hq. But any such neighborhood contains some product neighborhood Κ χΤχ with К compact and T\ a transversal. Finally, we claim that since f p is an embedding along Κ χ 0, it must also be an embedding along some neighborhood of the form KxT\ otherwise there would be sequences of distinct points {pj}, {qj} with second coordinates tending to 0 satisfying fp(pj) = fp(qj) for all j. Passing to a subsequence, we may assume the first coordinates converge; then they must converge to the same point in K. But since fp is a local diffeomorphism, this means that we must have Pj = qj for j sufficiently large, which is a contradiction. Now we return to the proof of the theorem. Let L be any other compact set satisfying W С L С int K.
§7. Leaf Holonomy 89 We may choose p:VxRq -+R with values in [0,1] to be 1 on (V - Κ) χ R9 and 0 on L χ R9. Then we replace / by fp as in the lemma and restrict the size of R9 again, and we have the following picture. U- Щ ffi MVxR4) That is, for (x, y) G L χ R9, we have tpk+ifp(x,y) — (x,y)- Thus, we can extend fp over (W U V) x R9 with the required properties. ■ For the case of a compact leaf with trivial holonomy, we see the local triviality of the foliation near this leaf. If the leaf is noncompact, such an argument is unavailable. Nevertheless, if U is any relatively compact submanifold of a closed leaf L such that the holonomy of L vanishes on the image of πι({7) in πι(Χ), the foliation is once again trivial near this part of the leaf. Here is an application of this result which amplifies an aspect of the fundamental theorem of calculus discussed in Proposition 1.4.19. Theorem 7.10. Let Μ and N be smooth manifolds with Μ connected, and lett:T(N) —> V be a trivialization of the tangent bundle of N. Suppose that fn, f:M —> N are smooth maps satisfying (i) f*t = f*tforalln = l,2,..., (ii) lim„—с» fn(p) = f(p) for some ре М. Then linin^oo fn{x) — f(x) for all χ 6 Μ. Proof. Let χ G Μ be an arbitrary point. Since Μ is connected we can always find a smooth immersion σ: (7,0,1) —► (Μ,ρ, χ). Now by replacing fn and / by fna and fa, respectively, and ρ and χ by 0 and 1, respectively, we may assume that Μ has dimension one. Set η = f*t and note that the graphs Ln and L of fn and / are leaves in the one-dimensional (and hence integrable) foliation on 7 χ TV given by π\*η — π2*£ = 0. Clearly, each leaf passes from 0 to 1 in the first coordinate and all are homeomorphic to the interval [0,1] by projection. By the previous result, we can find a neighborhood U of L in [0,1] χ TV and a leaf-preserving diffeomorphism LxT —► [/, where Τ is a transversal at ρ G L. By (ii), for sufficiently large n, Ln С {/; so under the diffeomorphism U —> L χ T, Ln corresponds to a leaf of the form L χ и for some и еТ. Then the result is clear. ■
90 2. Looking for the Forest in the Leaves: Foliations §8. Simple Foliations In this final section we study the simple foliations, which are indeed one of the simplest kinds of foliations. Definition 8.1. A foliation on Μ is called simple if its leaves are the level sets of a submersion /: Μ —> TV. In particular, this means the level sets of / are required to be connected. Φ Our aim is to characterize a simple foliation in terms of its holonomy and its leaf space. The leaf space is obtained as follows. If Μ is a foliated manifold, then there is an equivalence relation given on it by declaring x ~ у <^ χ and у lie in the same leaf of M. Definition 8.2. The quotient space M/~ equipped with the quotient topology is called the space of leaves. We denote by π: Μ —» M/~ the canonical projection. * The characterization of simple foliations is given in the following theorem. Structure Theorem 8.3. (A) If a foliation is simple, then each leaf has trivial holonomy, and the leaf space is Hausdorff. (B) Suppose a foliation has these properties: (i) each leaf has trivial holonomy; (ii) the leaf space is Hausdorff; (iii) each leaf has a finitely generated fundamental group. Then the foliation is simple. We begin the proof of this result with a technical lemma. Lemma 8.4. Let Μ be a foliated manifold and S С М. (i) If S is a union of leaves, then so is S. (ii) 7/5 is a union of leaves, then so is int(5). (iii) Let U be an open set of M. Then the union V of all leaves meeting U is open. (iv) The canonical projection π: Μ —> M/~ is an open mapping.
§8. Simple Foliations 91 Proof, (i) Suppose the leaf L meets S. We must show that L С S. Let L = LC\S. Step 1. It suffices to show that L is open in L. L = LP\S is nonempty and closed in £, so if it is also open then it must be a component. But L is connected, so it has only one component. Thus L = L and hence L С S. Step 2. L is open in L. Choose ρ G L = Χ Π S and take a chart ([/, φ) of the foliation. Then, as we can see from the following diagram, original sequence translated sequence if Pi,P2» · · · is a sequence in S approaching p, since S is a union of leaves we can slide the whole sequence in any direction along the leaves (i.e., add some small vector to the leaf coordinates Rm_9) to get sequences on S approaching points near ρ on the leaf L. (ii) Since Μ — S is a union of leaves, so is Μ — S = Μ — int(5), and hence so is int(5). (iii) Since int^) is a union of leaves and it contains C/, it also contains V (i.e., mt(Y) D V). But int(F) С V, so int(F) = V. (iv) Let U be open in M. We must show that tt(U) is open. Now n(U) is open & n~ln(U) = V is open, and V is the union of all leaves meeting {/, which is open by part (iii). ■ Exercise 8.5. Give an example to show that if С is a closed set in M, then the union U of all leaves meeting С may not be closed. □ Proof of the structure theorem. (A) Let us first show that a foliation on Μ defined by a submersion f:M —► N has trivial holonomy. If ha:Tq и I\ is the slide map along a curve σ tangent to the foliation in M, then we claim to have the following commutative diagram.
92 2. Looking for the Forest in the Leaves: Foliations к X? N This is obviously true for transversals in a single chart; so it is also true in general, by the multiplicativity of the slide maps with respect to path multiplication. But / is injective on small transversals; thus it follows that for a closed curve, ha — id, and so the holonomy is trivial. Now let us see that the leaf space is Hausdorff. The submersion /: Μ —► TV factors through the leaf space M/~, giving a continuous bijection M/~ —► TV. Since TV is Hausdorff, it follows that M/~ is also Hausdorff. Note that, for a foliation defined by a submersion, the continuous bijection M/~ —> TV is actually a homeomorphism since submersions are open. This fact indicates how to proceed in the proof of (B): it suggests that we look for a smooth structure on the leaf space M/~ with respect to which the canonical projection is smooth and a submersion. (B) Let us assume that Μ has a foliation with trivial holonomy and Hausdorff leaf space M/~. The latter condition implies in particular that each leaf is closed. Pick a leaf L and let ρ G L. Take a chart (£/, φ) of the foliation around p. Let Τ be the transversal at ρ arising from this chart. Since π\(£) is finitely generated and the holonomy is trivial, we may assume that Τ is so small that it meets each leaf at most once. Thus π | Τ: Τ —► M/~ is injective. Now let V be a neighborhood of ρ in L. Then, for Τ and V sufficiently small, the set Τ χ V (in the product structure of the chart) is an open set in U. Since π is an open map, it follows that 7г(Т) = π (Τ χ V) is open in the leaf space. In fact, more generally it is clear that π | Τ: Τ —► M/~ is an open map and hence is a homeomorphism onto its image. This shows that M/~ is locally Euclidean of dimension equal to the codimension of the foliation. Moreover, two charts of the type we have constructed differ by the sliding maps, which are diffeomorphisms (where defined). Thus the atlas is smooth, and the projection map π: Μ —» M/~ is clearly a submersion: locally, it is just the projection map Τ xV —> T. ■ Corollary 8.6. A foliation whose leaves are all compact and all have tnvial holonomy is simple. Moreover, if the ambient manifold Μ is connected, then the projection to the leaf space π: Μ —» M/~ provides Μ with the structure of a smooth fiber bundle. Proof. Since each leaf is compact, it follows that each leaf has a finitely generated fundamental group. Thus, by the structure theorem, to see that the foliation is simple, we need only show that the leaf space is Hausdorff. Let Xj, j = 1, 2, be distinct leaves. Now, according to Theorem 7.8, we can
§8. Simple Foliations 93 find trivially foliated product neighborhoods Uj of Lj, j = 1,2, together with foliation-preserving diffeomorphisms fji Lj xR9->[/j,j = l,2. Now the compact sets Lj, j = 1,2, can be separated by open sets and, again by the compactness of the Lj, we may choose these open sets to be of the form fj{Lj χ Tj) С Uj. Clearly, these project to disjoint open sets in the leaf space separating the leaves L\ and Lz. The proof that the projection to the leaf space equips Μ with the structure of a smooth fiber bundle follows easily from the existence of the trivially foliated product neighborhoods. ■ Exercise 8.7. Consider the components of n~l(y) for all у G R, where π: R2 — (0,0) —» R is given by projection on the second factor. Show that these components are the leaves of a foliation on R2 — (0,0) such that every leaf is closed and has trivial holonomy. Show also that the leaf space is not Hausdorff. □ The phenomenon behind this example depends on the lack of "transverse completeness." The following result shows what can be done with a strong transversal completeness condition. Proposition 8.8 ([C. Ehresmann, 1961]). Let π: Μ —► TV be a submersion of connected manifolds such that, for each point ρ G N, there are complete vector fields on Μ which project to complete vector fields on N whose values at ρ span Tp(N). Then π: Μ —> N is a smooth bundle. Proof. Fix ρ G N and choose complete vector fields X\,..., Xn on Μ that project to complete vector fields Yi,...,Vn on TV and such that Y\\p, · ·. ,Yn\p form a basis of TP(N). Let φ\ and ψΐ denote the one- parameter groups generated by Xj and Yj, respectively. Since π maps the vector field Xj to the vector field Yj, clearly it maps the integral curves of Xj to those of Yj and we have π ο φ^ = ip3t ο π. Now the map (Rn,0)->(W,p) given by (ίι,...,ί„)~<°</£°···°<(ρ) is a diffeomorphism on some neighborhood U of the origin in Rn to some neighborhood of ρ in TV, and π~ι(ρ) xU -+ Μ defined by (q, tu ..., tn) ^ φ\χ ο φ\ ο · - · о φ^ (q) is also a diffeomorphism onto its image and covers the previous map. Thus, it is a bundle chart. ■
3 The Fundamental Theorem of Calculus Every time we integrate a function J f{x)dx we are concerned with a 1-form f(x)dx on an interval [a, b] with values in the Lie algebra of real numbers, and the integral is an element of the Lie group of real numbers. -common room conversation The main theme of this chapter is the discussion of a nonabelian analog of the elementary fundamental theorem of calculus. We now sketch the main ideas followed in studying this theme. The problem is to characterize the smooth maps f:M —» G, where Μ is a smooth manifold and G is a Lie group with Lie algebra g (defined in §2). It turns out that the tangent bundle of G has a canonical trivi- alization ug-T(G) —» g. This trivialization may be regarded as a 1-form on G with values in g and is called the Maurer-Cartan form. Using this form, we can reinterpret the derivative f*:T(M) —» T(G) as the composite cl>g/*:Τ(Μ) —» g, which "forgets" the images f(x) G G and remembers only the linear part of the map. This composite cjg/* = /*^g is a g-valued 1-form on Μ called the Darboux derivative (cf. §5). It turns out that the Darboux derivative determines the map / up to translation by a constant element of G, the analog of the constant of integration of elementary calculus. The latter part of the chapter analyzes what conditions a g-valued 1-form a; on Μ must possess for it to be the Darboux derivative of some map Μ —> G. Such a map is an indefinite integral or primitive of ω. From
96 3. The Fundamental Theorem of Calculus another point of view, it is a period mapping determined by ω.1 There are two conditions for the existence of a primitive, one local and one global. The local condition is an integrability condition (the structural equation), which is a special case of the integrability condition for distributions. The global condition is a monodromy condition, which is automatically satisfied for simply connected M. Taken together, this material constitutes the nonabelian fundamental theorem of calculus (Theorem 7.14) and is indeed fundamental for our later understanding of Cartan's geometries. The second theme of this chapter is the study of the elementary theory of Lie groups. In particular, we obtain a full picture of the correspondence between Lie groups and Lie algebras.2 The third theme of this chapter is the characterization of a Lie group in terms of its Maurer-Cartan form (Theorem 8.7). It is this description that forms the basis for Cartan's generalization of Klein geometries. In particular, it will allow us to classify the Cart an space forms in Chapter 5. §1. The Maurer-Cartan Form In Euclidean space, parallel translation of vectors allows us to find a canonical trivialization of the tangent bundle. This notion is so fundamental that some authors define two vectors based at different points of Rn to be the same if there is a translation carrying one of them to the other. The possibility of doing this depends on the existence of a group—in this case, the group of translations—acting smoothly and simply transitively on Rn. Left and Right Translation The circumstances mentioned above apply to any Lie group G, because the left translation Lg:G —> G given by Lg(a) = да is a diffeomorphism (with inverse Lg-\) so that the induced maps on the tangent spaces Lg-u:Tg{G)-+Te{G) are all isomorphisms of vector spaces. This yields a canonical trivialization of the tangent bundle T{G). In fact, we could equally well use right translation Rh'.G —> G given by Rh(a) = ah. It follows that there are two ways (which are generally distinct if the group is not abelian) of identifying the space of tangent vectors at any point g G G with the space of tangent vectors at the identity e. Here is an example showing how these identifications work for the general linear group. *See Remark 8.12 at the end of the chapter. 2See the end of §2 for an outline of this.
§1. The Maurer-Cartan Form 97 Example 1.1. Regarding Gin(R) as an open submanifold of the vector space Mn(R), the "geometric" interpretation of T(G/n(R)) discussed in §4 of Chapter 1 yields the identification T(G/„(R)) = GZ„(R) χ Mn(R). Thus the tangent bundle is manifestly trivial in this way, but the trivialization uses the parallel translation in the vector space Mn(R). We are going to use this trivialization to calculate the one we are really interested in, which arises from the group structure on G/n(R). For g G G/n(R), we calculate the derivative Lg*:Gln(R) x M„(R) —► G/„(R) χ M„(R) as follows. Let (α,ν) e Gin(R) x Mn(R). Then -{Lg(a + tv) - Lg(a)} = -Lg(tv) = gv. Thus Lg*(a,v) = (ga,gv). Similarly, Rg*(a,v) = (ag,vg). ♦ The Map /i*: T(G χ G) -> T(G) According to Exercise 1.4.16 we can identify the tangent bundle of G χ G with T{G) χ T(G) by means of the diffeomorphism πι* χ n2*:T(G χ G) —► T(G) x T(G). In the following proposition we use this identification to calculate the derivative of the multiplication μ-.G χ G -+ G (cf. Definition 1.1.23). Proposition 1.2. Define θ by the commutativity of the following diagram. Щ* X 7Г2* T(G x G) -^-> 7(G) x 7(G) μ* 7(G) Tften # satisfies 0((g, u) χ (ft, v)) = (gh, Rh*u + Lg*v). Proof. First note that the formula for θ makes sense in that since и G Tg(G), it follows that Rh*u G Tgh(G), and since ν G ХЦС), it follows that b<>*^ G T^(G). The spaces T(G χ G), T(G) χ T(G), and T(G) are all vector bundles, and the three maps are bundle maps that cover the maps in the following diagram. id GxG > GxG \ / /
98 3. The Fundamental Theorem of Calculus Thus, it suffices to verify the formula for 0|, the restriction of θ to an arbitrary fiber, say the fiber over (#, h) G G χ G. Since θ\ is a linear map, it must be of the form e\:Tg{G)xTh(G)-+Tgh(G), (u,v) A(u)+B(v) where the maps A and В are linear (and may vary with g and h). To calculate A and B, we shall make use of the two maps Xh'.G 9 * GxG, pg:G h t GxG. (9>h) Now Ah and pg are connected by multiplication to left and right translations by the formulas μο\Η = Rh and μορ9 = Lg. The chain rule applied to the first equation yields the following commutative diagram (where Ch'-G —> G is the constant map with value h). no 1СЦ-Х Cftf: (g,u) id* χ c/ji|c T(G) < £— T(G) x T(G) (gh, R^u) <r^ . fc, u) x (A, 0) It follows that A(u) = Rh*u. Similarly, B(y) = L9*v. Maurer-Cartan Form Let us continue our study of the trivialization of the tangent bundle of a Lie group determined by left translation. Let % — Te(G). Definition 1.3. Let G be a Lie group. Then the left-invariant Maurer- Cartan form ωg'· Τ(G) —» g is defined by ujg(v) = Lg-i*(v) for ν e Tg(G).3 3The left-invariant Maurer-Cartan form is the grandfather of all the left- invariant forms on G. Taking the pth exterior power yields a left-invariant Ap(g)- valued p-form on G, Xp(ug)' Ap(T(G)) —► Ap(g). Any left-invariant R-valued p-form may be obtained from this one by composition with some linear map λρ{&) —* R· In particular, if ρ = η = dim G, then An(g) has dimension 1 and Xu(ug) is the Haar measure on G.
§1. The Maurer-Cartan Form 99 The term left-invariant refers to the fact that uG is invariant under left translation, which may be seen as follows. Since ν G Tg(G) implies that Lh*(v) e Thg{G), we have (Lh*uG)v = uG(Lh*(v)) = L(hgyi*(Lh*(v)) = Lg-i*(v) = uG{v). To see that uG is a smooth form, we note that it may be written as a composite of smooth maps T{G) ^dGx T(G) l-3(?x T(G) ^ T(G) χ T{G) -^ T(G). (g,u) ^ 9*(9,u) ι-»· g~lx(g,u) ^ {g~l ,0)x (g,u) ^ {e,Lg-i„u) Example 1.4. G = R. The Maurer-Cartan form is exactly the form dx defined in Proposition 1.5.3. ♦ Example 1.5. G = (R+,·). Here e = 1 &ηάω(χ,ν) — Lx-\^[v) = (l/x)v — (l/x)dx(v). Thus ω = (l/x)dx, where dx is as above (but restricted to R+). ♦ Example 1.6. G = Sl = {z e С \ \z\ = 1}, so T(Sl) = {(eie,ireie) |r,0e R}. In particular, g = Te(Sl) = {(l,n) | r e R}. The left action of 51 on T(Sl) is given by S1 xT(Sl)-+T{Sl). Let us calculate the Maurer-Cartan form. We have uG{ew ,ireie) = Le-ie^eie,ireie) = (l,ir). Te(G) = Q This description of the Maurer-Cartan form is extrinsic since the group has been regarded as sitting in С = R2. For an intrinsic view, see Example 1.10. ♦ Example 1.7. G — Gln(R). The Maurer-Cartan form at a point ν G Tg(G) is (cf. Example 1.1) ω(ν) — Lg-i*(g,v) — {e,g~lv). Or, identifying the
100 3. The Fundamental Theorem of Calculus Te(G) with Mn(R), we have ω(υ) = g~lv for ν G Tg(G). The classical way of writing the Maurer-Cartan form on Gln(R) is g~ldg. This has the following meaning. The factor g~l is an abbreviation for Lg-i*. Regard g G Gln(R) as "the general point," that is, g is the identity map on G/n(R). Then dg is the identity map on the tangent bundle. If ν G Tg(Gln(R)), then g~ldg(g,v) — g~l(g,v) = (e,g~lv); namely, g~ldg is the Maurer-Cartan form. Let us write this out explicitly in the case η — 2. (The case η — 1 appears above.) Let g — (хц) so that the x^ are the coordinate functions on Gl2(R). Then dg — (dxij) and g~ldg — (xij)~l(dxij). Writing out the matrices, we have (x\\ χ\ί\ fdxn dxi2\ \^21 X22J \dx2l dx22 J = / Χ22/Δ -χι2/Δ\ fdxn dxi2\ \-x2i/A хц/А J \dx2i dx22)' where Δ = detg = хцх22 — ^12^21· In its full glory, this is _ ( [x22dx\\ - x\2dx2i)/A (x22dxi2 - xi2dx22)/A V (—^21^11 + xndx2i)/A {-x2\dxi2 - xndx22)/A Thus we have represented ω explicitly as a 2 χ 2 matrix of 1-forms defined on Gl2(R) which is constructed from the coordinate functions x^ (and their exterior derivatives) on M2(H) restricted to Gl2(R). Clearly, we can do the same for any Gln(R). Behavior of Maurer-Cartan Forms Under Hornornorphisrn Proposition 1.8. Let φ: G\ —> G2 be a homomorphism of Lie groups. Then φ*ω2 — φ^ωχ, where the left-hand side is the pullback of the Maurer- Cartan form on G2 to G\ via φ, and the right-hand side is the Maurer- Cartan form on G\ with values interpreted in g2 — Te(G2) via the derivative of φ ate,(p*e:Te(Gi)->T2(G2). Proof. Let ν G Tg(Gi). Then {φ*ω2)ν = ω2(φ*9(ν)) = Ιφ{9)-ι*(φ*9(υ)) = tp*e(Lg-i*(v)) = φ*9(ωι(υ)). All the equalities are obvious except perhaps for the third one. This equality, which depends on the fact that φ is a homomorphism, follows from the commutativity of the following diagram. Ψ Gx >G2 L8~l 1 φ 1 L(p(s)-] Gx >G2
§2. Lie Algebras 101 Corollary 1.9. Let Η be a Lie subgroup of G. Then u;# = ug\H. Example 1.10. Consider the homomorphism exp: R —» Sl given by exp(#) = егв(в e R). According to Proposition 1.8, we have exp* CJ51 = exp^o ^R1 · But we know that ωηι — άθ and that exp^g is an isomorphism. Therefore, identifying the tangent spaces at the identity of R and 51 via exp*o> we таУ write the equation simply as exp* CJ51 = άθ. This may be further reinterpreted as follows. Instead of regarding θ as the identity map on R, let us regard it as the "function" Θ: S1 —> R sending егв ι—► θ. Of course, this is a "multivalued function" with many smooth branches differing from each other by constants, integral multiples of 2π. But, even though θ itself is not a function on S1, nevertheless d9:T(Sl) —» R is a well-defined 1-form. With this interpretation of άθ, the equation reads exp* CJ51 = exp* d<9, i.e., ω3ι = άθ. ♦ Example 1.11. 50n(R). Now 50n(R) is a Lie subgroup of G/n(R). By Corollary 1.9, the Maurer-Cartan form on SOn(R) is just the restriction of the Maurer-Cartan form on Gln(R). Thus, as above, it may be written as g~ldg, and the explicit coordinate expression is the same as above. Similar statements hold for all Lie subgroups of Gln(R). ♦ §2. Lie Algebras Although the axioms for a Lie group G are quite simple, they are very strong. In this section we show how, by combining these axioms and using the various maps which they guarantee, we obtain a multiplication on the tangent space Te(G) which turns it into a (nonassociative) algebra, the Lie algebra of G. Our practice will be to designate the Lie algebra of G by g, of Я by I), and so forth. The multiplication is constructed by identifying g with the vector space of left-invariant vector fields on G which can be "multiplied" by the bracket operation. The Lie algebra g should be regarded as an infinitesimal version of the group,4 as it turns out to encode and control almost every aspect of the group G (cf. the "primer" at the end of this section). Left-Invariant Vector Fields Definition 2.1. A left-invariant vector field on the Lie group G is a vector field satisfying either of the properties in the following lemma. * Lemma 2.2. Let G be a Lie group and let X be a vector field on G. The following two properties of X are equivalent: 4Indeed, Lie called it an "infinitesimal group."
102 3. The Fundamental Theorem of Calculus (i) ug(X) is a constant (as a ^-valued function on G); (ii) Lg*Xa = Xga for all a,g G G. Proof. uG{X) is constant & La-\ + (Xa) = L^ga)-i^(Xga) \/a,g eG (applying Lga* to both sides) & Lg*(Xa) = Xga\/a,g G G. Ш Proposition 2.3. (i) The left-invariant vector fields form a vector subspace of the real vector space of all vector fields on G. (ii) The mapping ( vector space of ] I left-invariant > > g I vector fields on G) Χ ι > Xe is a linear isomorphism. (iii) If X and Υ are left-invariant vector fields, then so is [X,Y]. Proof. Part (i) is an obvious consequence of Lemma 2.2(i). For part (ii), note that the mapping is clearly linear. Also, if Xe — 0, then 0 = Lg*Xe — Xg by Lemma 2.2(H). Moreover, if ν G g, we may define Xg = Lg*v, so that Thus, X is left invariant with Xe — v, and so the map is surjective. Finally, for part (iii) we note that the left invariance of X may be rephrased by saying that X is Lg related to itself for all g G G, and the same may be said of Y. By Lemma 1.4.22, the bracket [X, У] is also Lg related to itself for all g e G; that is, [X, Y] is left invariant. ■ Definition 2.4. If ν G g, the Lie algebra of G, then v^ denotes the corresponding left-invariant vector field on G. ft
§2. Lie Algebras 103 Lie Algebras The vector space g can be given a multiplication. Two vectors u, ν G g determine two left-invariant vector fields υ) and v^ on G from which we may form the commutator [г^,г^], which, by Proposition 2.3 (iii), is also left invariant. Then [щ ν] G g is defined to be the value of the commutator at the origin. The vector space g, equipped with this multiplication, is called the Lie algebra of G. Proposition 2.5. The multiplication [ , ]:gx^->0 satisfies the following properties. (i) (Skew symmetry) [u,v] — — [v,u], Vu,v G g. (ii) (Bilinearity) [au + bv, w] — a[u, w] + b[v, w], Vtx, v, w G g, and Va, b G R. (iii) (Jacobi identity) [[u, v],w] + [[v, w],u] + [[w,u],v] = 0. Proof. These are all immediate consequences of the properties of brackets of vector fields. ■ We now give the abstract notion of a Lie algebra which does not a priori arise as the Lie algebra of a Lie group (although cf. Theorem 7.20). Definition 2.6. A Lie algebra consists of a finite-dimensional real5 vector space g together with a multiplication [ , ]:gxg->g satisfying the three properties of Proposition 2.5. A homomorphism of Lie algebras is a linear map φ: g\ —» £2 that preserves the multiplication, namely [ψ(ν),φ(υ)]Β2 =<p([u,v]gi). # Definition 2.7. A subalgebra of a Lie algebra g is a vector subspace ()Cg satisfying [fj,fj] С I). An idea/ of a Lie algebra g is a subalgebra i) с д satisfying [f),g] С I). # Example 2.8. The Lie algebra of Gln(R) is denoted by gln(R). As a vector space it is just Te(G/n(R)), which we have identified with Mn(R), the vector space of η χ η matrices. To find the multiplication, we note that if A = (aij) G Mn(R), then the left-invariant vector field corresponding to 5The restriction to vector spaces over the real field is of course unnecessary. Lie algebras over the complex numbers or even arbitrary fields are commonly considered. The case of infinite-dimensional Lie algebras has also attracted much interest.
104 3. The Fundamental Theorem of Calculus it is given by Ag = (#, gA). As a derivation of functions (writing g — (xij)), we have , _v- д A9 ~ 2^Xiiaik dxik Now let us calculate the commutator of two such vector fields corresponding to A = (aij) and В = (bij) e M„(R). We have Ed v^ д xij^jk о » / j XpqDqr о . . . OXik OXpr i,J,k P,Q,r ^ EdxPQh д ^~-у dxjj д г,j,к ^ i,J,к ^ p,q,r P,q,r Now terms in the first sum on the right vanish unless ρ — i and q — k. Similarly, terms in the second sum vanish unless ρ — i and r = j. Thus, in the first sum we may replace ρ by г and q by k. After this is done, we exchange the indices к <-» r. In the second sum we replace ρ by г and r by j. After this is done, we exchange the indices q <-» j and then replace the new q by r. Then the right-hand side becomes Σ 1 Σ x^' zJ(a^rfc ~~ bjrark) .... , dxifc г,к ^ j Clearly, this is the left-invariant vector field corresponding to the matrix А В — ΒΑ. Thus, the multiplication in the Lie algebra of matrices Mn(R) is given by the formula [A, B] = А В - ΒΑ. Φ Proposition 2.9. (i) The derivative at the identity of a homomorphism of Lie groups φ: G\ —» G<i is a homomorphism of the corresponding Lie algebras (ii) If φ: G\ —» G2 and ψ: G<i —» G3 are homomorphisms of Lie groups, then (iii) 7/ 26?: G —> G is the identity map, then so is id*e: g —> g. (iv) 7/ <£> 25 an isomorphism of Lie groups, then φ* is an isomorphism of Lie algebras. (v) If G\ is connected and φ* is an isomorphism of Lie algebras, then ker φ is a discrete central subgroup of G.
§2. Lie Algebras 105 Proof, (i) Let u, ν G (ц. By Proposition 2.3(ii) we may uniquely extend these vectors to left-invariant vector fields Χ, Υ on G\ (i.e., such that Xe — u^ ye — v). Set υ! — φ*{η) and ν' — φ*{ν) in g2 and extend these to left- invariant vector fields X' and У on G<i. Since φ is a homomorphism, it follows that (poLg = Σφ(9)οφ. Differentiating this equation yields φ*οΣ9* — L<p(g)* °Ψ*ι so tnat φ*{Χ9) = V>*(L9*u) = Ap(0)*(<P*M) - L^g^iu1) = X^}. Thus, X and X' are <p related. Similarly, У and Y' are <p related. It follows from Lemma 1.4.22 that [X, У] and [Х',У] are also (^-related. In particular, φ*([υ,,υ]) = [φ*υ,,φ*υ]. (ii) This is just the chain rule (Theorem 1.4.7b). (iii) This is obvious from the definition of the derivative. (iv) From (ii) and (iii) we see that if the inverse of φ is ψ, then the inverse of (£*e is ф*е. (v) From (i) we have φ+9 = L^g^e οφ+Β °L~}e for all g so that φ*9 is an isomorphism for all g. It follows that φ is a local diffeomorphism at each point and hence that φ~1(β) is discrete. ■ Corollary 2.10. The Lie algebra of a subgroup Η of a Lie group G is a subalgebra of the Lie algebra of G. Corollary 2.10 makes it possible to describe the structure of the Lie algebras of any subgroup of G/n(R); for example, any of the subgroups given at the end of Chapter 1. In each case the bracket operation on the corresponding Lie algebra of matrices is just the commutator [A, B] — AB - В A. One-Dimensional Lie Subgroups One-dimensional Lie groups have quite special properties and are important, as they appear in large quantities as subgroups of general Lie groups. Proposition 2.11. Let G be a Lie group with Lie algebra g, and let ν G g. Then there is a unique homomorphism of Lie groups φ: R —» G such that <p*e(d/dt) = v. Proof. By Proposition 2.3(H), there is a left-invariant vector field X corresponding to ν (i.e., so that Xe — v). Let φ: ((α, Ъ), 0) —» (G, e) be a maximal integral curve for X through zero. For с G (a, 6), we consider the function /c: (a + c, 6 + c) —> G defined by fc(t) = φ(ό)φ(ί - с). Now we have
106 3. The Fundamental Theorem of Calculus /c* dt ^V(c)* ° Ψ* t0j dt L<p(c)*(X<p(c)<p(t0-c)) - Xf(t0)- From this equation we can draw the following two conclusions. First, the equation says that fc is an integral curve for X through /(c) = (^(c), so by uniqueness f(t) = φ{ί) for all t in the common domain. But then we may patch these integral curves together to get an integral curve with a larger domain of definition than that of φ. In particular, by taking с > 0 we see that b can be made bigger, and by taking с < 0 we see that a can be made smaller. These yield contradictions unless a — — oo and b = oo. Thus φ: (R, 0) -» (G, e). Second, the equation says that ip(s)ip{i) — Lp^ip^s+fj — s) = f3(s+t) — φ(β + £), so that φ is a homomorphism. Since it is also smooth, it is a homomorphism of Lie groups. ■ Corollary 2.12. Every left-invariant vector field on a Lie group is complete. (Of course, the same is also true for right-invariant vector fields.) Proof. The integral curve through e is a homomorphism φ:ΈΙ —» G; in particular, it is defined for all parameter values. Now consider the curve Lg(p: R, 0 -> G, g. We have (ί9φ)* dt = Lg* ο φ* dt - Lg^(X^to)) - XLgip(t0) tQj so that Lgtp is an integral curve through g defined for all parameter values. Thus, the left-invariant vector field X is complete. ■ Example 2.13. Consider G = R2/Z2, where Z2 is the lattice of integer points in R2. This is a compact Lie group that is not simply connected. The universal cover is G — R2. The covering projection is π: G —» G, sending (x, 2/) ι-* (е27Ггх,е2?гг2/). It is clear that in G a one-dimensional subgroup Η is just a one-dimensional subspace of R2; in particular, they are all closed subgroups of R2. However, the image π(Η) = Η of such a subgroup (which can be any one-dimensional subgroup of G) is not necessarily closed. In fact, a necessary and sufficient condition for a subgroup π(Η) — Η to be closed is that the line Η have rational slope in the standard coordinates of R2. ♦ Example 2.14. The group G in the preceding example sits as a closed (diagonal) subgroup in the simply connected group Sls(C) via the homomorphism (егх,егу) —» diag(e2X,e22/,e~^x+2/)), so this latter group also contains one-dimensional subgroups that are not closed. ♦ Exercise 2.15.* Let Ρ be a smooth manifold, Η a Lie group, and μ: Ρ χ Я->?а smooth right action. If X G f) (the Lie algebra of Я"), we define a
§2. Lie Algebras 107 vector field X1" on Ρ by (X^)p — μ*(Ρ)β)(0,Χ). (This is the vector field that describes how each point ρ G Ρ moves under the action of "the infinitesimal element X e #.") (a) Show that in the case where Ρ — G is a Lie group with Η С G a subgroup, and μ: G χ Η —> G is the group multiplication, that X1" here agrees with that of Definition 2.4. (b) Let Q be another smooth manifold and v\ Q χ Η —» Q be another smooth right action. Suppose that φ: Ρ —> Q is Η equivariant, that is, the following diagram commutes. PxH^P φ x id i Ι φ QxH^Q Show that the vector fields on Ρ and Q corresponding to X G f) and arising from the actions of Η are φ related. ♦ A Primer on the Lie Group-Lie Algebra Correspondence In this subsection we give a preview of certain facts concerning Lie groups and algebras whose proofs are consequences of the fundamental theorem of calculus, which we will supply later in the chapter, and Ado's theorem. Given a Lie group G, the universal cover G exists. By Corollary 8.11 G has a unique6 Lie group structure such that the projection map is a homomorphism. Moreover, the kernel of this homomorphism is a discrete central subgroup of the universal cover by Proposition 2.9(v). By Proposition 2.9 this homomorphism induces an isomorphism of the corresponding Lie algebras. The simplest example of this situation is the homomorphism R-> 51 sending θ ^ βίθ. The correspondence that associates to each Lie group its Lie algebra (Proposition 2.5) and to each homomorphism of Lie groups the corresponding homomorphism of Lie algebras (Proposition 2.9) is a functor from the category of Lie groups and homomorphisms to the category of Lie algebras and homomorphisms. It turns out that this functor is onto in the sense that every abstract Lie algebra is realized as the Lie algebra of some Lie group (Proposition 7.20), and every homomorphism of Lie algebras is the derivative of some homomorphism between some realization of the algebras (Proposition 7.15). It is not quite unique. There is no uniquely determined identity element. Any point lying over the identity element of G may serve as the identity. Once this choice is made, the rest of the Lie group structure is determined.
108 3. The Fundamental Theorem of Calculus However, the functor is not injective, as we saw earlier, since a Lie group always has the same Lie algebra as its universal cover. Nevertheless, for connected Lie groups this is the only thing that can go wrong. Two connected Lie groups with the same Lie algebra have the same universal cover (Proposition 7.20). If we restrict the functor to the Lie groups G that are connected and simply connected, it becomes an equivalence of categories. For subgroups, it may be shown [H. Yamabe, 1950] that if G is a Lie group and Η is any subgroup, then there is a unique smooth structure on Η such that the inclusion map is a weak embedding, and Η is a Lie group with this smooth structure. §3. Structural Equation Let G be a Lie group with Lie algebra g. The structural equation for the Maurer-Cartan form leads to the first of two necessary conditions that a g-valued 1-form on a manifold Μ be the Darboux derivative (cf. §5) of a map Μ —► G. Let us calculate the exterior derivative of the Maurer-Cartan form uq of a Lie group G. The calculation is merely an application of the general formula άω0{Χ,Υ)=Χ{ω0{Υ)) + Υ{ω0{Χ))-ω0({Χ,Υ}), derived in Chapter 1 (Lemma 1.5.15), which holds for any 1-form and any pair of vector fields. In our case we take X and Υ to be left-invariant fields so that lug(X) and ug(Y) are constant. Then the first two terms on the right vanish. As for the last term, since X and Υ are left invariant, the bracket [Х,У] is also left invariant, and so we have ug([X,Y]) = [X-, Y]e- But [X, Y]e = [Xe, Fe], where, by definition, the bracket on the right is the Lie bracket in g. Thus the last term is —\ug{X)^g{X)]· This yields the following equation, known as the structural equation: \c!wg(X,Y) + [ujg(X),u;g(Y)}=0\ We have derived this equation for left-invariant fields X and Y. But it is a linear equation relating 2-forms, and hence it must hold for any pair of vectors u, ν G Tg(G) that are the restrictions of left-invariant vector fields on G. But this is true for arbitrary vectors и and v, since by Proposition 2.3(ii) any vector may be extended by left translation to a left-invariant vector field of G. Thus, the structural equation holds for arbitrary vector fields X and Y. The structural equation may also be written (Lemma 1.5.21) as d^G + -[ω0,ω0] =0.
§3. Structural Equation 109 The structural equation may be thought of as merely a formula for the exterior derivative of the Maurer-Cartan form. But it is much more than just that. In Theorem 8.7 we show that it provides a local characterization of a Lie group. Thus, it determines the structure of the group locally and can be regarded as a fundamental defining property of the Lie group. If G is abelian, the second term in the structural equation vanishes and the equation reads άωο — 0. Example 3.1. G — Gln(H). Let us check the structural equation using the explicit formula for the Maurer-Cartan form ug = (xij)~l(dxij). We work in the differential algebra of η χ η matrices over the ring of R-valued differential forms on Mn(R). We may write (Xij)uJG = (dXij). Taking the exterior derivative of this equation of matrices of 1-forms kills the right-hand side and leaves (dxij) AuG + (xij)dwG = 0. Now replace (dxij) by (xij)^G to get (Xij)uJG Λ UG + (Xij)duG = 0. Since (xij) is invertible, we finally get due H- ujg Л ug — 0. To compare this with the original structural equation, we write ug — (^ij) so that ug Л uG = {uij) Л (uij) = (EfcCJifc Л ukj). Now EkuikALukj(X,Y) = Ek{uik{X)wkj(Y)-Ljik(Y)u>kj{X)}. So7 (Е»Л^)(1,Г) = (uik(X))(ukj(Y)) - (uik(Y))(ukj(X)) or uG Л u;G(X, Y) = uG{X)uG(Y) - uG{Y)uG{X) = [ug(X),ujg(Y)}. 7Here (uij(X)) denotes the matrix with components Uij(X).
ПО 3. The Fundamental Theorem of Calculus §4. Adjoint Action The adjoint action is a bookkeeping device that allows us to compare the left and right translations. More precisely, it is the derivative of the conjugation action of G on itself. It will allow us (cf. Proposition 4.10) to pull back the Maurer-Cartan form via the multiplication and inversion maps μ and ι. Let us see how it arises. A Lie group G acts on itself on the left by conjugation GxG->G. (g,h) ни. Ad(g)h=ghg-1 This smooth map induces, for each g G G, the homomorphism of Lie groups Ad(g):G->G, h ι-»· ghg~l called the inner automorphism induced by g. The map Ad: G -» Aut(G) g ^ Ad(fif) is called the (left) adjoint action ofG on itself. It is not necessary to concern ourselves with the question of whether Aut(G) is a manifold (it is!), so we shall put the map Ad in the background and concentrate on the maps Ad(g). For each g G G this map is a isomorphism of Lie groups, and so by Proposition 2.9 its derivative at the identity is an isomorphism of Lie algebras. It is this "infinitesimal" version of the adjoint action that interests us. Definition 4.1. Let G be a Lie group with Lie algebra g. Define Ad(g) = Ad(g)*e G Gl(g). The (left) adjoint representation8 is the map Ad:G —» Gl(g) sending g \~* Ad(g). Equivalently, we may speak of the (left) adjoint action of G on g as the map G x g —» g sending (g, ν) ι—► Ad(g)v. * Proposition 4.2. (a) Ad(#):g —» g is an isomorphism of Lie algebras, and the map Ad: G —» Gl(g) is a homomorphism of Lie groups. (b) If X is a left-invariant vector field on G, then so is Ad(g)*X for all gcG. 8If G is a Lie group and У is a (finite-dimensional) real or complex vector space, then a (finite dimensional) representation of G on V is a homomorphism of Lie groups p:G —► Gl(V).
§4. Adjoint Action 111 (c) R*gUG — Ad(# ι)ωο· (Note that Ad(g l) acts on the values of uq and these values lie in g.) Proof, (a) This is just a special case of Proposition 2.9(i). (b) We calculate Lh*(Ad(g)*X) = Lh* о Lg* о Ед-1„(Х) — Rg-l* О Lh* ° Lg*(X) = Rg-l*(X) = Rg-i*oLg*(X) = Ad(g)*X. (c) Let ν e Th(G) so that Rg*(v) e Tgh(G). Then R^G(v) = uuG(Rg*v) = L(hg)-i*Rg*v = (Lg-i*Rg*)(Lh-i*v) = Ad(g-luG(v). Ш Example 4.3. G — G/n(R). Let us compute the adjoint action in this case. We have Ad(g) = Lg о Rg-i. Thus Ad(g)*X = Lg* о Rg-i*X. By Example 1.1, calculating the derivatives Lg* and Rg-i* shows us that Ad(#)*: M„(R) -» Mn(R) sends υ »-► gvg~l. ♦ Definition 4.4. For g a Lie algebra, (i) AutLie(g) = {Te Gl(g) I T[u,v] = [Tu,Tv], Vu,v e fl}, (ii) fl[Lie(fl) = {T e fl[(fl) I 2>,ν] - [Tu,t;] + [u,Tt;], Vu,t; G fl}. * Exercise 4.5. Show that (a) AutLie(^) is a Lie group. (b) the Lie algebra of AutUe(g) is 0iLie(0)· (c) Let g be the Lie algebra of G. Show the derivative of Ad:G —» AutLie(^) at e is the map ad:g —» д1ыеЫ defined by ad(tz)v = [tx,v]. (Hence G abelian <ίΦ Ad trivial => Ad trivial => ad trivial => [ , ] trivial.) □ Corollary 4.6. (i) The map Ad: G —» G/(g) iafces z£s values in Autue(g)· (ii) Ad(Ad(#))(ad) = adoAd(#).
112 3. The Fundamental Theorem of Calculus Proof, (i) This is just Proposition 4.2(a). (ii)By (i) wehdLveAd(g)[u,v] = [Ad(g)u, Ad(g)v], or Ad(g)[u, Ad^"1)^] = [Ad(g)u,v], so that Ad(Ad(g))(ad(u)) = ad(Ad(g)u). ■ Exercise 4.7.* Let G be a Lie group with Lie algebra g and let Η С G be a closed subgroup with Lie algebra f). Show that (a) if Η is normal in G, then f) is an ideal of g, (b) if f) is an ideal of g and if Η and G are connected, then Η is normal inG, (c) if Η is normal in G, then (Ad(tf) -/)gCi). □ Exercise 4.8.* (a) Show that Hom(A2(g/f)),g) is an Η module under the action №){v,w) = M{h№Ad{h-1)v,Aa{h-l)w), heH. (b) Show that if Η is connected, then Hom„(A2(0/f,,0) = {φ e Hom(A2(g/f)),g | &ά(μ)φ(ν,νϋ) = <^(ad(tx)v,iy) + <p(v,ad(tz)ii;)} is the submodule of if invariant elements in Hom(A2(g/f),g) under the action given in (a). (c) For the case of the Euclidean algebra g = eucn(R), t) — son(R), take the basis e» = J5», e^· = £^· — E^. Write <p G Hom(A2(g/f)), t)) as φ = Σ aijkie* A e*· (g) e^, where a^ is skew symmetric in each of the first and last pairs of indices. Show that φ e Homf,(A2(g/f)), fj) <ίΦ a^j = с for some с and ацы vanishes when {г,^} Φ {кJ}. [Hint: Treat the cases η = 2, η = 3, and η > 3 separately.] (This result will be basic for our study of constant-curvature Riemannian spaces in Chapter 6.) Show also that Hom0n(R)(A2(g/f)), f>) = Ноть(А2(д/*)), f,). □ Exercise 4.9.* Show that the adjoint representation Ad: G —» G/(g) pulls back the Maurer-Cartan form mgi(q) according to Αά*ωοΐ(Β) = ad^c)· [Note: ad(a;G) is the g tie (β)-valued form on G given by ad (u;g) ν = οά(ωο(ν)) for υ G T(G). In more detail, this reads (a,d(uG)v)w = [ug(v),w] for ν e T{G) and w e g] □
§4. Adjoint Action 113 Product and Quotient Rules Using the adjoint action of G on g, we obtain the following fundamental formulas showing how the Maurer-Cartan form behaves with respect to multiplication and inversion. Proposition 4.10. Let μ-.G χ G —> G be multiplication and l:G —> G be the inverse. Then9 (i) μ*ωα = (π^Αά'^πΙωο) + n$uG, i.e., μ*ω0{ιυ) = Аа~1(Н)(ис(тгиъи)) +ис(к2*и]) forw е Τ^λ)(ΰ χ G), (ii) l*ug = -Adcjc, i.e., l*ljg{v) = -Ad(g)uG(v) for ν e Tg{G). Proof. Recall from Proposition 1.2 the commutative diagram щ* χ 7Γ2* T(G x G) -^-> 7(G) x 7(G) 7(G) where #((#, w) χ (ft, v)) — (gh, Rh*u + Lg*v). Let г/; G T(G x G) correspond under πι* χ π2* to the element (g, u) χ (ft, г;) G T(G) χ T(G). We calculate (μ*ω0)νο = ω0(μ*νο) = uG0((g,u) x (ft, v)) = uG(gh, Rh*u + Ly*v) = L^gh)-i^{Rh^u + L^v) = Lh-\^Lg-i^Rh^u + Lh-i*Lg-i*Lg*v — Lh-i^R^Lg-\^u + Lh-\^v = Lh-i*Rh*LjG(u) +ujg(v) = Ad(ft_1)o;G(w) + cjg(v) = Ad(ft_1)o;G(^*ii;) + u;g(7t2*w) - <(Ad(ft_1)o;G)ii; + π$(α;σ)ΐί7, 9Of course, m:G χ G —* G denotes projection to the ith factor, Ad 1: G —► Aut(g) is the map sending # —► Ad(flf)-1, and π^Αά-1 acts on the values of n*UG-
114 3. The Fundamental Theorem of Calculus which verifies (i). To see how (ii) is a consequence of (i), consider the composite map G-^+GxG^GxG-^G. 9 ·-»· (9,9) ·-»· (Sf,Sf_1) ·-»· 1 Now, this composite is the constant map, so the pullback of the Maurer- Cartan form vanishes. Thus ((id χ ι)Α)*μ*ωο = 0, and so by (i), 0 = ((id χ 0Δ)*Κ(Ad(g)uG) + π*2ωα) = (πι(id x L)A)*(Ad(g)ujG) + (π2(ίά χ ήΑ)*ωα) = Ad(g)(jjG + L*ujG. Ш Corollary 4.11. Let /ι,/2: Μ -» G, and set h(x) = f1(x)f2(x)~1. Then h?uiG = Ad(f2(x)){Ku>G ~ />g}· Proof. Write h(x) as the composite M^MxMfl-H2 GxG ^ GxG -^ G. χ ~ (x,x) ~ (fl(x),f2(x)) ~ (fi(x)j2(x)-1) - fi(x)f2(x)-1 Then h*u;G = ((id x 0(/i x /2)A)*/i*u;G = ((id χ 0(/i x /2)A)*K(Ad(/2)u;G) + 7r^G) - (πχ(id χ 0(/i x /2)A)*Ad(/2)o;G + (π2(ϊά χ ι)(/ι x /2)A)*o;G = /fAd(/2)u;G + /2n*a;G = /*Ad(/2)o;G - /2*Ad(/2)o;G = Αά(/2){/Γ(ωσ)-/2*(ϋ;σ)}. ■ Exercise 4.12.* Let /1, /2: Μ —► G, and set Л(ж) = fi(x)f2(x). Show that /**u;G - Ad^*)"1)/*^ + />G. □ Exercise 4.13. Let Η be a Lie group, V a vector space, and ρ: Η —» Gi(V) a representation. Let C/ be a manifold, X a vector field on [/, and h:U —> H, f:U —>V smooth functions. Show that Xi/Kh-1)/) = P(h-l)X(f) - (Aa(h)(h*wH(X)))f. □ Exercise 4.14.* Let G be a Lie group with Lie algebra g and let Η be a Lie subgroup of G. Let θ be a g-valued 1-form on the manifold U and k: U -> H. Show that
§5. The Darboux Derivative 115 ά{Αά(^λ)θ} = Aa{k~l)de - [Аа{к~1)в, к*ин] (cf. Lemma 1.5.21 for the bracket of two g-valued 1-forms). □ The last two exercises are a good test of your ability to differentiate. Can you find a common generalization of them? We note that Exercise 4.14 is fundamental for the elementary properties of Cartan gauges in Chapter 5. §5. The Darboux Derivative Although we could have defined the Darboux derivative in §1, it is not until now that we can show anything of its usefulness. Let ( G be a Lie group, g be the Lie algebra of G, < uq be the Maurer-Cartan form on G, Μ be a smooth manifold, I /: Μ —» G be a smooth map. Definition 5.1. The (left) Darboux derivative of / is the g-valued 1-form cjf = f*(uc) on M. * The map / itself is called an integral or primitive of cj. It may also be called a period map associated to ω. It is obvious from the naturality of d that cjf satisfies the analog of the structural equation, that is, άωί(Χ,Υ) + [ωί(Χ),ωΙ(Υ)]=0. Why is this 1-form Uf called a derivative? In general, if /: Μ —» AT, then we have called the induced map f*:T(M) —» T(N) "the derivative of /." However, this map is not exactly analogous to the usual derivative for functions on R since it has the original map / built into it. In the special case when N — G, a Lie group, we may follow /* with the Maurer-Cartan form uq to obtain the map cjg/* = /*^g = ω!'·Τ(Μ) —» Τ (Ν) —» д. The precise effect of this composition is to "forget" the underlying map / and keep only the tangential information. Thus, ω/ is the precise analog of the usual derivative of functions on R. The main property of the (left) Darboux derivative ω/ is that it determines / up to left translation by a constant element of G. Theorem 5.2 (Uniqueness of the primitive). Let Μ be a connected manifold and /i, /2: Μ —» G be smooth maps such that Uf1 — Uf2. Then there is an element С G G (the constant of integration) such that /2(я) = С · fi(x) for all χ e M.
116 3. The Fundamental Theorem of Calculus Proof. Consider the map h: Μ —> G given by h(x) = f2(x)fi(x)~1. According to Corollary 4.11, h*(uc) — Ad(/i)(/2 u;g — /i^g)· Since this latter expression vanishes by assumption and h*(uc) — ^g^*, we see that h*:T(M) —» T(G) induces the zero map on each tangent space. It follows from Proposition 1.4.18 that h is the constant function. Thus, we get h(x) — С for all χ Ε Μ, where С is some fixed element of G. In particular, f2(x) = С · /i(x) for all χ e Μ. Μ Corollary 5.3. (a) Let G be a Lie group with Lie algebra g. Then Ad: G —» Gl(g) is the unique map f: G —» Gl(g) satisfying the two properties 0) f(e) = e (ii) /*o;G/(0) = ad(o;G) (b) Let g:I,0 —> G, e be a path on G. Then Ad(g): I, e —» G/(g), e г*5 i/ie development, starting at the identity, of ad(#*(u;G)) on G/(g). Proof, (a) First we note that Ad does indeed satisfy the two properties: the first is obvious and the second is Exercise 4.9. On the other hand, Theorem 5.2 guarantees the uniqueness of /. (b) We have (Ad(#))*u;G/(g) = #*Ad*u;G/(g) = g*(&d(u;G)) = 8id(g*ujG). ■ Exercise 5.4. Show that if G is a connected Lie group and φ: G —» G is a homomorphism of Lie groups that induces the identity map on the Lie algebras, then φ itself is the identity. □ Exercise 5.5. Show that the map /: G —» Gl(g) defined by f(g) — Ad(#_1) is the unique map satisfying (i) /(e) = e and (ii) /*ujgi(q) — —ad(a;G)· Q §6. The Fundamental Theorem: Local Version How can we characterize those g-valued 1-forms ω on Μ that are Darboux derivatives? In the last section we showed that a Darboux derivative satisfies the structural equation. Now we show that this necessary condition is also sufficient, at least locally. The beautiful proof of this is due to Elie Cartan. Theorem 6.1. Let G be a Lie group with Lie algebra g. Let ω be a g- valued 1-form on the smooth manifold Μ satisfying the structural equation άω + §[ω,ω] = 0. Then, for each point ρ G Μ, there is a neighborhood U of ρ and a smooth map f:U—>G such that ω \ U = Uf. Proof. The meaning of the Darboux derivative of a map / is that if we are sitting at a point ρ Ε Μ and we want to move in the direction w G TP(M),
§6. The Fundamental Theorem: Local Version 117 then, even though we don't know the image point /(p), nevertheless the vector ljp(w) G g, interpreted as a left-invariant vector field on G, tells us how the image point f(x) should move no matter where it is situated. The idea of Cartan's proof is to construct, locally, the graph of / in Μ χ G. If there is such an /, then a point (p, f(p)) on the graph can move only in certain directions if it is to remain on the graph. In fact, a tangent vector to the graph (v,w) must satisfy the condition f*(v) — w. G fip) Ρ This is equivalent to the condition ωο(/*(ν)) — ujg(w), which in turn is just the condition Uf(w) — ujg(v). That was for the case when / exists. If we are merely given the form cj, then the condition ω(w) — ujg(v) determines a distribution on Μ χ G. We shall show that this distribution is involutive and that any leaf is, locally, the graph of a possible /. Now we pass to the formal proof. Let πο· G χ Μ —» G and пм'· Gx Μ —> Μ denote the canonical projections. Let Ω = π*Μ(ω) — ttq{ug) and let <D — ker Ω be the distribution defined by the kernel of Ω. Actually, we don't yet know that it is a distribution. By Proposition 2.5.1, it suffices to show that it has constant rank. Fix a point (p, g) £ Μ χ G. We are going to show 7ГМ*(9,р) I ®(9,P): ®(9,P) -* Tp(M) is an isomorphism. In particular, this will verify that ker Ω has constant rank = dim M. If 7Гм*К^) — 0 f°r some (v,w) e £>(g,p) — {(v,w) G TP(M) χ Tg(G) | ω(υ) = ujg(w)}, we have the implications nM*(v,w) — 0 => ν — 0 =>10 ug(w) = 0 => w — 0 => (v,w) — 0, and so we see that 7rM*(£f,p) I 2) is injective. Conversely, if υ G Tp(M), then {v,Uq1{u{v))p) G 0(3,p), and so пм*(д,Р) | 0 is surjective. Now we are going to show that CD is integrable. Calculating the exterior derivative of Ω, we get (ΚΙ = άπ*Μ(ω) - άπ^(ωα) = π*Μ(άω)-πο(άω0) (ν, w) 'This follows since ω(ν) = ug(w) for (v,w) in the distribution.
118 3. The Fundamental Theorem of Calculus = π*Μ \-^[ω,ω]) ~ πο I-^g^g]) Now make the replacement π*Μ{ω) = Kq(ujg) + Ω so that11 άΩ = ~[π*0ω0,Ω] - \{ΩΧ0ω0] - ±[Ω,Ω]. Thus, с?0(Х,У) - 0 whenever Ω(Χ) = Ω(Υ) = 0. Therefore, by Proposition 2.5.3, the distribution ker Ω is integrable. Finally, we are going to construct, for each ρ G Μ, a neighborhood U of ρ in Μ and a smooth map f:U—>G such that ω | U = Uf. Let X be a leaf through the point (p, g) Ε Μ χ G. The derivative of the restriction of 7Гм to X induces the isomorphism 7гм*: ®(д,р) ~* TP(M) studied above, and so π μ | X is a local diffeomorphism of a neighborhood of (p,g) to some neighborhood U of ρ G Μ. Let F: U —» X be the inverse map. Since kmF — id(/, F must have the form F(p) = (p, /(p)), where /:{/—* G. Now ^*(Ω) = Ω^* = 0, since the image of F is tangent to the distribution on which Ω vanishes. Thus, we have 0 = ^*(Ω) = (ttmF)*u) - (ttgF)*ug = ω- ί*(ωο)- That is, ω \ U = uf. Ш Exercise 6.2. Let g be a Lie algebra and [)Cga Lie subalgebra. Let ω be a g-valued 1-form on Μ such that ω~ι(\)) is a distribution 2) on Μ (i.e., dima;~1(i}) is a constant, independent of x). Show that if άω-\-^[ω,ω] takes values in f), then 2) is integrable. □ Exercise 6.3.* Let G be a Lie group with Lie algebra g and let f) С g be a Lie subalgebra of g. Show that there is a unique connected Lie subgroup Η С G with Lie algebra I). □ §7. The Fundamental Theorem: Global Version Passing from the local existence to the global existence of the nonabelian integral requires the entry of a new idea. We assume that the reader is 11 The reader should resist the urge to put the sum of the first two terms in the following expression equal to zero. In fact, the expression [cji,cj2] is symmetric in the two 1-forms cji,cj2 (cf. Exercise 1.5.20).
§7. The Fundamental Theorem: Global Version 119 familiar with the notion of the fundamental group of a space, in the sense of algebraic topology. We shall also use the following facts. (1) For Μ a smooth manifold, every continuous map λ: (7,0,1) —» (Μ, ρ, q) is homotopic to a smooth map. (2) If λ0, λχ: (7,0,1) —» (Μ,ρ, q) are smooth maps that are continuously homotopic, then they are smoothly homotopic. Thus, for a smooth manifold Μ, the fundamental group π\ (Μ, b) can be regarded as the group of smooth homotopy classes of smooth loops on M. We are going to show that a g-valued 1-form a; on Μ that satisfies the structural equation άω + ^[ω,α;] =0 determines a homomorphism Φω:πι(Μ,6)-*σ called the monodromy representation. It will turn out that this representation is trivial if and only if ω is the Darboux derivative of some map Μ —> G. Thus, the monodromy will be the last obstruction to the global existence of a primitive for ω. Development Let ω be a g-valued 1-form on M. The dim Μ — 1 case is quite exceptional in that, for it, the structural equation is always satisfied, there being no nonzero 2-forms on a one-dimensional manifold. Thus, the local existence theorem always applies to this case. Let us assume for definiteness that Μ — I — [a, 6]. With this special assumption, we have the following global existence theorem. Theorem 7.1. Let ω be a smooth g-valued 1-form on I = [a, b]. Then there is a unique map /:(J,a)->(G,ff) with Darboux derivative ω. Proof. By Proposition 5.2, if / exists, it is unique. By the local fundamental theorem, Theorem 6.1, each point lies in an open interval on which ω is the Darboux derivative of a map into G. Since I is compact and these intervals form an open cover, it follows that I is covered by finitely many open intervals Ui, 1 < i < n, on each of which ω is the Darboux derivative of a map into G. In fact, we may assume that UiPiUj = 0 unless \i—j\ < 1 and that tG = a e Ui. Choose U Е^П C/i+i for 1 < i < n.
120 3. The Fundamental Theorem of Calculus 1 . Pf u2 Us __rV^ Un-i —I i\ Un LJ tn=a ϊΛ Ы-\ t„=b We inductively choose primitives fi\ Щ —* G such that /ι(α) = #, fi+i(ti) = fi(U) for 1 < г < η — 1. By the uniqueness of the primitives, we have /г+ι = /г on Ui Π {/г+i. Thus we may patch these maps together, setting /(0 — /г(0 for t Ε Ui to obtain a map /: 7 —» G. Clearly, /(a) = # and /><?)= ω. ■ Definition 7.2. The unique smooth map f:I —> G satisfying /(a) = g and /*(^g) = ω is called the development of ω on G along I starting at g. (It is, of course, merely a suggestive name for a primitive of ω.) ft Remark 7.3. It suffices that ω be merely piecewise smooth on 7 for there to be a unique development. It will be continuous everywhere and smooth where ω is smooth. Indeed, if a = to < t\ < · · · < tn — b is a partition of 7 and ω \ [ti-i,U] is smooth for all г — Ι,.,.,η, then we can choose developments /*:[^-ιΛ] -* G so that /x(a) = g, Λ+ι(^) = fi(U) for 1 < г < η — 1. Patching together, we set f(t) = fi(t) for t G [^г-ъ^г] to obtain a continuous map /:7 —» G that satisfies /*(cjg) = cj except when t = ti, 1 < г < rc — 1. Next we pass to a more general situation. Definition 7.4. Let 7 = [a, 6], let Μ be a manifold of arbitrary dimension, and let ω be a smooth g-valued 1-form on M. Given a piecewise smooth path σ: (7, α, b) —» (Μ,ρ, g), let σ: (7, α) —» (G, #) be the development of the g-valued 1-form σ*(ω), which means σ*(α;^) = σ*(ω). We call σ the development of ω along σ starting at g. ft Exercise 7.5.* Let 7, Μ, and ω be as in Definition 7.4. Show the following: (a) Let σ: (7, α, 6) —» (M,p,q) have development σ: (7, α, 6) —» (G, #,ft). Then (i) &σ: (7, α, 6) —» (G, &#, kh) is also a development of σ. (ii) σ_1:(7, α, 6) —» (G, ft, #) is a development of σ_1:(7,α,6) —» (M,p,q). (Here inverse refers to the reverse path rather than inverse in the group G.) (b) Set J = [6, c], and let σ: (7, α, 6) —> (Μ,ρ, q) and ρ: (7,6, с) -» (Μ, д, г) have developments σ: (7, α, b) —> (G, #, ft) and ρ: (7, α, 6) —» (G, ft, fc).
§7. The Fundamental Theorem: Global Version 121 Then σ * p: (7, a,c) —» (M,p,r) has development <7*p:(7,a,c) —> (G,M). ^ Exercise 7.6. Show that the endpoint σ(1) of the development of a smooth path is invariant under smooth reparameterization of the path. □ In general, however, the endpoint σ(1) of the development depends not just on <7, but on the whole path σ joining ρ to q. Nevertheless, we now show that when the 1-forma; on Μ satisfies the structural equation άω-\-^[ω, ω] — 0, then σ(1) depends only on the homotopy class of σ (and the starting point g). Theorem 7.7. Let Μ be a smooth manifold and let σο, o\: (7, a, b) —» (M,p,g) be smooth homotopic paths. If ω is a smooth g-valued 1-form on Μ that satisfies the structural equation, then its developments, starting at g, along σο and σ\ have the same endpoints. Proof. Let h: (7 χ 7, a x 7, b χ 7) —>· (Μ, ρ, q) be the homotopy joining σο and σχ. Οχ —*►- σ0 7x7 Now h*u is a g-valued 1-form on 7 χ 7 that satisfies the structural equation. By Theorem 6.1, each point of 7 χ 7 lies in a square neighborhood on which h*u is the Darboux derivative of a map into G. Given a point t G 7, the compact set 7 χ t is covered by finitely many of these open sets, and, by an argument like the one above for the interval, we obtain an open set of the form 7 χ С/, with U an open interval, on which h*u is the Darboux derivative of some map into G. A similar analysis allows us to fit these maps together to get, finally, a map H:I χ I —> G such that H(a,a) — g and H*(ug) — h*u. Since h(b,u) = q is constant for и G 7, it follows that h*u — H*(ug) vanishes along the right edge of 7 χ 7, and hence Η is also constant along this right edge. In particular, H{b,a) = H(b,b), and so the developments of ω along σο and σ\ both end at the same point in G. ■ Exercise 7.8. Show that Theorem 7.7 remains true for homotopic piece- wise smooth paths. [Hint: Show that a piecewise smooth path has a reparameterization that is a smooth path.] □
122 3. The Fundamental Theorem of Calculus Monodrorny As before, Μ is a smooth manifold and a; is a g-valued 1-form on Μ that satisfies the structural equation. Let I = [0,1], fix a base point b G M, and let λ: (7, dl) —» (M, 6) be a smooth closed loop.12 According to the last result, we may develop ω along λ starting at e, and the endpoint φ (Λ) = λ(1) of this development will be a point of G that, by Theorem 7.7, is independent of the specific loop λ chosen within a homotopy class. In this way the development yields a well-defined map, Φ^: πι (Μ, Ь) —» G. Definition 7.9. The map Φω:πχ(Μ,b) —» G described above is called the monodrorny representation of cj. The image of Φ^ denoted by Г С G is called the group of periods {period group), or the monodrorny group. Φ Proposition 7.10. The monodrorny representation is a group homomor- phism. Proof. The reader can easily construct a proof by referring to the following picture. Oft.!* Аз) = (V АзХ!) = Φ(λ!)Α2(1) = Φ(λί Φ(*2) In this diagram the left translation Φ(λι)λ2 is, of course, the development of λ2 starting at Φ(λι). ■ Period Group Only Defined Up to Conjugacy The fact that the fundamental group depends on the choice of base point is reflected in the fact that the period group is only defined up to conjugacy. First we recall the effect of change of base point from bo to b\ on the fundamental group of a path-connected space. If σ: (/,0,1) —» (Μ, bo,b\) is a path joining the two points, then there is an isomorphism !a/ = {o,i}.
§7. The Fundamental Theorem: Global Version 123 οσ:πι(Μ,6ι)-»πι(Μ,6ο) induced by the map on loops sending λ ι—► σ*λ*σ_1. Let σ: (7,0) —» (G, e) be the development of a g-valued 1-form a; on Μ along σ starting at e. Then we have the conjugation map ca^y.G —» G sending # ι—► σ(1)^σ(1)_1. Theorem 7.11. Lei Φ^·:πι(Μ,6j) —> G, j = 0,1, denote the monodromy representations with respect to two base points. Let σ: (7,0,1) —» (Μ, fro? &i) 6e α ραί/ι joining &o to b\. T/ien £fte following diagram commutes. ^(M,b0) —->G Φο λγ,(Μ,*,) —^G Proof. Let λ: (7,97) —» (Μ, bi) be a loop on Μ representing the element [λ] e 7Ti(M,fei) and let σ:(7,0) -» (G,e) and λ: (7,0) -» (G, e) be the developments of σ and λ, respectively, starting at e. σ(1)λ(1) = σ(1)Φ1([λ]) The diagram shows that the development of σ * λ * σ~ι is σ * (σ(1)λ) * (σ(1)Φ1([λ])(σΓί)), which ends at а(1)Фх([А])(^)(1) = ^^([ADSil)-1 = ^^([λ]). The equation (σ_1)(1) = σ(1)_1 used here comes from Exercise 7.5(a). ■
124 3. The Fundamental Theorem of Calculus Proposition 7.12. Let Μ be a smooth connected manifold, letb G M. and let /: (M, b) —» (G, g) be a smooth map. Set ω = /*α;^. ТЛега for any curve σ: (7,0,1) —» (Μ, Ь,х), ί/ie development, starting at g, ends at f(x). Proof. Since (/σ)*ωο = cr*f*ujG — σ*ω, the development of ω along σ: (7,0,1) -» (Μ, b, χ) is just /σ, and /σ(1) = f(x). Ш Corollary 7.13. The monodromy representation of a Darboux derivative is trivial. Proof. Write ω = /*cjg, where /: (M,b) —» (G,g). Replacing / by Lg-\f if necessary (which doesn't alter the Darboux derivative), we may assume that /: (M, b) -> (G, e). For any loop λ: (7,0,1) —► (Μ, 6,6), the proposition says that the development, starting at e, of u; along λ ends at /λ(1) = /(b) = e. Thus Φω([λ]) = β. ■ The Fundamental Theorem Now we can assemble our results into the following statement, the fundamental theorem of nonabelian calculus. Theorem 7.14 (Fundamental theorem of calculus). Let G be a Lie group with Lie algebra g. Let Μ be a smooth connected manifold and let ω be a g-valued 1-form on M. Then {ω is the Darboux derivative} J \.L J7 2L ? J ·> I г Д/Г ^ } ^ \ (n) the monodromy representation > of some map Μ -> G / | ^ ^^ ft) ^ ^ ^ ^.^ J Moreover, if these conditions are satisfied, the integral of ω is unique up to left translation by a constant element of G. Proof. The last statement is just the uniqueness of the primitive, Theorem 5.2. For the implication =>, (i) is shown in the discussion following Definition 5.1 and (ii) is Corollary 7.13. Let us verify the implication <=. Define /: (M,b) —» (G, e) by declaring f(x) to be the endpoint of the development of ω along any path from b to x. This definition makes / well defined under homotopies of the path by Theorem 7.7 and under general changes in the path by the very definition of monodromy, Definition 7.9. We need to check that ω = Uf. Note that the value of f(x) may be obtained in two steps as follows. If xo G Μ, we may choose a path from b to xq and then another from χ ο to χ. Then if the development of the first path starts at e and ends at go, while the development of the second starts at e and ends at g, then the development of the composite path will start at e and end at g$g.
§7. The Fundamental Theorem: Global Version 125 Now the local existence theorem guarantees that, for each point χ ο G Μ, there is a connected open set U about xo and also a smooth map fu-U -+ G satisfying f^uc — ω. After a left translation of fu by some element of G, we may assume that fu(xo) — f(xo)· Now the development of ω, starting at f(xo), along any curve σ: (7,0,1) —► (U,xo,x) ends at fu(x)· By the remarks in the last paragraph, it also ends at f(x), so fu(x) — f(x) for all x eU. Thus, /*cjg = fu^G = ω on U. Ш Some Remarks About the Fundamental Theorem The fundamental theorem certainly generalizes the usual fundamental theorem of elementary calculus concerning maps /: (a, b) —► R. In that case, conditions (i) and (ii) are automatic, so the theorem reads that every smooth R-valued 1-form on [a, b] is the derivative of some smooth map / that is unique up to an additive constant. It also generalizes the theorem that says that a vector field v: U —► Rn on a simply connected, open set U of Rn that satisfies the equation dvj dvj = dxj dxi for all г < i,j < n, has a potential function. To see this, write ν — (v\,... ,vn), and set ω — Y^Vidxi. Then ω is an R-valued 1-form on [/, and we have l<i,j<n г l<i<j<n Ч г 3/ Moreover, since R is abelian, [ω, ω] = 0 and hence the structural equation holds. Thus, by the fundamental theorem, a function V: U —► R exists such that V*{dx) = ω. Now (V*(dx))ei = dx(V*{ei)) = dx(dV/dxi) = dV/dxi. Thus ω = У*б?х = Y^{dV/dxi)dxi, that is, ^ = dV/dxi, for all 1 < г < η. It is worthwhile to see (at least once) the coordinate expression of the structural equation for a nonabelian group. We shall do this for the group G/n(R). Of course, the same expression then works for any subgroup. Let us assume that ω is an Mn(R)-valued 1-form on some open set U of Rm. Thus, for ν G T(U), we have ω(υ) = (Aij(v)), where the entries Aij are R-valued 1-forms on U. We use the basis др = д/дхр, 1 < ρ < га, and set Ap = (Aij(др)) so that Ap: U -+ Mn(R). Then dw(dp, dq) = dpu(dq) - dqu(dp) - ш[др, dq] = dpu(dq) - dqu(dp) = UpAq — UqAp and
126 3. The Fundamental Theorem of Calculus 1 2 Thus, the structural equation becomes -[u,u](dp,dq) = [ш(др),и(дч)] = [Ар,Ая\- ^-^ + [Ap,Aq} = 0 iorl<p,q<m. In terms of the entries Ацр (where Ap = ^]κ· ·<η Л^ре^·), the structural equation is ~Q^ ~Q^ Ь Z_^ {AikpAkjq - AikqAkjp) = 0 l<fc<n for 1 < p, q < m and 1 < г, j < n. The fundamental theorem tells us that whenever we see equations like these we may interpret the Aps or AijPs as the coordinate expression of an Mn(R)-valued 1-form ω that on U is locally (and globally if U is simply connected) the Darboux derivative of a Gln(H)-valued function В on U. That is, B~ldB = ω. Moreover, if U is connected, then such a B is unique up to left translation. This is an existence and uniqueness theorem for solutions of a certain partial differential equation. In coordinates it says that for U connected and simply connected, and provided the structural equations hold, the system dB -— = BAV for 1 < ρ < m oxp or, the same as above but in more detail, the system дВц У^ BikAkjP for 1 < ρ < m and 1 <i,j <n dxP i<k<n has a solution,13 which is unique up to left multiplication of the matrix В by a constant matrix. The Lie Group-Lie Algebra Correspondence We end this section with some applications of the fundamental theorem of calculus to the Lie group-Lie algebra correspondence outlined at the end of §2. Proposition 7.15. Let G and Η be Lie groups with Lie algebras g and I). // G is connected and simply connected, then every homomorphism of Lie algebras φ-Q -^ Ь is induced by a unique homomorphism Ф: G —> H. The solution is a "nonabelian potential function."
§7. The Fundamental Theorem: Global Version 127 Proof. The composition ω = φωο is an \)-valued, left-invariant 1-form on G that satisfies the structural equation (since άω + \{ω,ω] — άφωο + \{φωο,φωο] — φάωο + ^[<^g^g] — 0). Thus there is a unique smooth map Φ: (G, e) —► (H,e) satisfying Φ*(ω#) = ω. We claim that Φ is a homomorphism. To see this, consider /: (G, e) —► (H,e) defined by f(g) = Ф(#о)_1Ф(#о#) for some fixed g0. Now / = L<p(go)-i ο Φ ο Lgo, and so f*uH = 1;Ф\о).1Уй = Ь*доФ*шн = Vgu = ω. Since / and Φ agree at the identity and have the same Darboux derivative, it follows that / = Φ and hence Ф{д) = Ф(до)~1Ф(дод), that is, Ф(#о)Ф(#) = Ф(#о#) for all g and for all g0. Ш In particular, connected and simply connected Lie groups with isomorphic Lie algebras are themselves isomorphic. More precisely, we have the following result. Corollary 7.16. Let G and Η be connected and simply connected Lie groups with Lie algebra, g and \) and let φ.%-+\) be an isomorphism of Lie algebras. Then φ is induced by a unique isomorphism of groups Φ: G —> H. Proof. The existence and uniqueness of a homomorphism Φ: G —► Η is guaranteed by Proposition 7.15. Since φ is an isomorphism, it follows that the kernel of Φ is a discrete central subgroup and so Φ is a covering map. Since Η is simply connected, the covering must be trivial and hence Φ is an isomorphism. ■ We may also apply these ideas to obtain the correspondence between the representations of a Lie group and the representations of its Lie algebra. Definition 7.17. A representation of a Lie algebra g on a finite vector space V is a Lie algebra homomorphism p: g —► Ql(V). * Corollary 7.18. Let G be a connected and simply connected Lie group with Lie algebra g. Let V be a finite-dimensional vector space and let p: g —> gl(V) be a representation. Then there is a unique representation R:G —> Gl{V) satisfying R*e = p. Proof. Apply Corollary 7.16. ■ Finally, we want to show that every (real) Lie algebra is the Lie algebra of a Lie group. For this we use (the real version of) Ado's theorem, which we state without proof. Theorem 7.19 (Ado's theorem). Every real Lie algebra g is isomorphic to a subalgebra of the Lie algebra gl(V) for some finite-dimensional real, vector space V.
128 3. The Fundamental Theorem of Calculus Proof. Cf. [M. Postnikov, 1986] or [W. Fulton and J. Harris, 1991]. ■ Theorem 7.20. Let g be an arbitrary, finite-dimensional, real Lie algebra. Then there is, up to isomorphism, a unique connected and simply connected Lie group G whose Lie algebra is isomorphic to g. Proof. By Theorem 7.19, we may regard the Lie algebra g as a subalgebra g С gl(V), where gl(V) is the Lie algebra of the Lie group Gl(V). By Exercise 6.3, there is a unique connected Lie group Go С Gl(V) with Lie algebra g. Let G be the universal cover of Go- Then by Corollary 8.II,14 G is also a Lie group with Lie algebra g and it is connected and simply connected. By Corollary 7.16, if G' is another connected and simply connected Lie group with Lie algebra g, then G and G' are isomorphic. ■ §8. Monodromy and Completeness In this, the final section of this chapter, we study the monodromy group from the viewpoint of covering spaces. In certain cases we relate the completeness of a Lie algebra-valued 1-form to the discreteness of the period ёгоирГ = Фы(тг1(М,Ь))сС. Monodromy from the Viewpoint of Covering Transformations Let π: Μ —> Μ be the universal cover of M, and let b G Μ be a choice of base point lying over b. Because π is a local diffeomorphism, the form π* (ω) satisfies the structural equation when ω does. In this case, since Μ is simply connected, its monodromy is trivial. Thus, by the fundamental theorem, there is a unique primitive /: (M,b) —► (G,e) for π*(ω). Proposition 8.1. There is a unique homomorphism Φ: Qaf(M/M) —» G satisfying fT = Φ(Τ)/ for every Τ G Qat(M/M)}b Proof. Every covering transformation Τ G QaC(M/M) preserves the form π*(ω) (since Τ*(π*(ω)) = (πΤ)*ω = π*(ω)), and so we have (fT)*uG = T*(f*(uG)) = Т*7г» = π» = f*uG. Thus, by the fundamental theorem, fT and / differ by left multiplication by a constant element Ф(Т) G G, that is, fT — Ф(Т)/. In particular, 1 We do not appeal to the statement of Theorem 7.20 in any other proof of this chapter. 15By QaC{M/M) = {T e Diff(M) | ποΤ = π}, we denote the group of covering transformations.
§8. Monodromy and Completeness 129 fT(b) = Ф(Т)/(Ь) = Ф(Т). To see that Φ is a group homomorphism, note that Φ(5Γ) = f(ST(l·)) = Ф(5)/(Г(Ь)) = Ф(5)Ф(Г)/(Ь) = Ф(5)Ф(Т). ■ Now we show that the isomorphism σ: QaC{M/M) —► πι(Μ, b) sending Τ ι—> [π(λ)] (where λ is any path on Μ joining b to T(b)) relates the two versions of monodromy. We have the following result. Proposition 8.2. The following diagram commutes. QaC(M, b) •|\ 1 φ ^ щ(М,Ь) ► G Proof. The equation fT(b) = Φ(Τ) identifies Φ(Τ) as the endpoint in G of the development of π*ω along a path in Μ joining b to T(b). Clearly, development of ω along the image loop in Μ starting at b yields the same endpoint (in G). This fact is the commutativity of the diagram. ■ Completeness Next we introduce the important notion of a complete 1-form. Definition 8.3. Let a; be a F-valued 1-form on the manifold M. A vector field X on Μ is said to be ω-constant Ίϊω(Χ) is constant on M. The form ω is said to be complete if every cj-constant vector field X on Μ is complete. Example 8.4. Every F-valued 1-form on a compact manifold Μ is complete. This is a consequence of Proposition 2.1.6, which says that every vector field on a compact manifold is complete. ♦ Example 8.5. Let a; be a complete V-valued 1-form on the manifold M. Let TV be a closed submanifold of M. Then ω \ Ν is complete. Note that this example may be of small content since there may not be any cj-constant vector fields on TV. ♦ Example 8.6. The Maurer-Cartan form on a Lie group G is complete. This follows since the ω constant vector fields are just the left-invariant fields, which we have shown to be complete (cf. Corollary 2.12). ♦
130 3. The Fundamental Theorem of Calculus Characterization of Lie Groups The final result, (Theorem 8.7), shows that in a certain sense the Maurer- Cartan form determines the group up to some covering. (It is this description of a Lie group that will generalize to the notion of a Cartan connection on a principal bundle. However, in this we are getting somewhat ahead of ourselves.) Theorem 8.7. Let Μ be a connected smooth manifold and let g be a Lie algebra. Let ω be a g-valued 1-form on Μ that satisfies the conditions (i) άω + ±[ω,ω] = 0, (η) ω: Τ (Μ) —» g is an isomorphism on each fiber, (iii) ω is complete. Then (a) The universal cover of M, π-.G —> M, has, for an arbitrary choice e G G, the structure of a Lie group with identity element e and Lie algebra g whose Maurer-Cartan form is π*ω. (b) The period group Γ = Φω(πι(Μ, b) С G acts by left multiplication on G as the group of covering transformations for the cover π: G —> Μ. Proof. The case πχ(Μ) — 1. Here we have G = M. This case involves five steps. The first shows how to construct a map /: Μ —» Gl(g) that will eventually be given by f(g) = Ad(g). The second step constructs, for an arbitrary choice of e G G — Μ, a "multiplication" map μ: Μ χ Μ —► Μ satisfying the conditions /i(e,e) = e and μ*ω = (тг^/)-1?Γία; -\-^2ω· The third step shows that / is a homomorphism with respect to the multiplication μ. The fourth shows that μ is associative. The fifth step constructs the inversion map l: Μ —» Μ, and shows that ι is inversion with with respect to μ. Step 1. Corollary 5.3 shows that /: Μ —> AutLie(s) should have Darboux derivative 77/ = ad(cj) (G glue(&)) and satisfy /(e) = e. Thus, the form ω will determine /, provided η — ad(cj) is integrable. But dr7+ ybliV] = ad(du;) + -[ad(u;),ad(u;)] = ad ( du + -[ω,ω] J = 0, and since Μ is simply connected, it follows from the fundamental theorem, Theorem 7.14, that η is the Darboux derivative of a unique map /: Μ —> Gl(g) satisfying /(e) = e. In particular, df = /oad(o;) (where the product on the right is composition of elements of Ql(g)).
§8. Monodromy and Completeness 131 Step 2. We construct the multiplication map. Proposition 4.10 shows that μ: Μ χ Μ —> Μ should have Darboux derivative at(x,i/)EMxM given by η = πί(Λ%_1)ω) + π*2ω = it\{f{y)-lbj) + -π\ω = {ττ*2Γι)ττ\ω + τ*2ω, where ж^.М χ Μ —* Μ denotes projection on the ith factor. Thus, the form ω will determine μ, provided η is integrable. Now άη + \[η,η] = ^(Г^ЖИ) + тг2*(Г ^(eb)) + тг2*(<М + ^[τ2(Γ1)(πί(ω)),π2*(/-1)(πί(ω))] + [^(/_1)(πί(α;)),π5α;] + |[π*ω,π>] = π2*Κ/-1))«(α;)) + π2*(/-1)«(ίίω + ^μ,ω])) + тг^еЬ + jjkw]) + Η(/_1)Κ(ω)),π2*ω]. Since d(/_1) = -adw о /-1 and16 π^-αάωΓ^πΚω)) = -7г2*^)тг2*(Г ^«(ω)) = -[7Γ2*ω,π2*(/-1)(7Γί(ω))] = -[т5(Г1ЖН),7г2*и;], it follows that c/77 -f |[т7,77] = 0. Since Μ is simply connected, the fundamental theorem shows that 77 is the Darboux derivative of a unique map μ: Μ χ Μ -> Μ satisfying /i(e, e) = е. 5iep 3. We show that /: Μ —» Μ is a homomorphism with respect to μ in the sense that /(μ(χ,ν)) = f(x)f(y) for all x,y e M. We shall verify the commutativity of the following diagram. MxM /x/ Gl(g)xGl(g) -» Μ Mc/ / * G/Ce> It is clear that this diagram commutes at (e,e). Since Μ is connected, by Proposition 1.4.19 it suffices to show that the Maurer-Cartan form wqi on Note that by Exercise 1.5.20, the bracket is symmetric on 1-forms.
132 3. The Fundamental Theorem of Calculus G/(g), with values in qI(q), pulls back to the same form on Μ χ Μ along the two possible routes. That is, we must show that (/ ο μ)*ω0ι = (μοι ο (/ χ f))*uGi- Calculating the left-hand side at (x,y) G Μ χ Μ, we have (ί°μ)*ω& = μ*ί*ωοι = μ*(&ά(ω)) (where ad(cj)?; = [ω,υ]) = &ά(μ*(ω)) = ad(^*(/-V^) + 7r2*u;) = ad((/-1(j/)«w)+^u;) = Ad(/_1(2/))«ad ω) + ^ad ω (since f~l(y) G AutLie(g)), while the right-hand side is (μαι ° (/ x f))*UGi = (/ x /)>^ ^g/ = (/ x /H^Ad"1)^*^) + tt*u;g/} = Ы о (/ х /)Τ(Αά-ι)(πΙωοι) + (тг2 ο (/ χ /))*uG/ = (/ ο ^)*(Ad-x)((/ ο πχγωοι) + (/ ο tt2)*u;g/ = (Adi/^)-1)^/*^) + τγ*/*^ζ = (Ad(/(j/)_1)«ad ω) + ^ad ω. This verifies that / is a homomorphism with respect to μ. Step 4- Now we show that μ is associative. We shall verify the commu- tativity of the following diagram. uxid Mx Mx Μ >MxM id χμ > MxM- -> Μ It is clear that this diagram commutes at (e,e,e). Since Μ is connected, it suffices to show that the form ω on Μ pulls back to the same form on Μ χ Μ χ Μ along the two possible routes. That is, we must show that (μο(μχίά))*ω = (μο(ίάχμ))*ω. Calculating the left-hand side at (x,y,z) G Μ χ Μ χ Μ, and using pi'. Μ χ Μ χ ΑΙ —> Μ, Pij- Μ χ Μ χ ΑΙ —>· Μ χ ΑΙ to denote projection to the ith or zjth factors, we have (μ ο (μ χ id))*£j = (μ x id)*/i*u; = (μ x Ίά)*(/(ζ)-ιπΙω+πΖω) = f(z)-l(m ο {μ χ id))*£j + (π2 ο (μ χ id))*u;
§8. Monodromy and Completeness 133 = /(2)_1(μ° Ρΐ2)*ω + ρ%ω = /(«)_1Ρί2(/(ί/)_1τίω + π» + ρ%ω = f(z)-lf(y)-l(*i ο Ρΐ2)*ω + }{г)-\ж2 о ρ12)*ω + ρ*3ω = }{ζ)-ιί{υ)-ιρ\ω + ί{ζ)-ιρ*2ω + ρ*3ω. On the other hand, (μ ο (id χ μ))*ω = (id χ μ)*μ*ω = (id χ μ)*(/(νζ)-ιπΙω + π$ω) = f(yz)~l(ni ο (id χ μ))*ω + (π2 ο (id χ μ))*ω = f(yz)~lplu + (μ ο ρ23)*ω = f(yz)~lp*u + p22V*u = f{yz)~lp*u + Р2з(/(^)~17ГГ^ + *2fc>) = f(yz)~lp*U + /(^)~1(7Г1 ° P23)*^ + (ТГ2 О p23)*^ = f(yz)~lp*u + f{z)~lp2U + Рз^ = f(z)-lf(vTlPlu + /CO"1/*" + р5"· This verifies the associativity. 5iep 5. Finally, we construct the inversion map. Proposition 4.10 shows that i\M -* Μ must have Darboux derivative given by 77 = —Ad(x)o; = — /ω. Thus the form ω will determine 6, provided we can show that —fu is integrable. But since άη + 2fo> r/] = -d/a; - / da; + -/[ω,ω] = -f a,d(u)u - f 1--[ω,ω]1 +-f[u,c = -/[ω,ω] + /[ω,ω} = 0 and Μ is simply connected, it follows from the fundamental theorem that 77 is the Darboux derivative of a unique map l:M —> Μ satisfying u(e) = e. However, we must still verify that ι is the inverse map, that it satisfies μ(ί(χ),χ) = e. For this we shall verify the commutativity of the following diagram. MxM iXld>MxM 4 ϊμ Μ >e eM
134 3. The Fundamental Theorem of Calculus It is clear that this diagram commutes at e. Since Μ is connected, for the general commutativity it suffices to show that the form ω on Μ pulls back to the same form on Μ along the two possible routes. Since it obviously pulls back to 0 via the lower route, we must show the same for the upper route, namely, that (μ(ι χ id)A)*u; = 0. Calculating at χ G Μ, we have μ ο (l χ id) ο Δ)*ω = Δ*(^ χ id)*/i*u; = Α* (ι χ id)*(f(x)-lnlu> + π» = A*(f(x)-l(t χ id)*7r*u; + (tx id)*^a;) = A*(f(x)-1{n1 o(lx id))*cj + (π2 ο (l χ id))*cj) = A*(f(x)~l(L ο 7Γι)*ω + π^α;) = Δ*(/(χ)_1π^*α; + π» = Δ*(/(χ)"1πί(-/(χ)ϋ;)+π5ϋ;) = Δ*(-7Για; +π^ω) = -(πιΔ)*α; + (π2Δ)*α; = -α; + α; = 0. Thus, ι is an inverse for μ. The case πχ(Μ) φ 1. Here the form a; on Μ pulls up to the form π*α; on the universal cover G of M. Clearly, π*α; satisfies the same three conditions on G that ω satisfies on M. Thus case 1 applies to the pair (G,cj), and we see that for any fixed choice of e G G there is a Lie group structure on G for which e is the identity and 7T*u; is the Maurer-Cartan form. By Proposition 8.1, there is a unique injective homomorphism Φ: QaC{G/M) —► G satisfying Ф(Т)д = T(g) for all Τ G QaC(G/M). Thus, the group of covering transformations is identified with the period group, a discrete subgroup Г С G. ■ Remark 8.8. Dropping condition (iii) in Theorem 8.7 still leaves us with the manifold M, which is locally a Lie group. Remark 8.9. If we drop condition (i), we are left with what may be regarded as a "deformation" of a Lie group. For example, we might consider altering the form ω, which does satisfy (i), (ii), and (iii), by the addition of a small form η such that ut = ω -\- ίη still satisfies (ii) for t G [0,1] so that ω χ is a deformation of ω. The new form may also satisfy (iii), but it will not, in general, continue to satisfy (i), and so we will not even get a Lie group locally. Remark 8.10. Except in special cases, we cannot weaken condition (ii) to the following (ii)7 ω:Τχ(Μ) —► g is an injection for all χ
§8. Monodromy and Completeness 135 and still expect to obtain a discrete monodromy group. For example, we could take the case where G = Sl3(C) (so that g = M3(C)0, the 3x3 matrices of trace 0), Μ = S1, and ω is the g-valued 1-form on 51 which is the composite φωΜ, where им is the Maurer-Cartan form on S1, and φ: R —► g is the Lie algebra homomorphism given by χ ι-* diag(zx, л/2гх, —(1 Η- л/2)гх). Of course, locally, φ is the derivative of the locally defined map S1 -> T2 С 5/3(C) sending eix .-> diag(eix, e^ix, е-(1+^)^). The form ω satisfies the structural equation automatically, since the dimension of Μ is 1, and since Μ is compact ω is complete. Moreover, condition (ii)' is obviously satisfied. What is the monodromy group Γ of ω? Clearly, it is the subgroup generated by the element diag(l, e2v/2i7r, e~2^iir) G T2, and this subgroup is not a discrete subgroup of G. Corollary 8.11. Let G be a connected Lie group and let π:0 —> G denote the universal covering space. For any choice of ё G n~l(e), there is a unique Lie group structure on G such that ё is the identity and π is a homomorphism of Lie groups. Moreover, the kernel of π is a discrete central subgroup ofG. Proof. First we note that if such a Lie group structure exists on G, then (by Proposition 2.9) the Maurer-Cartan form on G is π*ωο· Thus, we are led to define a g-valued 1-form on G by ω — 7t*o;g, and this form satisfies the structural equation since άω + -[ω, ω] =άπ*ωα + -[тт*и;с,7г*и;с] = π*άωα + ^[ω0,ω0] = π* (a\jG + -[uG,uG] = 0. Since π is a local diffeomorphism, not only is ω: T(G) —» g an isomorphism on each fiber but also ω is complete, since integral curves for uG on G will lift to integral curves for ω on G. By Theorem 8.7, given ё G 7r_1(e), G has a unique Lie group structure with ё as the identity and for which ω is the Maurer-Cartan form. By Theorem 5.2, π is the unique map G, ё —> G,e pulling back ω from uG, and then by Proposition 7.15, π must be a homomorphism. Finally, since ker π С G is a discrete subgroup of a connected group, the function fz: (G, e) —» (ker π, e), where ζ G ker π, must be constant. Thus, fz(g) — fz(e) — z-> and hence ker π is central in G. Ш
136 3. The Fundamental Theorem of Calculus The Classical Period Mapping Remark 8.12. This final remark constitutes a brief discussion of the classical period mapping and is addressed to those who know something about complex analysis. Let Μ — Mg be a Riemann surface of genus g. Then Hl(Mg, C) = Ί?9 is spanned by g closed (i.e., dw = 0) holomorphic 1-forms cji, ... ,ω9 and their conjugates. Let ω = (ω\,... ,ω9). This is a closed 1- form on Mg with values in C9 = R2fif which is the (abelian) Lie algebra of C9 (so that [ω, ω] = 0). Thus, ω satisfies the structural equation. Moreover, it is complete since Μ is compact. It is known from complex analysis that ω:Τρ(Μ) —» С9 is complex linear and injective for every ρ G Μ and that the period group is discrete. However, the period group must contain a basis for C9, for if it were merely to span a proper (complex) subspace V, then the composite period mapping projection Μ PP *>C8/L—· > C8/ V would give a nonconstant holomorphic function on Μ, which, by the maximum modulus principle, is impossible for Μ compact. Thus, L is a lattice in C9, and we get the map Mg —► C9/L, which is classically called the period mapping. The receiving group C9/L is called the Jacobian variety.
4 Shapes Fantastic: Klein Geometries Let us now do away with the concrete conception of space ... and regard it only as a manifoldness of η dimensions .... By analogy with the transformations of space we speak of transformations of the manifoldness; they also form groups. But there is no longer, as there is in space, one group distinguished above the rest by its signification; each group is of equal importance with every other. The following comprehensive problem then arises as a generalization of geometry: "Given a manifoldness and a group of transformations of the same; to investigate the configurations belonging to the manifoldness with regard to such properties as are not altered by the transformation of the group." -Felix Klein, 1872 During the 19th century, non-Euclidean geometry made its appearance with the independent discovery by Gauss, Bolyai, and Lobachevski of hyperbolic geometry in the plane. This geometry is very close to Euclidean geometry, satisfying as it does all the axioms except for the existence of a unique line through a point parallel to a given line. Elliptic geometry, the geometry of antipodal pairs of points on the sphere, appeared and again was non-Euclidean in this strong sense. In fact, another geometry, spherical geometry, had been studied by navigators for centuries, without being considered as "non-Euclidean" since it has models in Euclidean 3- space and perhaps also because, considered in its two-dimensional aspect, it more violently violates the axioms of Euclidean geometry. In the course
138 4. Shapes Fantastic: Klein Geometries of a few decades, the number of these new geometries proliferated to include affine geometry, projective geometry, Mobius geometry, Lie sphere geometry, Laguerre geometry and so on. Each geometry has its own set of theorems, that is, its own theory. In addition to this there are relations among the geometries. For example a theorem that seems to belong to Euclidean geometry might still be true in projective geometry. Moreover, a purely projective proof, which could employ the larger group of projective symmetries, might well be simpler.1 It was Felix Klein's idea to bring order to these new geometries by means of the unifying notion of the principal group2 of a geometry. He noticed that each geometry, Μ say, is a connected manifold and has a Lie group G of "motions" acting transitively3 on it and, moreover, all the properties of figures studied in the geometry remain invariant under these motions. If, in addition, the action is effective,4 we may speak of an effective Klein geometry. It is this group G of motions that Klein called the principal group of the geometry, or more briefly, the group of the geometry. In the case of Euclidean geometry, the properties studied are angle and length, and the group is the group of rigid motions; for projective geometry, the properties are concurrence of lines and collinearity of points and the group is the group of projective transformations, and so forth. Thus, Klein extended the notion of geometry by defining a geometry to consist of a Lie group G, a smooth manifold Μ, and a smooth, transitive, and effective action of G on M.5 The study of such a geometry is the study of those properties of figures in Μ that remain invariant under the action of G. 1 Associated to each of these geometries is not only a simple theory (a goniom- etry) in which one studies figures like triangles but also a differential geometry in which one studies the "figures" consisting of smooth curves and surfaces in the geometry. A great deal of effort was spent in studying these generalizations of differential geometry in the last decades of the 19th century and in the first four decades of this century. 2Klein called this group the haugtgruppe of the geometry. In his translation of Klein's Erlangen program [F. Klein, 1872], Haskell renders this as principal group. This has the two-fold advantage that the modern notion of a principal bundle generalizes this group and that the competing English term fundamental group, which many authors have used, is easily confused with Poincare's completely accepted use of the latter term for πι(Μ). 3Let G χ Μ —» Μ sending (ρ, χ) —> gx be a group action. Recall that the action is said to be transitive if, for any pair of points x,y G M, there is an element g G G such that gx = y. 4An action is said to be effective if gx = χ for all χ G Μ implies that g is the identity element of G. 5The formal definition of a Klein geometry is given in §3. We shall not insist on an effective action.
§1. Examples of Planar Klein Geometries 139 Klein's generalization of geometry allows us to shift the emphasis from the space Μ to the group G in the following manner. Let us fix a base point x G M. Then there is a map π: G —» Μ sending g to gx. By the transitivity of the action, this map is onto; but it is not one to one. The inverse image π~λ(x) = {x £ G \ gx = x} = Hx is clearly a closed subgroup of G. It is called the stabilizer of χ. Moreover, if у G M, it is clear that тг~1(у) — {g G G | gx = y} — 9οΗχ, where go is an arbitrary element of тг~1(у). It follows that π induces a bijection G/Hx —» M. The condition that the action be effective translates into saying that the subgroup TV = {g G G \ gp = ρ for all ρ G M} is trivial. Exercise 0.1. If we do not assume that the action is effective, show that this subgroup N is the largest subgroup of Hx that is normal in G. □ Thus, to summarize, we may say that instead of describing a geometry with base point as a pair (M,x) together with its principal group G, we could equally well describe it as the pair (G, ϋΓ), where Η — Hx is a closed subgroup of G. Of course, all this depends on the choice of base point x. This base point may be regarded as "a constant of integration," which disappears in the infinitesimal description of Klein geometries. Exercise 0.2. Show that different choices of χ lead to conjugate subgroups tf. □ It is clear that in these Klein geometries it is impossible to distinguish a point from any other by properties of the geometry since, by definition, the transitive action of G preserves these properties. Thus, the Klein geometries are fully homogeneous. We might say that they are "flat" or that they have no "lumps" in them. Although we lend the name "Klein geometry" to the geometries studied in this chapter—and the justice of this should be clear from the quotation at the beginning of the chapter—Klein himself was interested in only the few cases that correspond most closely to our experience: projective geometry and its sons and daughters, the parabolic, Euclidean, and elliptic geometries. Thus, he would have felt our present point of view to be too general. It was that remarkable pioneer Wilhelm Killing whose passion it was to try to understand these geometries in general. Our aim in this chapter is not to study the theories of individual Klein geometries, but rather to try to understand their nature in general.
140 4. Shapes Fantastic: Klein Geometries §1. Examples of Planar Klein Geometries Let us see how Klein's ideas, described in this chapter's Introduction, work in some cases of more or less familiar two-dimensional geometries.6 Each of these planar geometries is merely a single example drawn from an infinite series of higher-dimensional analogs. "School Geometry" (the Euclidean plane). In this case Μ — R2. The Lie group of symmetries or congruences or rigid motions7 is G = £7uc2(R) = |^ 0 (cos* -άηθ\ Λ/ι v ' l sin θ cos θ J \v2 which acts on Μ by the formula г»/л\ ) -x = R(0)x + v, where χ — [ l V Щв) J K ' \X2 An easy calculation shows that the stabilizer of the origin of R2 is the subgroup of rotations Η = 502(R) = {(J д^рен}. Then we have the basic identification Euc2(R)/S02(R>) ~ R2· In fact, Euc2(R) и 50г(К) х R2 (the semidirect product with respect to the standard representation of 50г(К) on R2). Next follow the two non-Euclidean8 geometries. Hyperbolic Plane. In this case Μ = {ζ — χ + iy G С | у > 0}. The group is the group of Mobius transformations G — ^^(R) = {A E M2(R) | det A = 1}, which acts on Μ by the formula a b \ az + b ζ — с d J cz + d' It is easy to check that this is indeed an action on Μ and also that the stabilizer of the point г G Μ is the circle group 6 A complete list of all 23 families of two-dimensional Klein geometries (13 of which are singleton families) may be found in [B. Komrakov, A. Churyumov, and B. Doubrov, 1993]. 7Of course, in picking this group we have rather arbitrarily chosen to ignore the orientation-reversing congruences. In fact, there are actually two distinct interpretations of "school geometry." 8That is, non-Euclidean in the strong sense. See the introduction to this chapter.
§1. Examples of Planar Klein Geometries 141 S<32(R) = {A e M2(R) | det A = 1 and AAl = /}. Elliptic Plane. This is the real projective space P2(R), which may be regarded as the set of unoriented one-dimensional subspaces of three-dimensional Euclidean space R3, with coordinates xo1x\1X2- (The coordinate charts of P2(R) "are" the 2-planes in R3 that avoid the origin, cf. Example 1.3.) The group is SOs(R), whose action on the lines in R3 is induced by the standard action of SOs(R) on the ambient space R3. The stabilizer of a line, say the xo-axis, is 02(R)? which is embedded in SOs(R) via det A~l 0 0 A Complementing these, we have a series of six other geometries. Geometry of Similarity. This geometry is a relative of Euclidean geometry that may also be regarded as a part of "school geometry." Here Μ = R2 and the group is the two-dimensional conformal group given by Q — Η χ R2 (semidirect product), *-{ A e Glt(R) r2 0 0 AAl = ( л 2 ) for some r G R ' An element A G Η has the form A — ( . a ) acting by matrix multiplication on the plane. r sin θ r cos θ Lorentz (or Minkovski) Plane. Here Μ — R2 and the group is the oriented two-dimensional Lorentz group given by G = Η χ R2 (semidirect product), where Я = 0^1(К) = {Ле5/2(К)|Л*ЕЛ = Е}, where Σ = ( J Д V An element A G Η has the form ^4=1 . , й ,fl (a boost) acting by matrix multiplication on the plane. Affine Plane. Here Μ — R2 and the group is the two-dimensional affine group given by G = Aff+(R) = |Q °) e GI+(R)\a e Gi2+(R), ν G R2J . The action on Μ is given by the formula ( \ · χ = Ax + v, where χ and ν are column vectors. Again, it is easily verified that the stabilizer of the origin is
142 4. Shapes Fantastic: Klein Geometries A G G/+(R) \ и GZ+(R) and G ~ Я χ R2 (semidirect product). Real Projective Plane. The space is the real projective space P2(R) just as in the elliptic case, but the group is now the full general linear group Gls(R). The group Gls(R) acts on R3 in the standard way and so it also acts on the "lines" in R3. The stabilizer of the xo-axis is #, where A gGZ2(R), veB2, AeR*| and TV = {XI G Gls(R) | λ G R*} is the largest normal subgroup of G in #. Mobius Plane. Next we consider the Mobius plane. Here the space Μ is the standard 2-sphere, 52 = {x G R3 | x\ + x\ + x\ — 1}. The principal group is the oriented Lorentz group L3ji(R) = {A e SU(R) \ ΑιΣΑ = Σ}, where / 0 0 0 -1\ 0 10 0 1 Ιο οι or \-l 0 0 0/ This time, however, it is more difficult to describe the action. The Lorentz group G = Z/35i(R) acts on R4 so that the level sets of the quadratic form —2x0^3 -f x\ H- x\ are preserved. In particular, the light cone L, given by —2x0^3 -f x\ H- x\ = 0, is preserved. Now the light cone is homogeneous in the sense that if χ G L, then Xx G L for all λ G R. Thus, L is a union of lines through the origin, and the linear action of G on R4 restricts to give an action on this set of lines. Each such line meets the hyperplane ^о + ^з — λ/2 in exactly one point, so the set of lines in the light cone may be identified with the intersection Ln{xGR4|xo+^3 = \/2} = {x G R4 | x0 + хз = л/2, -2x0^3 + x\ + x\ = 0}. The change of variables xo = (1 +2/o)A/2,ffi = 2/1,^2 = 2/2, x3 = (l-!/o)/>/2, identifies the set of lines in the light cone with the 2-sphere {</GR3|</? + ^+^ = l}. H = {(1 a)^gi^ H = {(o l)^WR) This yields the action of G on S2.
§1. Examples of Planar Klein Geometries 143 Lie Sphere Plane (cf. [Т.Е. Cecil and S.-S. Chern, 1987] or [Т.Е. Cecil, 1980]). Prom one point of view, the space here is the quadric Q in P3(R), given by —XQ-\-xf-\-x% — x% = 0. Prom another point of view, it is the space Λ of lines on the quadric Q. Λ appears to be more important for Lie sphere geometry then Q (cf. the references cited above). However, dim Λ = 3, so this space is not a two-dimensional geometry, and we will restrict our remarks to Q, which does have dimension 2. Q is the projective completion9 of a hyperboloid of 1-sheet in R3 given by x\ -f x\ — x\ — 1 and so is diffeo- morphic to a torus. These facts can be pieced together from the following diagram. θ + π θ + π The finite points of Q have the interpretation as oriented 0-spheres on Sl as follows. For a finite point η G Q, we have the oriented orthogonal plane n1- meeting the circle on the light cone given by x3 = 1 in two (distinct) oriented points (one plus, one minus), that is, in an oriented zero spehre σ. As η tends to infinity, the two points of the corresponding sphere σ move together to yield an unoriented point sphere in the limit. the circle in the plane jc3= 1 _2_ 9Briefly, given any set of polynomials Pj(x) = 0, j = l,...,r, where χ = (χι,... ,xn) G Rn, we have, on the one hand, the locus X С Rn given by I = {xG Rn | Pj{x) = 0, j = 1,...,r}. On the other hand, we may homogenize the polynomials Pj to get polynomials Qj in the variables у = (xo, · · · ,a?n) by setting Qj(y) = Xq Qj(x/xo), where the powers dj are chosen just larger enough to make the Qs polynomials. Then the equations Q = 0 have a locus Υ С Рп, and Υ Π Rn = X. Then Υ is called the projective completion of X.
144 4. Shapes Fantastic: Klein Geometries (A point sphere can be perturbed in a pencil of spheres containing two copies of each unoriented sphere, one for each orientation; a point sphere may be thought of as the moment of change of orientation in a pencil of infinitesimal oriented spheres.) The group is G = 0^2(R) = {A e 5/4(R) Ι ΑιΣΑ = Σ}, where Σ = diag(-l,l, 1,-1). G acts transitively on the quadric, and we take Η to be the stabilizer of the line spanned by (1,1,0,0) corresponding to the equatorial 0-sphere S° consisting of the two points (±1,0,0) in the figure above. We mention one other plane geometry, which, although identical to one already considered, in its higher-dimensional incarnations becomes an independent geometry. Complex Projective Line. Here Μ — S2, the Riemann sphere (or the complex projective line), and the group G is the group of Mobius transformations of M, G = Sh(C), with H = <AeSl2(C) ::)}· Exercise 1.1. Show that S72 (C) is isomorphic to the Lorentz group and that the complex projective line is the same as the Mobius plane. □ §2. Principal Bundles: Characterization and Reduction The examples of the last section make it clear that the study of "non- Euclidean geometry" really is the study of the left coset spaces G/H, where Я is a closed subgroup of the Lie group G. There are, however, some technical questions that begin to arise. For example, it is not so obvious that a geometry in Klein's sense is always a smooth manifold. The purpose of this section is to deal with this difficulty and also to show that the map G —» G/H is a principal Η bundle.10 In fact, in anticipation of later need, we shall prove something still more general. We are going to characterize the principal Η bundles Ρ —» Μ in terms of properties of the action PxH —> P. At the end of the section we give a brief study of the notion of a reduction of a principal bundle. Definition 2.1. Let Ρ be a smooth manifold, Η a Lie group, and PxH —> Ρ a smooth action. 10Such bundles constitute the first and motivating examples of principal bundles; the Hopf bundle S1 —► S3 —► S2 studied in the first chapter is, of course, among these.
§2. Principal Bundles: Characterization and Reduction 145 (i) The action is called free if ph — ρ for some ρ G Ρ => h = e. (ii) The action is called proper if A and В compact => {ft G Я | Ah Π Β φ 0} is compact. ft Proposition 2.2. Lei G be a Lie group and Η С G a closed subgroup. Then the action G χ Η —> G given by multiplication is free and proper. Proof. Free action: If gh = g, then multiplying by g~l yields ft = e. Proper action: Let A and В be compact subsets of G, and let К = {ft G Я | Aft ПВ/Й}. We show that К is compact. Suppose that fti, ft2,... is an infinite sequence in K. Then there are infinite sequences αι,α2,.·· in Л and bi,&2,··· m В such that a^-ftj = bj for all j. Since A is compact, we may pass to a subsequence J С {1, 2,...} such that {aj}j^j converges to a. Passing to a further subsequence L С J, we may also assume that {bj}jeL converges to b. Thus {ftj = (aj)-1bj}jeL converges to ba~l and every sequence in К has a convergent subsequence. Thus К is compact. ■ Next we verify that right principal Η bundles have these properties. Proposition 2.3. Let ξ = (Ρ, π, Μ, Η) be a right principal Η bundle. Then the action Ρ χ Η —> Ρ is free and proper. Proof. The fact that the bundle is locally a product U χ Η, with the canonical right Η action, shows that the action is free. The argument in Proposition 2.2 shows that, for any sequence {ft^·} in К — {ft G Η \ Ah Π Я/0}, there are sequences {o>j}jeL in A and {bj}jeL in В converging to a and b, respectively, and such that ajhj = bj and for all j. Now n(a,j) = 7r(ajhj) = n(bj) for all j, so π (α) = π(6). Thus a = hb for some ft G H. We claim that \im{hj}jeL = h. To see this, note that for large j all of the terms in the sequences lie in a single coordinate chart. This allows us to assume we have the trivial bundle Ρ = Μ χ Η. Projecting on the Η factor by ρ leads to the equation p(aj)hj = p(bj) in the group H. Thus {hj = p(aj)~lp(bj)}jeL converges to p(a)~lp(b). Thus, every sequence in К has a convergent subsequence. ■ Now we complete this cycle of ideas by showing that these properties characterize principal bundles over a smooth manifold. We have the following result whose proof, being rather technical, is given in Appendix E. Theorem 2.4. Let Ρ be a smooth manifold, Η a Lie group, and μ: ΡχΗ —> Ρ a smooth, free, proper right action. Then (i) P/H with the quotient topology is a topological manifold (dim P/H = dim Ρ - dim Я),
146 4. Shapes Fantastic: Klein Geometries (ii) P/H has a unique smooth structure for which the canonical projection π: Ρ —» Ρ/Η is a submersion, (iii) ξ = (Ρ, π, Ρ/Η, Η) is a smooth principal right H bundle. Example 2.5. It is important that the action be proper. For example, in the case of the right action G χ Я —» G by a subgroup Я С G, if the subgroup is not closed, then the action is not proper and the quotient G/H need not be a manifold. The standard example of this is the case of the torus G = T2 = R2/Z2 and an irrational subgroup Я of G. To describe the meaning of this, we pass to the universal cover G\ = R2 with the corresponding lattice Ζ2 = {(α, b) Ε R2 | a, 6 G Z} as kernel of the projection map R2 -» R2/Z2. Let HY = {(i, д/2 i) | * G R}. In Gx the subgroup H\ is a straight line, so it is closed. However, since y/2 is irrational, its image Я in G is a dense subgroup; in particular, Я is not closed. To see how bad the quotient topology on G/H is, note that if /: G/H —» R is any continuous function, then its lift to G is also continuous. Since / is constant on a dense subset of G, it is constant everywhere. Thus there are no nonconstant continuous functions on G/H. In particular, it is not a manifold. ♦ Example 2.6. From the preceding example we might be led to wonder if the problem arises only from the fact that the torus T2 fails to be simply connected. Might it be true that for simply connected groups G, subalge- bras of g may always be realized as closed subgroups of G? The answer is again no in general, for many simply connected groups contain the torus as a subgroup; and so the phenomenon of the previous example persists. For example, Sl^(C) is simply connected, and yet it contains the torus T2 as the subgroup diag(e^, βίφ', е"^+(^), θ,φβΈΙ. ♦ Exercise 2.7. Show that if G is a Lie group and Я is a closed normal subgroup of G, then G/H has a unique Lie group structure such that the projection map π: G —» G/H is a homomorphism of Lie groups. □ Exercise 2.8. Let ρ: Χ χ Я -» X (x,h) ι—► χ h be a right action of a topological group Я on a Hausdorff space X. Show that ρ is proper as an action if and only if the map X xH ->X xX (x,h) ι—► (χ,χ h) is proper. □
§2. Principal Bundles: Characterization and Reduction 147 Exercise 2.9. Let p: X x Я -» X (x,h) ι—► χ h be a proper right action of a topological group Я on a Hausdorff space X. Show that XxHxH -> X xH (x,h,k) ι—► (χ fc,fc_1/i) is also a proper action. □ Exercise 2.10. Let W be a real n-dimensional vector space. Let Gvs(W) denote the set of all s-dimensional subspaces of W. Show that Grs(W) can be given a unique smooth structure such that, if V G Grs(W) and W = U Θ V, then the canonical map Hom(V, {/) —» Grs(VF) sending / ι-* graph(/) is a smooth map onto an open subset. [Hint: Fix a basis and an inner product in W. Show that the map On(R) —» Grs(VF) sending Л to the subspace of VF spanned by the first s columns of A induces a bijection of sets On(R)/(Oe(R) x On_s(R)) -> Grs(W).] □ Definition 2.11. The manifold Grs(W) described in Exercise 2.10 is called the Grassmannian of s-dimensional subspaces ofW. Φ Reduction of Principal Bundles An important procedure in working with principal bundles is the notion of a reduction of a principal bundle. Definition 2.12. Let Я be a Lie group and Hq С Я a subgroup. Let Я -> Ρ -> Μ be a principal Я bundle over M. An Я0 reduction of this bundle is a submanifold Pq C. Ρ such that Po —* Μ is an Hq bundle and the action of Щ on Po is the restriction of the action of Я on P. * The following lemma is a preparation for studying one way in which reductions arise. Lemma 2.13. Let μ-.GxM -» Af (g,x) ^ g x be a smooth left action of a Lie group G on a connected smooth manifold M. Then every orbit X С М of this action is a submanifold. Moreover, if the action is proper, then X is a proper submanifold. Proof. Fix an orbit X С Μ and choose x0 G X. Set Я = {g G G \ gxo = xo}. It will suffice to show that Я is a closed subgroup of G, that the induced map
148 4. Shapes Fantastic: Klein Geometries G/H -* Μ gH y-> g x0 is an injective immersion with image X, and that, if the original action is proper, then G/H —» Μ is a proper embedding. S'iep i. Я is a closed subgroup of G, and the induced map G/H -^ Μ gH >-+ g x0 is injective with image X. Since μ~ι(χ0) is closed and Я χ χ0 — μ~ι(χο) ΓΊ (G χ Χο), it follows that Η χ x0 is closed in G χ Μ and hence Я is closed in G. The fact that the map is injective with image X is clear. Step 2. G/H-^ Μ gH »—► g Xq is an immersion. Define φ-.g-* μ ρ »—»· ρ·ζο and note that Ф_1(яо) = Я. Set V = ker Ф*е С Te(G) = g, the Lie algebra of G, and let i) denote the Lie algebra of Я. Since Φ is constant on Я, it follows that Ф*е(Те(Я)) = 0, and hence f)CF. We are going to show that V С f). Since the left translations h ι—► p/i a: i—► pa; are diffeomorphisms, and since ΦοΙ5 = ί3 ο f, it follows that Ψ*ρ Ο Lg*e = Ьд*х0 ° Ф*е> and hence ker Φ*ρ = Ьр5)секег(Ф5)се) = 1^*е V for all g Ε G. Thus on G, Φ has constant rank r = dim g—dim V, and so by Theorem 1.1.31 the components of the level surfaces of Φ foliate G with codimension r. The component of the identity in Φ_1(χο) = Η is the identity component subgroup Я0 with Lie algera I). Thus V С ί) and so К = I). Now consider the following commutative diagram. \ψ.β ГН(С/Д) —> Τ^Λί)
§2. Principal Bundles: Characterization and Reduction 149 Since the two downward arrows in the diagram both have the kernel Lg*eV·> it follows that the bottom map is injective, so G/#-> Μ gH !-► g.Xo is an immersion. Step 3. If the original action is proper, then G/H —» Μ is a proper embedding. Suppose that μ: G χ Μ —» Μ is proper. Let С С М be any compact set. Then *-1(C) = {geG\gx0eC}, and the latter is compact since С and {xo} are compact. Hence the image of Ф_1(С) in G/# is also compact. But this image is just the inverse image of С under the map G/H -» M. Ш We now apply this lemma to obtain a simple but useful sufficient condition for the existence of a reduction. Proposition 2.14. Suppose that Η —> Ρ —> Μ is a smooth principal Η bundle. Let Q be a manifold equipped with a smooth, proper, right Η action. Let f:P —> Q be a smooth equivariant map (i.e., f(ph) = f(p)h for all h G H). Fix a point qo G Q, and set Щ = {ho G Η \ qoho = qo}· Suppose that qo G Q lies in the image under f of each fiber of P. Then (i) Pq = f~l(qo) is aw #o reduction of P. Fix another point q\ G Q, and set Hi = {h\ G Η \ q\h\ — q\}. Suppose that q\ G Q also lies in the image under f of each fiber of P, so that by (i) Pi — f~l(q\) is also an Hq reduction of P. Then (ii) there exists an element h G Η such that Η ι = h~lHoh and, for any such element h, we have P\ = P^h. Proof, (i) First we show that P0 is a submanifold of P. For this it suffices to show that / has constant rank. Since qo lies in the image under / of each fiber of P, it follows that the image of / is the orbit X = qoH С Q. By Lemma 2.13, since the action Q χ Η —> Q is proper, all the orbits, and in particular X, are proper submanifolds of Q. Since / takes values in X, it follows that rankp/ < dim X. We show equality by verifying that, for each fiber, rank(/ | fiber) = dim X. But the map that / restricts to on any fiber is, up to change of coordinates, the map Η —» X sending h —» qoh, which has rank = dim X by the lemma. Thus / has constant rank and Po is a submanifold of P. Next we note that
150 4. Shapes Fantastic: Klein Geometries heH0<=>q0h = q0<=>f l(q0h) = f l(q0) & f \qo)h = f l(q0) <=> P0h = P0, so we see that Pq is stable under the induced action of Щ. Let p,pf lie in the same fiber of P0. Then there is an h e Η such that p' — ph e Pq. Thus q0 = f(ph) — f(p)h — q0h, and so h e H0. It follows that the induced action of #o on Pq is transitive on each fiber of Pq. Since Pq = f~l(qo) is closed in Ρ and Hq is closed in H, the induced action of Ho on Pq is proper and free. By Theorem 2.4, P$ is a principal Hq bundle over Μ and is therefore an Щ reduction of P. (ii) This part is easy, and so we leave it to the reader. ■ Exercise 2.15. Suppose that Η —> Ρ —» Μ is a smooth principal Η bundle. Let ρ: Η —» Gf(V) be a representation. Give sufficient conditions for the following subsets S С Р to be reductions. (i) Let f:P —> V transform according to f(ph) = p(h~l)f(p). Let v0 e V, and take 5 = f~l(vo)· (ii) Let Grn(VQ = the Grassmannian of η-dimensional subspaces of V and ρ denote the standard action of Η on Gr^F) induced by p. Let f:P —» Grn(F) transform according to /(pft) = p(h~l)f(p). Let Vb G Grn(F), and take 5 = /"Ч^о). §3. Klein Geometries Now we are in a position to give the formal definition for Klein geometries, which we have roughly described in the introduction to this chapter. However, to flesh out the context in which they appear, we first prove the following result. Proposition 3.1. Let G be a Lie group and Η С G a closed subgroup. Then there is a unique maximal normal subgroup К of G lying in H. Moreover, К is a closed Lie subgroup of H, the left action of G on G/H induces a left action of G/K on G/H, and there is a diffeomorphism (G/K)/(H/K) —» G/H commuting with the canonical left G/K actions. Proof. Let К be the group generated by all the normal subgroups of G that lie in H. Then К is clearly a normal subgroup of G that lies in Η and is, moreover, the unique maximal normal subgroup of G that lies in H. Since the closure of К is also a normal subgroup of G that lies in Я, it follows that К itself is closed. Now apply the general result of Kuranishi and Yamabe (cf. [H. Yamabe, 1950]) that a subgroup of a Lie group is a
§3. Klein Geometries 151 Lie group11 to see that К is a Lie group. Finally, by Theorem 2.4, the canonical map π: G/K —» G/H is a principal fiber bundle with fiber H/K. Thus π induces a diffeomorphism (G/K)/(H/K) —» G/Я commuting with the canonical left G/K" actions. ■ Definition 3.2. A Klein geometry is a pair (G, Я), where G is a Lie group and Я С G a closed subgroup such that С/Я" is connected. G is called the principal group of the geometry. The kernel of a Klein geometry (G/H) is the largest subgroup Κ οϊ Η that is normal in G. A Klein geometry (G, Я) is effective if /Γ = 1 and locally effective if K" is discrete. A Klein geometry is geometrically oriented if G is connected. The connected coset space Μ — G/H is called the space of the Klein geometry or sometimes, by abuse of notation, merely the Klein geometry.12 A Klein geometry is called primitive if the identity component He С Н is maximal among the proper closed connected subgroups of G. A Klein geometry is called reductive if there is an Ad H-module decomposition g = \) 0 p, where g and \) are the Lie algebras of G and Я, respectively. * We use the remainder of this subsection to discuss some aspects of this definition. If (G, H) is a Klein geometry with kernel K, then, for any closed subgroup N С К that is normal in G, by Proposition 3.1 the pair (G/7V, H/N) is also a Klein geometry with space (G/N)/(H/N) и G/Я. Of course, these geometries are all ineffective except when N = K. This leads to the following. Definition 3.3. Let (G, H) be a Klein geometry with kernel K. Then the Klein geometry (G/K, H/K) is called the associated effective Klein geometry. Φ Exercise 3.4. Let Ke be the identity component of K. Show that (G/Ke, H/Ke) is a locally effective Klein geometry with kernel K/Ke. □ The reader may be wondering why all this fuss about ineffective Klein geometries. Proposition 3.1 seems to indicate that if we are interested in the geometry of the coset space G/H we need only consider the effective case. Why not define Klein geometries to be effective? The first point is that doing this would eliminate the subtle phenomenon of spin. While we do not deal with this notion in this book, we can keep it within the scope of 11 What they show is that a path-connected subgroup of a Lie group is a Lie subgroup. Since we allow uncountably many components for our manifolds, this is enough to ensure that any subgroup of a Lie group is a Lie subgroup. 12Note that the Euclidean plane and the affine plane have the same space (i.e., R2) even though the geometries are distinct.
152 4. Shapes Fantastic: Klein Geometries our definition by allowing our Klein geometries to be merely locally effective rather than demanding effectiveness. We may say that the geometrically important cases of Klein geometries are the locally effective ones. The second point to be made is that ineffective geometries, even ones that are not locally effective, do turn up, so we might as well have a language that allows us to speak of them. For example, of the planar geometries given in §1, only four were given with an effective description (Euclidean geometry, hyperbolic geometry, geometry of similarity, and Lorentz geometry). We could have done this for the others, too, but then we would have had to describe the kernels. The space G/H of a geometrically oriented Klein geometry need not be topologically orientable in the sense of Definition 1.1.9. For example, the effective version of the real projective plane has connected principal group G — PS73(R) and so is geometrically oriented, but the projective plane is not topologically orientable. On the other hand, consider the subgeometry (G, H) of the geometry of similarity, where Η is generated by SO2(R) and the dilation R2->R2 ν *-+ 2v and G — Η χ R2 (semidirect product). The space here is R2, which is topologically orientable, but since G has infinitely many components the Klein geometry (G/H) is not geometrically orientable. Note that G acts on the space R2 by transformations preserving the topological orientation. The real meaning of the notion of a geometrically oriented geometry will become clear only in §4 of Chapter 5 when we extend the definition to include Cartan geometries. The following result shows that every Klein geometry (G, H) determines a geometrically oriented geometry with the same space but with a possibly smaller principal group. Proposition 3.5. Let (G, H) be a pair of Lie groups, where Η is a closed subgroup of G, and suppose that G/H is connected. Let Ge be the identity component of G, and set Щ — Η DGe. Then (i) G = Ge-H, (ii) G/H = Ge/H0. Proof, (i) It suffices to show that G С Ge · H. Let g G G. Since G/H is connected, there is a path in G/H joining gH G G/H to eH G G/H. Since the map G —» G/H is a bundle, we can lift this path to a path joining g G G to some element h G H. Thus, gh~l G Ge and g = (gh~l)h G GeH. (ii) By (i) the map j: Ge —» G/H is surjective. Suppose g\, g2 G Ge. Then 92l9\ e Ge and j(9i) = ЗШ <=>9ι£ 92Η & g2~lgl G H.
§3. Klein Geometries 153 Thus, j induces a diffeomorphism Ge/H0 —» G/H. Ш Exercise 3.6. Let (G, H) be a pair of Lie groups, where Я is a closed subgroup of G. Let Ge be the identity component of G, and set Hq = HilGe. (i) Show that the set of products Ge · Η = {gh e G \ g e Ge, h e H} is a subgroup of G. (ii) Show that the inclusion Ge С G induces a smooth inclusion Ge/Ho С G/H whose image is a component of G/H. □ We remark that the bundle Ho —> Ge —> Ge/#o is an example of a reduction of the bundle Η —> G —> G/H. Lie-Theoretic Properties of'Klein Geometries As motivation for the bundle definition of Cartan geometries given in Definition 5.3.1, we draw the reader's attention to the following data associated to a Klein geometry (G, H): (a) the smooth manifold Μ = G/H; (b) the principal Η bundle Η С G -» G/H; (c) the Maurer-Cartan form ug'-T(G) —► g satisfying (i) cjg is a linear isomorphism on each fiber, (ii) Щис = Ad(h~l)uG for all h e Η (Proposition 3.4.2, part c)13, (Hi) cjg(X+) = X for all X e \) (cf. Exercise 3.2.15). The general usage in modern literature is to use the term homogeneous space for the coset space G/H, where nothing is assumed about the Lie groups Η С G, except that Η is closed in G (cf. e.g., [M. Golubitski, 1972]). Older literature uses the term Klein space. Geometrical Isomorphism and Mutation Definition 3.7. Klein geometries (G\,Hi) and (G2,H2) are called geometrically isomorphic if there is a Lie group isomorphism φ: G\ —» G2 such that φ(Ηχ) = H2. * In particular, the pairs (G, H) and (G, gHg~l) are geometrically isomorphic. A useful generalization of isomorphism is the notion of mutation. Actually, this is true for all h € G, but Rh arises from right Η action on the principal bundle
154 4. Shapes Fantastic: Klein Geometries Definition 3.8. Klein geometries (Gi,#i) and (G2,H2) are mutants (of each other) if there is an isomorphism of Lie groups φ: H\ —» H2 such that the induced algebra isomorphism </?*e:f)i —» f)2 extends to a linear isomorphism λ: gi —» £2, which is an ii module map (i.e., a map satisfying X(Ad(h)v) = Ad(ip(h))(\(v))). The pair (</?, λ) is called a mutation from (Gb^)to(G2^2). * Exercise 3.9. Show that (EuCn(R),SOn(R)), (SOn+1(R),SO„(R)), and (Ln>i(R), 50n(R)) are all mutants of each other. (They are Euclidean space Rn, elliptic space 5n, and hyperbolic space Hn, respectively.) □ Locally Klein Geometries Here is a generalization of the notion of a Klein geometry. Definition 3.10. Let (G, #) be a homogeneous space, and let Г С G be a discrete subgroup such that Г acts effectively by left multiplication as a group of covering transformations on the space G/H with Г \ G/H connected. Then the triple (Г, G, H) is called a locally Klein geometry. It is called geometrically oriented if G is connected. The double coset space Г \ G/H is called the space of the locally Klein geometry or, by abuse of notation, merely a locally Klein geometry. ft In the case that Г is the identity, this reduces to the Klein geometry (G,H). Note that in the definition of a locally Klein geometry (Г, G, H) it is not assumed that (G,H) is a Klein geomtery, that is, it is not assumed that G/H is connected. A consequence of the following exercise is that we can always replace (Г, G, H) by a locally Klein geometry (Го,Со,#) with the same space, and for which (Gq,H) is a Klein geometry. Exercise 3.11. Suppose that (Г, G, #) is a locally Klein geometry. Set Go = Ge · Я, where Ge is the identity component of G, and Г0 = G0 Π Γ. Show that Go is a Lie group, Go/Η is connected, and the inclusion Go С G induces a diffeomorphism ф: Г0 \ Go —> Г \ G. □ Just as in the case of Klein geometries, locally Klein geometries also have an associated principal bundle. This fact depends on the following lemma. Lemma 3.12. Let Г and Η be Lie groups with free commuting left and right proper actions (respectively) on the smooth manifold X. Then (Γ \ Χ) χ Η -> Γ \ X is proper ^ Γ χ (Χ/Η) -> Χ/Η is proper Proof. By symmetry it suffices to prove only the implication =Φ.
§3. Klein Geometries 155 Step 1. The action (ГхЯ)х1-> X ((g,h),x) »-+ gxh~l is proper. Let A,B С ХЪе compact, and let С = {(#,ft) G Γ χ Η \ gAhПВ/0}. We must show that С is compact. It suffices to show it is sequentially compact. Let {gi,hi) G Г х Я, г = 1, 2,... be a sequence in C. Let Л' and B' denote the images (which are compact) of A and Я, respectively, in Γ \ X. Since (Г\1)хЯ-> T\X is a proper action, the set C" = {ft G Я | A'ftnB' ^ 0} is compact. Moreover it is clear that hi G С", г = 1,2, Hence {ft^} has a convergent subsequence. Replacing (gi,hi) Ε С, г = 1,2,... by the corresponding subsequence, we may assume that (i) limftt = ftoo e С Since (gi,hi) G С we may choose points a^ G A and bi Ε Β such that gidihi = 6^, г = 1,2, Since Л and В are compact {a^} and {bi} have convergent subsequences. Replacing (дг,Ы) G С, г = 1,2,... by a corresponding subsequence, we obtain, in addition to (i), (ii) lima^ = a^ G A (Hi) Нт&г = Ьоо е В. Now let /i С X be a compact neighborhood of a^hoo so that a^fti G K", i> N. Since Γ χ Χ —> X is a proper action, the set C" = {g 6 Г | <^П {бос} φ 0} = {<? 6 Γ Ι ,Γ1^ € К) is compact. Now lim^ lboo = lim^ lbi = \im aihi = aooftoo G X. Thus g^lbOQ Ε К, г > Μ and hence gi e C", i > M. Thus {^} has a convergent subsequence, and we may pass to this subsequence to obtain, in addition to (i), (ii), and (Hi): (iv) lim^i = g^ G Γ It follows that the original sequence (gi, hi) G Γ χ Я, г = 1,2,... has a convergent subsequence, and hence С is sequentially compact. Step 2. The action Γ χ (Χ/Η) -> Χ/Η is proper. Let А, В С X/Η be compact. We must show that {g G Γ | gA Π Β φ 0} is compact. If we write A — Ui<2<a^b ^ ~ Ui<7<6 Bj-> w^n ^i and Bj compact, then it suffices to show that {g G Γ | gAi Π Bj φ 0} is compact for all г, j. Thus it suffices to assume that A and В are small sets (i.e., subordinate to some open covering of X/H). Now, by Theorem 2.4, the map X —* Х/Я is a principal Я bundle, so there is an open covering of
156 4. Shapes Fantastic: Klein Geometries X/H such that over each open set of this cover the bundle X —> X/H is trivial. We assume that A and В are subordinate to this cover. Applying a local section to A and B, we can obtain compact sets Af and B' in X with images A and В in Х/Я. Thus, {(g, К) е Г х Я | gA'h-1 Π 5' / 0} is compact. Therefore, the image of this set under the canonical projection Γ χ Η —» Γ is also compact; but this image is clearly {деГ\дАпВ^Щ. Ш Corollary 3.13. Let (I\G, Η) be a locally Klein geometry. Then the map Γ \ G —» Γ \ G/H is a principal Η bundle. Proof. Apply Lemma 3.12, taking X = G, and use the left and right multiplication actions of Γ and H, respectively, on G. Since Γ and Η are closed subgroups, these actions are proper. Since (Γ, G, H) is locally Klein, it follows that the action Γ χ (G/H) —► G/H is free and proper. So by the lemma, (Γ \ G) χ Η —» Γ \ G is also free and proper. The result then follows from Theorem 2.4. ■ We remark that, according to Exercise 3.11, the principal bundle Y\G —» Γ \ G/H is the same as the principal bundle Γ0 \ Go —► Г0 \ Gq/H. Example 3.14. Consider the case of the hyperbolic plane Μ — 5/2(R)/ S02(R). It is known that any closed, orientable surface Mg of genus g > 1 can appear as the quotient Γ \ Μ for some choice of Г С S^R)· Let Yg be such a subgroup. Then the quotient Mg = Yg \ Μ is a locally Klein geometry in the sense of the definition above. ♦ We note that if (Γ, G, H) is a locally Klein geometry, then the Maurer- Cartan form ug'-T(G) —> g, because of its left invariance, induces a form urXG:T(F\G)-+Q. Exercise 3.15. Show that the form ur\G has properties analogous to those of the Maurer-Cartan form given on page 153. □ Lie Algebras of a Klein Geometry For each Klein geometry (G, #), we have the corresponding pair of Lie algebras (g, i)). If {G,H) is effective, then \) contains no nontrivial ideal of g (cf. Exercise 3.4.7). Definition 3.16. An infinitesimal Klein geometry, or more briefly, a Klein pair, is a pair of Lie algebras (g, f)) where f) is a subalgebra of g. The kernel I of (g, f)) is the largest ideal of g contained in f). If £ = 0, we say the pair (g, \)) is effective. If there is an i)-module decomposition g = i) 0 p, we say (g,i)) is reductive. *
§3. Klein Geometries 157 Let us consider the correspondence {Klein geometries} —» {Klein pairs}. (G,H) ~ (fl,b) This correspondence is not surjective, since Remark 3.8.10 gives an example of an effective Klein pair (g, f)) for which f) cannot be realized as a closed subgroup in any realization of g. We could define the closure of \) of f) to be the Lie algebra of the closure of the realization of f) in the connected and simply connected realization of g. Then it is clear that the image of the correspondence consists of the Klein pairs (g, f)) where f) is closed (i.e., Ϊ) — f)). It is easy to see that the Lie algebra I of the kernel К of a Klein geometry (G, H) is the kernel of the associated Klein pair (g, f)). Exercise 3.17. Let (g, i)) be an effective Klein pair with \) closed. (a) Show that every realization is locally effective. (b) Show that there is an effective realization. (c) Study the relationship among the collection of all realizations of (g, \)). □ One-Dimensional Effective Geometries It is a remarkable fact that the real line supports exactly three effective Klein geometries. These are as follows. Definition 3.18. (i) The Euclidean line is (R, 0) with Klein pair (g,f)) = (R,0). (ii) The affine line is (Afft(R), Glf(JEL)), with Klein pair (g,f)) = (afh(R), flli(R)), where Afft(R)={(l 2)|ceR,dGR+J,G^(R)= j(j 2)|deR+}' aff+(R) = {(° ^|cGR,rfGR+},0[+(R)={(° y|d6R+}. (iii) The universal cover of the projective line is (G, H) with Klein pair (g,i)), where G — Sl2(R) = the universal cover of G = S72(R), Η = the identity component of the inverse image in G of Η С G, and where *={(V !;)hR*'&eR}·
158 4. Shapes Fantastic: Klein Geometries The Lie algebras of G and Η (and of G and H) are *={{~cd !0Ьйек}'Н("о" S)lMeR}· * The proof of this depends on the following classification of Klein pairs. Proposition 3.19. Let (g, f)) be an effective Klein pair with dim g/f) = 1. Then (g, f)) is isomorphic to one of the three Klein pairs described in Definition 3.18. Proof. See Appendix C. ■ We use this proposition to prove the following slightly more general result than that stated at the start of this subsection. Proposition 3.20. (i) There are exactly three isomorphism classes of effective Klein geometries on the real line: the Euclidean line, the affine line, and the universal cover of the real projective line. (ii) There are two families of isomorphism classes of effective Klein geometries on the circle. (a) One is the family of Euclidean circles (R/(/Z),0) with the same Klein pair as the Euclidean line. The parameter I G R+ is the length of the circle. (b) The other is the family of finite covers of the projective line (PS/2(R)(n), #(n)) with the same Klein pair as the projective line. The parameter η G Z"1" {the degree) indexes the subgroup пЪ 9* nm(PSl2(B.)) С π1(Ρ5ϊ2(Κ)) = Ζ determining the covering PS^CR-)^ —» PSfaiR), and H^ is the component of the identity of the preimage of Η in P5/2(R)(n). Proof. Suppose (G, H) is a one-dimensional effective Klein geometry with Klein pair (g, f)). Then by Proposition 3.19, (g, f)) must be the Euclidean, affine, or projective Klein pair. First assume that (g, f)) = (R, 0). Then the universal cover of G is R; in particular, G is abelian, and so Я is a normal subgroup. Since the geometry is effective, Η — 0. The preimage of Η in the universal cover is a discrete subgroup that must then take the form VL С R. Thus G = R/(iZ). This is the real line if / = 0 and is the circle otherwise.
§3. Klein Geometries 159 Next assume that (g, f)) is the affine Klein pair. Then g is realized by the connected and simply connected group Afff(R). As a simple calculation shows, Aff^(H) has no center and hence Aff^(R) is the unique realization of g. Now f) С g has G/^~(R) as its unique connected realization in Afff(R). Another simple calculation shows that G/+(R) is a maximal subgroup in Afft(R) and hence Я = G/+(R). Thus (G, Я) = (ДГ/+(К), G/^"(R)). In particular, G/H = R. Thus, there are no affine Klein geometries on the circle and just one on the line. Finally, assume that (g, f)) is the projective Klein pair. Then g is realized by the connected and simply connected universal cover of S72(R) denoted by 5/2(R). In fact, the projection S/2(R)_-> S/2(R)/{±/} = PS/2(R) is a 2-fold covering so that the composite S72 (R) —» 572(R) —* PS72(R) displays S72(R) as the universal cover of PS/2(R). Now the kernel of the covering homomorphism S72(R) —» PSi2(R)) is 7Ti(PS72(R)) = Ζ and is a central subgroup. On the other hand, PSi2(R) has no center, and so Ζ is the whole center of S72(R). Now if G is an arbitrary connected realization of g, there is a covering homomorphism p:G —» PSi2(R) determined by the subgroup p*K\{G) = nZ С ^(P5i2(R)) = Z, that is, by the integer n. Thus, we may write G = PS72(R)^· Since the pair (G,H) is effective, Η Π center(G) = {e}. Thus, Η is isomorphic to its image H0 in P5i2(R), and hence Щ contains the connected subgroup of PS72(R) with Lie algebra f). This latter subgroup is isomorphic to the image of 6GR, deR+j cSi2(R). under the projection map S72(R) —» PSi2(R). However, a third simple calculation shows that this subgroup is a maximal subgroup of S72(R). Thus, it is isomorphic under the projection map to Hq С PS72(R). In fact, since Щ is connected and simply connected, there is a canonical homomorphism Щ С PS72(R)(n) lifting the inclusion for each η and hence we have the following diagram. H0czPSi2(R)^R 55 Φ Φ >l· Hq^PSI^RJ^S1 «J, J, J, «-foldcover Я0е PSlflL) -> Sl From this it follows that there is just one isomorphism class of projective geometries on the real line, while for the circle there is one for each positive integer. ■ Exercise 3.21. Verify the three "simple calculations" referred to in the previous proof. □ ν ί
160 4. Shapes Fantastic: Klein Geometries Exercise 3.22. (a) Show that the Euclidean circles (Definition 3.18(ii), part a) are isomorphic as Klein geometries if and only if they have the same length. (b) Show that two projective circles (Definition 3.18(ii), part b) are isomorphic as Klein geometries if and only if they have the same degree n. □ Exercise 3.23. Show that the geometries in Definition 3.18(i) and (ii) are reductive but that in 3.18(iii) is not reductive. □ §4. A Fundamental Property In this section we prove a basic technical result about the connection between a Klein geometry and its associated infinitesimal geometry. Theorem 4.1. Let (G,H) be a Klein geometry with kernel К and associated Klein pair (g, f)) with kernel t. Then К = {h e Η \ Ad(h)v — vet for all ν e g}. Corollary 4.2 (Fundamental property of effective Klein geometries). If (G, H) is an effective Klein geometry, N is a subgroup of H, and η is the Lie algebra of N, then N = {he Η I Ad(h)v - ν e η for all υ e g} => N = {e}. Our proof of this result and its corollary depend on the following two lemmas. Lemma 4.3. Let η С f) С g be Lie algebras and N С Н Lie groups realizing the inclusion nc(). Assume N is normal in Η, and let N' = {he Η I Ad(h)v -v en for all ν e g}. Then Nf is also a normal subgroup of H. Proof, (i) Clearly e e TV'. (ii) Next, a e N' => Ad(a)v - υ e η for all υ e g => υ - Αά(α~ι)ν e Ad(a-1)n, for all υ e g; but TV is normal in H. Thus Ad(a-1)n С п. Hence α"1 e Nf. (Hi) α,βεΝ' => Αά(αβ)ν -ν = Αά(α)(Αά(β)ν - ν) + (Ad(a)v - υ) e Ad(a)n -f η = η for all ν e g. Hence αβ e N'. Thus TV' is a group. Finally, we verify it is normal in H.
§4. A Fundamental Property 161 (iv) aeN',heH=> Ad(h~lah)v -v = Ad(h-l){Ad(a)(Ad(h)v) - (Ad(h)v)} e Ad(h~l)n = η for all υ e д. Ш Lemma 4.4. Let f) С g be an embedding of Lie algebras and Η realize f). Define a sequence of subgroups of Η inductively by Νχ = {he Η I Ad(h)v - ν e n0 for all ν e g}, where no = Lie algebra of Nq = f), N2 = {he Η \ Ad(h)v - ν e tii for all υ e g}, where xi\ = Lie algebra of N\, Nk = {he Η I Ad(h)v -v e nk-i for all ν e g}, where π^_! = Lie algebra of Nk-\. Then Nq D Ni D N2 D ... D Nk D ... are all Lie groups that are closed and normal in Η and, after finitely many steps, the sequence stabilizes at a group N^ whose Lie algebra Поо is an ideal in g and satisfies N^ = {h e Η I Ad(h)v - ν e n^ for all υ e g}. Proof. Applying Lemma 4.3 inductively, we see that each of the groups Nk is normal in H. Also, if we assume Nj D Nj+ι (which holds for j = 0), then xij D tty+i, so that η e Nj+2 => Ad(n)v - ν e nJ+i for all υ e g => Ad(n)v — ν exij for all υ e g => η e Nj+ι, and hence A^+i D Nj+2- Now set TVqo = HTVj. Since each is Nj normal in Я, so is N^. We note that the sequence of Lie algebras xij and hence also the identity components of the NjS form a decreasing sequence which must become constant after finitely many steps. It follows from their definition that the JV^s themselves must also stabilize in the same finite number of steps, and so Noo is again a Lie group. Clearly N00 = {he Η \ Ad(h)v - υ e Поо for all υ e g}. ■ Corollary 4.5. Let Η be a closed subgroup of G realizing the inclusion f) С g. Assume that G is connected. Then in addition to the result of Lemma 4.4, TVqo ^ normal in G.
162 4. Shapes Fantastic: Klein Geometries Proof. We continue from the end of the proof of Lemma 4.4. Writing the condition NQO = {he Η \ Ad(h)v -v eN^ for all ν e $} infinitesimally, we see the Lie algebra Поо of N^ satisfies {η^,β} С 0, so that Поо is an ideal of g. By Exercise 3.4.7 the identity component of TVqo is a normal subgroup of G. For η G iVoo, we have Ad(n)(Ad(g)v) - Ad(g)v G n^, so that Ad(g-lng)v-v = Ad(g-l){Ad(n)(Ad(g)v)-Ad(g)v} G AdQT1)^ = η*,. (The latter equality comes from the fact that the identity component of Noo is already known to be normal in G.) Thus, g~lng G N^. It follows that TVqo itself is normal in G. ■ Proof of Theorem 4.1. It suffices to show that К — N^. The inclusion К D TVqo follows from the maximality of К (cf. Definition 3.2). We need to show that К С N^. The fact that К С N0 = Η is automatic. Let us assume inductively that К С Nj. Thus t С xij. Now the fact that К is normal in G implies, by Exercise 3.4.7, that &d(k)v — ν G Ϊ for all к G К and for all ν G 0. Thus, К С {h G Η | ad(ft)v - υ G t for all υ G g} С {h G Η | ad(/i)i; -vGtij for all ν e &} = A^+i- It follows by induction that К С TVqo? and hence K" = iVoo. ■ Proof of Corollary 4.2. The second half of the preceding argument shows that any subgroup TV satisfying the condition TV С {h G Η \ &d(h)v — v en for all υ e g} lies in N^. But TVoo = 1, since it is a normal subgroup of G lying in H. Ш §5. The Tangent Bundle of a Klein Geometry- Let us consider a bundle chart (U, ψ) for the principal Η bundle π: G —» G/Я. Thus we have the diffeomorphism ψ:17 χ Η —> n~l(U) С G. Clearly, this yields a commutative diagram. % tft/* x тгя* Т{п\и))<г^г T(UxH)-^T(U)xT(H)
§5. The Tangent Bundle of a Klein Geometry 163 The top right-hand diffeomorphism is the one described in Exercise 1.4.15. From this diagram it follows that, if χ = 7г(#), the diagram ι " π W :г—> 0 I 1 TX{GIH) I8 > fl/f, is commutative,14 with exact columns. Since f) is only a subalgebra of g (not an ideal), the quotient space g/f) is merely a vector space (not a Lie algebra). The linear isomorphism φ9 is the unique map making the diagram commute. We can identify the tangent space of G/H at χ with g/f), but the identification φ9 depends on the choice of g G G over x. In fact, since TiRfi — 7г, the relation Щ^н — Ad(/i)_1u;# implies that φ9\ι — Αά(Η)~ιφ9. It follows that the identification oiTx(G/H) with g/ί) is determined only up to the adjoint action of Η on g/ί). This fact accounts for the frequent occurrence of the adjoint action in the sequel. Proposition 5.1. T(G/H) и G χ я g/ί) (as vector bundles over G/H). Proof. Define a map φ: G χ g —» T(G/H) by <p(<7 χ ν) = (π(^),π*Ι/^*τ;), where π: G —» С/Я" is the canonical projection. Clearly, φ is smooth, sur- jective, and linear for fixed g. Moreover, for ν G Te(H), Lg*v G Tg(gH) G ker π*. Thus, (^(# xd) = (<7,0). There is also another fact about φ. We have ip(gh,ka(h~l)v) = (Tr(gh),K*Lgh*ka(h~l)v) = (n(g),K*(Lgh*Lh-i*Rh*v)) = (n(g),Lg*n*(Rh*v)) = (n(g),Lg*(nRh)*v) = (n(g),Lg*(n*(v)) =φ(9,ν). Thus, φ induces a smooth quotient mapping φ-.G χ я β/ϊ) —* T(G/H). Moreover, this map is injective since φ($ xv) — (p(gf χ ν') implies g' = gh for some h G Η and n*Lgh*vf — n*Lg*v. The latter means that n*Lh*vf = π*ν and hence that v' — Ad(h~l^v G \). Thus, Of course, the top map is not really uh except in the case when g G H. However, for any choice of gh G gH, the composite Tg(gH) 9—> Th(hH) = Th(H) —^> f) is the same, so we may sensibly denote this map by uh-
164 4. Shapes Fantastic: Klein Geometries gf χ v' = gh x Ad(h l)v (mod \)) = g xv in G xH g/f)· Thus, φ is a vector bundle map covering the identity on the base G/H, which is linear on fibers. ■ This proposition gives an identification of each tangent space Tg(G/H) with g/f). This identification is canonical up to the adjoint action of Η on g/f). Since tangential information involves only derivatives of the first order, we can make the following rough division of Klein geometries into two types according to whether or not Η is faithfully represented by the adjoint action on the tangent space. Definition 5.2. A Klein geometry G/H is οϊ first order if the representation Аад/ъ'.Н —» G/(g/f)) is faithful (i.e., injective). Otherwise, the geometry is said to have higher order. Φ As one may see in Appendix C, the classification of primitive effective pairs (g, f)) breaks into two cases of quite different nature depending on whether or not adQ^: f) —» G7(g/i)) is injective. §6. The Meteor Tracking Problem In order to motivate the need for the notion of a gauge, we begin by describing the meteor tracking problem in a Klein geometry Μ — G/H. Suppose we live in a small open set U С М and a meteor flashes through U. We wish to describe the motion of the meteor. We assume the meteor is rigid and sufficiently complicated. By the term rigid in the given geometry, we mean that for any two of its positions, where it has the configurations15 Xo and X\ say, there is an element of G carrying X0 to X\. By sufficiently complicated we mean that the subgroup of G that fixes the body pointwise (its stabilizer) is the identity. It follows that if X(t) is the configuration of its points at time £, then there is a unique path g(t) in G, the group of "rigid motions" for X, such that X(t) — g(t)X(0). One way to describe the motion would be to specify the path g(t) itself. However, it is more useful for us, and closer to what is actually observed, to describe the motion differently. We first describe the motion of one of its points q G X(0), which we take to be e#, by a path q(t) — g(t)H on [/, and then describe the motion of the rest of the body as turning about this one point as it moves along. To describe this turning, we need the notion of a gauge. 15If the "abstract" meteor is the point set X, then a configuration Xo of these points in Μ is a map Xo: X —> M. This is what Cartan originally had in mind when using the term "moving frame."
§6. The Meteor Tracking Problem 165 What Is a Gauge? The following figure depicts the anatomy of a generic gauge.16 Something fixed, but arbitrary ν^Τ^ζ (the choice of gauge) /^ ~ 3 д Something moving "V млллллл (what we wish to measure) We see that this structure consists of two parts: (1) a part that is fixed but arbitrary, the numbered marks (the "choice of gauge") (2) a part that is moving and that we wish to measure by comparison to the fixed part. In our example of the meteor tracking problem, the moving part is the meteor itself. What is the fixed part? Let us assume that the open set U is small enough so that there is a smooth section a:U —> G. Of course, once there is one σ there will be many others also, and if σ\ and σ<ι are two such sections, then they will differ by a map h:U —> H, namely, σ2{η) = ai(u)h(u). The "choice of gauge" is merely a choice of one of these sections. The section is the gauge and the relation between σι and σ2 above is called a gauge transformation. A choice of gauge can be regarded as a choice of motions, varying smoothly with u, which maps the base point to u. The value of a gauge at a point may be called a frame at that point and the gauge itself may be called a moving frame. From this point of view the principal bundle G —» G/H is a bundle of frames. Now let us see how we may use the gauge a:U —> G to track the meteor. Since both a(q(t)) and g(t) lie over g(£), they must differ by an element k(t) G H. Thus, in the presence of a gauge, the motion may be described by giving q(t) and k(t). The latter describes the way the meteor turns 16Hermann Weyl's original use of the term gauge was in the restricted sense of the gauge of a railway track and refers to a scale factor. This appears as the special case Η = R+. It is a happy accident of language that the term gauge also has the more generic interpretation given here.
166 4. Shapes Fantastic: Klein Geometries about the point q(t). We emphasize that there is no intrinsic geometric meaning attached to the choice of gauge, or indeed to the function k(t) alone. Together, however, they complete the description of the motion of the meteor. §7. The Gauge View of Klein Geometries In this section we are going to formulate various aspects of the Klein geometry Μ = G/H from the gauge point of view. These remarks will serve as the motivation for the base definition of Cartan geometries given in §1 of the next chapter, which studies Cartan's generalization of Klein geometries. Gauge Picture of Bundle Charts Fix a bundle coordinate chart (t/,V0 f°r the principal Η bundle n:G —» G/H. Thus we have the diffeomorphism ψ-.υ χ Η —> n~l(U). Note that, since such a chart is by definition right Η equivariant, specifying ψ is equivalent to specifying a section σ over U. More precisely, setting a(u) — ψ(ν,, e), we have ф(и, h) = a(u)h. Let us study the coordinate change resulting from a change of sections over U. Suppose that σι and σ<ι are two sections of G —> G/H over U. Thus, there is a unique smooth map k:U —► Η such that a2(u) — o\(u)k(u) for и G U. If φι and φ>2 are the corresponding trivializations of 7г_1(£/), we have the diagram UxH-^ n~l(U) ^-UxH. (u,h) »—у ai(u)h=a2(u)h «—< (u,h) It follows that ^Vi(^?^) — (^,σ2(ϊχ)_1σι(ΐχ)/ι) = (u,k(u)~lh). Thus, it is possible to reconstruct the bundle G —► G/H if we are given only the transition functions (or gauge transformations) k:U —► H. Gauge Picture of Maurer-Cartan Form Let us consider now the shape taken by the Maurer-Cartan form on a coordinate chart given by a section σ of G —* G/H over an open set U. The coordinate chart corresponding to σ is ψ:ΙΙχΗ-^π-ι(υ). (u,h) и-* Rhcr(u) We calculate ψ*(ωο). Proposition 7.1. The following diagram commutes, where φ(ν, у) = Ad(h)~luG(v) + uH(y).
§7. The Gauge View of Klein Geometries 167 тш(ихщ Яда Х Яда Ψ*ωα φ UU) χ щн) Proof. By Exercise 1.4.15, the vertical map is an isomorphism with inverse Tu(U)xTh(H)^T^h)(UxH), (v,y) *-+ ih*(v)+ju*(y) where г^: U —» UxЯ sends χ —» (χ, ft) 8inaju:H —> UxH sends A; —» (гг, fc). Thus it suffices to verify the commutativity of TiUth)(UxH) *h* V*c% Тш(ихН) J и* UU) (p\UU) = Aa{h-x)<»G ЩН) J*(DG rq>\Th(H)=(uH = o>G\Th(H) For the first diagram, we have (fwG)oi^ = (фгн)*ис = (Rha)*LuG = σ* R*huG = a*(Aa{h-l)uG) = Ad(ft_1)ff*u;G, and for the second we have (ψ*ω0) ο ju* = (ψ3η)*ωα = {La{u)YuG = uG. ■ Infinitesimal Gauge Let σ: U —» G be a gauge, namely, a section of the principal bundle G —» G//Z" over an open set C/ in С/Я". Then σ pulls back the Maurer-Cartan form uG to a g-valued 1-form θ on [/: σ*{ωα) = θ. The structural equation for uG also pulls back to yield άθ+-[θ,θ]=0.
168 4. Shapes Fantastic: Klein Geometries We note that not only does σ determine θ but the reverse is almost true as well, since the fundamental theorem of calculus says that θ determines σ up to left multiplication by a fixed element of G. Indeed, for an effective Klein geometry, if σ is a section, then да is a section <=> g = e.17 For this reason we may refer to both σ and θ as gauges on M. If we wish to distinguish them, we shall refer to θ as an infinitesimal gauge. In fact, it will be the infinitesimal form of the gauge that will be of the most use to us exactly because it is "independent of the base point." We may further note that since σ is a section, by the diagrams in §5 identifying (noncanonically) the tangent space of G/H with g/ί), the composite mapping canonical is an isomorphism for every и Εί/. The infinitesimal gauge θ can be roughly interpreted as assigning to each tangent vector ν G TU(M) the infinitesimal motion I -f εθ(ν) G G (where ε is infinitesimal and we are thinking of G as a matrix group so that the addition makes sense) of Μ whose effect on и itself is to move it to и -f εν. By varying the trivialization ψ, or equivalently by varying the section σ, we change the infinitesimal gauge Θ. We may see the variation explicitly in the following way. Let σ\ and σ2 be two sections over U. Then, as we saw above, there is a smooth map h:U —> Η such that a2(u) — o\{u)h(u) for и G U. According to the product rule (Proposition 3.4.10), differentiating this formula yields that is, 02-Ad^-1)^-^^ More explicitly, if ν e TU(U), then θ2(ν) = Ad(/i~1)i91(i;) +uH(h*(v)). We call such a variation of the infinitesimal gauge θ an {infinitesimal) change of gauge, and the two infinitesimal gauges are said to be (infinites- imally) gauge equivalent. We denote this symbolically by θι =>h θ2. It is quite clear from the "integral" version of the gauge that this is an equivalence relation, at least among gauges with a common domain of definition. 17If да is also a section, then χ = n(ga(x)) and hence gx = χ for all χ G U. Thus f^gf G Я for all / G tt'^U) С G. This shows that {/ G G \ f~lgf G #} contains an open subset of G. But it is also an analytic subset of G, so it must be equal to G, at least for G connected. Thus g G Π/eG/^/-1, and this latter is a normal subgroup of G in Η and is therefore trivial.
§7. The Gauge View of Klein Geometries 169 Exercise 7.2. Let uq be the right-invariant Maurer-Cartan form on G, and set θ = σ*ώο· Show that this (right) infinitesimal gauge transforms according to θ2 = #i + Ad(ai)/i*o;#, where σ2 = o\h. □ To prepare for later work with Cartan geometries, we make the following definition. Definition 7.3. A gauge symmetry of a Klein geometry (G, H) is a smooth map b:G —> G such that (i) b(gh) = b(g)h for all h e Я, and (ϋ) %) Ξ ^Я^"1 for all g e G. Φ Exercise 7.4. Show that a gauge symmetry is just a bundle automorphism of the principal bundle Η —» G —» G/Я. □ Exercise 7.5. Show that a gauge symmetry b:G —► G determines and is determined by a smooth map /: G/H —» G such that /(#) e gHg~l for all ££G. □
Shapes High Fantastical: Cartan Geometries In the wake of the movement of ideas which followed the general theory of relativity, I was led to introduce the notion of new geometries, more general than Riemannian geometry, and playing with respect to the different Klein geometries the same role as the Riemannian geometries play with respect to Euclidean space. The vast synthesis that I realized in this way depends of course on the ideas of Klein formulated in his celebrated Erlangen programme while at the same time going far beyond it since it includes Riemannian geometry, which had formed a completely isolated branch of geometry, within the compass of a very general scheme in which the notion of group still plays a fundamental role. -E. Cartan, 1939 The universe appears to be a nice mixture of homogeneity and nonho- mogeneity. At almost any location and scale, one is likely to see nothing (i.e., homogeneity) except for some concentrations ("lumps") of something at some great distance. If we move to the center of one of these concentrations, the nonhomogeneity may become more apparent; but if we then change our scale, the nonhomogeneity disappears and we are left with virtual homogeneity again. In a broad sense this is the fractal nature of the universe. Human life on earth is quite exceptional in this regard. We appear to be located at a position and scale where the nonhomogeneity is quite manifest. Nevertheless, we have not been so completely drowned in nonhomogeneity that we were prevented from discovering Euclidean geometry, a
172 5. Shapes High Fantastical: Cartan Geometries totally homogeneous idealization of our circumstance. Until this century, physics was regarded as the study of events in the amphitheater (as it were) of Euclidean geometry.1 The present century has seen the appearance of "lumpy geometry" in physics, both in Einstein's theory of general relativity as well as in the gauge theories of electromagnetism and of weak and strong interactions. Here geometry sheds its passive appearance as backdrop and assumes increasingly the role of actor. Indeed the question arises, "Is geometry in its various forms the only actor?" The new geometries in the quotation from Cartan at the head of this chapter refer to his "espaces generalises" or what we are calling here "Cartan geometries." If Klein geometries represent perfect homogeneity, then Cartan geometries represent a perfect mixture of homogeneity and nonho- mogeneity. Riemannian geometry (see Chapter 6) had its origins in 1854 in Riemann's celebrated talk.2 It can be regarded as a nonhomogeneous version of Euclidean space. Cartan's generalization of Klein's geometries do for them what Riemann's generalization does for Euclidean space, that is, it adds "lumps." If a geometric theory is going to be of any use in a seamless description of a nonhomogeneous universe, it had better be "lumpy." Physical gauge theories describe fermions ("particles") as complex-valued functions on a principal bundle over space-time, and bosons ("forces") as connections on this principal bundle (cf. [Y.I. Manin, 1989]). Thus, it may become physically interesting to understand these notions in their proper geometrical context. Cartan geometries are modeled on, and named by, Klein geometries. For example, we refer to a Cartan geometry modeled on Mobius geometry as a Mobius geometry. This usage is justified, for instance, by the fact (cf. Theorem 5.1) that if the Cartan geometry is flat, then it is locally the same as the model space. There are, in fact, two forms of Cartan geometries, the local form and the global form.3 The two forms are equivalent only when the model Klein geometry is effective.4 Our first definition is of local character and is called the base definition. Here the geometrical side is least apparent and analysis comes to the fore. This appears to be the version most preferred by the physicists because of however, Gauss had his suspicions about the truth of this. After his discovery of hyperbolic geometry, he realized that it could equally well serve as the amphitheater of events. In his capacity as director of the project to survey the Kingdom of Hanover, he was aware that the measurement of earthly triangles would not settle the issue. This story is told in Chapter 9 of [W.K. Buhler, 1981]. It was Lobachevski's proposal to attempt to settle the question experimentally by studying stellar triangles. 2An English translation of this talk appears in [M. Spivak, 1975], p. 135. 3See Appendix A for the relation to Ehresmann connections. 4 Although this is a very important case, it does not exhaust the interesting possibilities. For example, spin geometries are based on ineffective models.
§1. The Base Definition of Cartan Geometries 173 the local and analytical character of measurement. The second definition is global and is called the principal bundle definition. Roughly speaking, it presents a Cartan geometry as "a deformation of a Lie group which fixes the cosets of a closed subgroup." In §1 we use the gauge idea to generalize the notion of a Klein geometry (G, #) to that of the local version of "a Cartan geometry modeled on (G, H)" We also introduce the curvature 2-form, which is a measure of the failure of the structural equation. In §2 we see how, for effective geometries, a Cartan geometry determines a principal bundle and a Cartan connection on it. In §3 we give the global definition of a Cartan geometry in terms of the data obtained in §2. At the same time this definition generalizes the Lie-theoretic properties of the Klein geometries given on page 153. Section 3 continues with the study of a number of concepts related to Cartan geometry, including the tangent bundle, the curvature function, Bianchi identity, tensors, differentiation, and special geometries. Section 4 introduces the notion of developing a curve in a Cartan geometry as a curve in the model space. In the case of closed curves, this leads to the notion of the holonomy group. This group is a subgroup of the group G of the model. In the case of a complete flat geometry, it is just the monodromy group. It is one measure of the failure of the geometry to be Klein. Also in this section, the notion of geometric orientation is generalized to Cartan geometries to prepare the way for the classification in §5 of geometrically oriented locally Klein geometries among Cartan geometries. In §6 we study the Cartan space forms, a class of geometries generalizing the Riemannian space forms. These again turn out to be locally Klein, but the model may change. The ideas in this section revolve around the notion of model mutation, which refers to the possibility of altering the model Klein geometry on which a given Cartan geometry is modeled. It is a change that may alter the curvature but in which there is no loss of information about the original geometry. Finally, in the short §7, we apply the classification of Cartan space forms to the case of symmetric spaces. These latter can be defined for any reductive model. §1. The Base Definition of Cartan Geometries In Chapter 4 we showed how a Klein geometry Μ gives rise to a gauge, or rather an equivalence class of gauges, on each sufficiently small open set. We are now going to generalize this notion to that of a geometry modeled on a Klein geometry. But first there is the question of what data to take as the model geometry. The apparently simplest way would be to use a Klein geometry (G, H) itself as the model. This approach, however, would result in later difficulties related to the fact that the influence of the model is essentially local. For this reason we take the following definition.
174 5. Shapes High Fantastical: Cartan Geometries Definition 1.1. A model geometry for a Cartan geometry consists of (i) an effective5 infinitesimal Klein geometry (g, I)), (ii) a Lie group Η realizing i), (Hi) a representation, denoted Ad: Я —» С7ые((0> extending Ad^.H —> G/Lie(W- The kernel К of the representation Ad: Η —» С7ые(в) ^s called the kernel of the model geometry. If K is trivial, the model geometry is called effective. A model geometry is called primitive is the subalgebra f) is a maximal subalgebra of g. It is called reductive if there is an Η module decomposition fl = i) Θ p. * We note that the hypothesis of effectiveness in Definition 1.1 (i) implies that the kernel К is a discrete subgroup of H. It is clear that an effective Klein geometry (G, #), or more generally a Klein geometry with discrete kernel, canonically determines a model geometry. From now on we assume that we have chosen once and for all a fixed model geometry. Cartan Gauge Definition 1.2. Consider a model geometry (g, i)) with group H. A Cartan gauge with this model on a smooth manifold Μ is a pair ([/, #f/), where U is an open set of Μ and #f/ (which we may abbreviate by Θ) is a g-valued 1-form on U satisfying the regularity condition that canonical is a linear isomorphism for each и G U. (One usually assumes that U is a coordinate neighborhood in Μ, although this is not strictly necessary.) ft Definition 1.3. If Μ is a smooth manifold, then a Cartan atlas on Μ is a collection Л = {(t/a,0a)} of Cartan gauges with model (g, fj) and group Η such that (i) (Covering) the Ua's form an open cover of Μ, (ii) (Compatibility) if (U,9u), (V,6v) £ -Я, then there exists a smooth map fc: 17 Π V -» Я such that6 6>v = Ad(A;"1)i9{/ + /с*и;я onUnV. 5It is possible to omit the effectiveness condition here, but there seems to be no gain in doing so. 6This equation is an abbreviation. It means that for each χ 6 U Π V we have {θν)χ = AdMa;)-1)^)* + {k*)xoJH.
§1. The Base Definition of Cartan Geometries 175 (This compatibility relation is also called gauge equivalence.) Φ Let us take a moment to study the compatibility relation between θυ and θγ. As in the last chapter, we denote this relation by writing θυ =>k θγ. The following lemma is not unexpected since it is obvious for the original ("integral") version of gauges given in the last chapter. Lemma 1.4. Suppose that (0^, U) are Cartan gauges for i = 1, 2,3. Then (i) #i =>id #i, (ii) #i =>k #2 implies 02 =H-i #ъ (iii) 0i =>h 02 and 02 =H #3 imply θ ι =>hk 03. Proof, (i) is obvious. (ii) 02 = Ad^-1)^ + k*(u)h) implies 0i=Ad(fc)(02-fc>tf)) = Ad(fc) 02-Ad(k)(k*(LjH))· Now by the quotient rule, Corollary 3.4.11, Ad(k)(k*(uH)) = -(к~1)*ин, so 0X = Ad(k)02 + (k-l)*uH, i.e., 02 =»fc-i 0i. (iii) 02 = Ad(/i-x)0i + h*(u)H) and 03 = Ad(/c-1)02 + /с*(^я) imply 03 = Ad^XAdt/T1)^ + Л*(а;я)) + к*{шн) = Ad((/i/c)-1)01 + Ad(/T ^(ад) + **(уя). Now by the product rule, Proposition 3.4.10, applied to the composite U-^+UxU^HxH-^H, we have (hk)*uH = ((Λ χ к)А)У*ин = ((h χ к)АУ(п1Аа(к-г)шн + тг*и;я) = (πι (ft x /c)A)*Ad(fc_1)u;tf + (π2(Λ x к)А)*ин = Ad(fc_1)ft*u;tf + &*ω# and so 03 = Ad(ft/c)"1)01 + Ad^"1)/**^)) + k*(uH) = Ad{{hk)-1)el + {hk)*uH. Ш
176 5. Shapes High Fantastical: Cartan Geometries Exercise 1.5. Suppose that Ad: Я —► Gl(g/l)) is surjective. Let (£/, Θ) be a Cartan gauge and let x: U —» g/f) be any local coordinate system. Show that each point ρ EU lies in a neighborhood V С U such that the Cartan gauge (V,0 | V) is gauge equivalent to a gauge (ν,ψ) with φ = dx mod i). □ Just as in the case of manifolds, bundles, and foliations, we call two atlases equivalent if their union is also an atlas, and we note that there is a unique maximal atlas equivalent to a given one. Definition 1.6. A Cartan structure on a smooth manifold Μ consists of an equivalence class of Cartan atlases on M. A Cartan geometry7 is a smooth manifold Μ together with a specified Cartan structure. A Cartan geometry is effective if the model is effective. Φ Definition 1.7. Let Μχ and M2 be two Cartan geometries with the same model geometry. A diffeomorphism φ: Μ\ —» M2 is called an isomorphism of geometries (or a geometric isomorphism) if for any Cartan gauge (С/, вц) on M2, the gauge (φ~ι(ΙΙ),φ*θυ) induced on Μχ is compatible with the Cartan atlas on Mx. & Curvature We note that in defining the notion of a Cartan geometry, we have not mentioned the structural equation. This is not an oversight. Rather, it constitutes Elie Cartan's basic insight as the means for generalizing the Klein geometries. In particular, for any Cartan gauge (U,0u), the g-valued 2-form θ{/ = άθυ Η- |[0£Λ θ и] on U need not vanish. As we shall presently see, θ{/ is a measure of the nonhomogeneity—the lumpiness—of a Cartan geometry. Definition 1.8. The form θυ — άθυ Η- \[θυ,θυ] on U associated to a Cartan gauge (0, U) is called the curvature with respect to this gauge. * Of course, like the gauge вц, the curvature is not intrinsically defined. Let us see how it transforms as we alter the gauge. Lemma 1.9. Suppose that (Oi,U) are Cartan gauges on U for i = 1,2. Then 0\ =>h 02 implies θ2 = Ad{h~l)Q\. Proof. Calculating the exterior derivative of the equation 7Later in this chapter we shall give another definition (3.1, page 184) of a Cartan geometry equivalent to this one only in the effective case. The later definition is the definitive one.
§1. The Base Definition of Cart an Geometries 177 02 = Ad(h-1)e1+h*LjH, we get, using the formula for d{Ad(h~1)9i} in Exercise 3.4.14 and the symmetry of the bracket on 1-forms (Exercise 1.5.20(H)), d02 = Ad(h-1)de1-hAd(h-1)euh*uH]-l[h*uH,A It follows that θ2=<»2 + ^[02,02] = Ad{h-l)dex - ^[Adih^OiXuH] - \[h*u>H, АЩ/Г1^] + h*duH + ^[Ad{h-X)ei + /i*u;tf,Ad(/r1)01 +h*uH] = Adih-1) 1мг + \[θι,θΛ + ft* ldwH + ^я,"я]} = Ad(ft_1)6i since d<;# + -[u;#,u;#] = О by the structural equation given in §3 of Chapter 3. ■ A simple consequence of this is that the vanishing of the curvature is an intrinsic condition, independent of the choice of gauge. Definition 1.10. A Cartan geometry whose curvature vanishes at every point is called flat. Φ The analog for a Cartan geometry of the structural equation for a Lie group is flatness. While the strucural equation always holds for a Lie group, not all Cartan geometries are flat. Example 1.11. The simplest examples of Cartan geometries are, of course, the Klein geometries G/H themselves equipped with the Cartan gauges described in Chapter 4. We showed there that these charts were all regular and compatible. This is called the canonical Cartan geometry on G/H. ♦ Example 1.12. The next simplest examples are the open subsets of a Klein geometry. These clearly inherit the structure of a Cartan geometry by restriction. In general, these will not be Klein geometries. Nevertheless, they still satisfy the Maurer-Cartan equation, so they are flat Cartan geometries. Moreover, if such a flat Klein geometry is not simply connected, each of its covering spaces will have the structure of a Cartan geometry induced on it by the covering map. These geometries are again flat. ♦ Example 1.13. The final example we give now is the case of a locally Klein geometry T\G/H, where Γ is a subgroup of G that acts on G/H by left
178 5. Shapes High Fantastical: Cartan Geometries multiplication as a group of covering transformations. The transformations in Γ act as geometric isomorphisms on G/H. Since the projection G/H —» Y\G/H is a covering map, it is a local diffeomorphism, and we can transport the Cartan atlas from G/H to Γ \ G/H. The result is a Cartan atlas on Γ \ G/H, and the covering projection is locally a geometric isomorphism. In particular, the curvature on Γ \ G/H that vanishes so it is a flat Cartan geometry. ♦ Since all of these examples are flat, the reader may well be wondering if there are any nonflat examples. Of course, there are such examples. In particular, in §6 we will study cases where the curvature is nonzero but, in a certain sense, constant. More generally, it often happens that a submanifold of a Klein geometry can be canonically equipped with the structure of a Cartan geometry, and these are seldom flat. Examples of this are given in Chapters 6 and 7. Finally, we define the notion of a gauge symmetry of a Cartan atlas. Definition 1.14. Suppose that Я = {(E/Q,0Q) | a G A} is a maximal atlas for a Cartan geometry on the manifold M. A gauge symmetry of the Cartan geometry is a permutation b: A —> A such that (i) Ub(a) = Ua, and (ii) b satisfies the compatibility condition: if Ua = Up and θα =ϊ^ θβ, then 0b(a) =>h 0b(/3). * The full meaning of this rather subtle notion of symmetry will become evident only after we study, in §2, the principal bundle associated to a Cartan geometry. In any case it is clear from its definition that a gauge symmetry does not alter the geometry in any way. §2. The Principal Bundle Hidden in a Cartan Geometry In the case of a Klein geometry Μ = G/H, there is a canonical principal bundle η over Μ given by Η —> G —> Μ with the property that the Cartan gauges are all obtained by pulling back the Maurer-Cartan form on G via sections of η. It is natural to ask whether a similar bundle exists canonically for any Cartan geometry. The answer is that it does, provided the model geometry is effective, and in this section we describe it. The construction depends on the fundamental property of Klein geometries given in §4 of Chapter 4. Proposition 2.1. Let U support a Cartan geometry modeled on (g, \)) with group H. Let К be the kernel and let Qj (j = 1,2) be two compatible Cartan
§2. The Principal Bundle Hidden in a Cart an Geometry 179 gauges on U. Then θχ =>k #2 for a smooth function k:U —» Η that is unique up to multiplication with a smooth function l:U —+ K. In particular, if the Cartan geometry is effective, then к is unique.8 Proof. If k\ and k2 are two such fcs, then by Lemma 1.4, Thus it suffices to show that θ =Η θ implies к takes values in K. That is, we must show that every solution к to the equation θ = Ad(/Tx)<9 + k*uH, where k:U -» #, (2.2) takes values in K. Although this may look like a nasty problem in differential equations, it turns out that it can be solved by means of linear algebra by using the fundamental property of Klein geometries. Recall that we showed in Lemma 4.4.4 that the series of groups 7V0 = H, with Lie algebra no = F), TVi = {ft e Я I Ad(h)v — ν G no for all ν G g}, with Lie algebra tii, N2 = {ft G Η | Ad(h)v — ν G Πι for all υ G g}, with Lie algebra n2, Nk = {ft G Η | Ad(h)v — ν Ε η^_ι for all υ G g}, with Lie algebra n^, are all Lie groups that are closed and normal in H. They satisfy Л^ Э TVi D A^2 D ... D Ν^ D ..., and this chain stabilizes after finitely many steps at a group Л^ whose Lie algebra Поо is an ideal in g; moreover, Noo = {he Η \ Ad(h)v -v епоо^у е g}. In the present case, since η^ С I), it follows (cf. Definition 1.1 (i)) that Поо = 0. Thus, A^oo is discrete and Noo = {ft g Η | Ad(h)v -v = 0 for all ν G g} = ker(Ad:#->GI(fl)) = JT. We now show by induction that к takes its values in Ns, for 5 = 0,1, Since No = H, we see that к: U —» iVs for 5 = 0. Assume inductively that fc: E/ -» Ns. Fix w G t/. Then Eq. (2.2), rewritten as Ad(fc"ι)θ-θ = -ωΗΚ, says that for any ν G θ{Τη(ϋ)) we have Ad(k(u)~l)v — ve im8Lge(uuHk*u) С ns. 8We remark that this result is true even in cases where К is not discrete. Thus, it applies to the more general notion of a Cartan geometry with model as described in footnote 5.
180 5. Shapes High Fantastical: Cartan Geometries But since Ns is normal in #, by Exercise 3.4.7(c), the same conclusion also holds if ν G i). By the regularity condition on 0,g = f) + 0(Tu(U)), which implies Ad(k(u)~l)v -v ens for every ν G 0. Thus, fc(tx) G iVs+i. But w was arbitrary here, so in fact k:U-+ N3+x. It follows that к: U -+ TVoo = К. Ш The Principal Bundle of an Effective Cartan Geometry Now we are ready to pass to a description of the principal Η bundle associated to an effective Cartan geometry. Let Zl = {U} be a cover of Μ by sufficiently small, connected open sets (i.e., so that each is contained in the domain of a Cartan gauge and the intersection of any two of them is connected). The principal bundle we seek is obtained by glueing together the products U χ Η for all U G 11. For each U G £i, we choose a representative connection 1-form вц- Now if U\,U2 G U have corresponding forms #i,#2, then along U\C\U2 there is a gauge equivalence θ\ =>k θ2 given, according to Proposition 2.1, by a uniquely determined smooth map к: U\ Г1С/2 —► H. We can glue ί/ι χ Я to ί/2 χ Я along the common (U\ П[/2)хЯЬу making the identification (tx, /г) <-> (tx, k{u)~lh) as in Figure 2.2. It is easily checked that these identifications are compatible along the intersections of three open sets, so they fit together to give a right principal Η bundle P. Of course, the right action of Я in a coordinate patch is just right multiplication on the second factor, which clearly commutes with the identification maps. UiXH U2xH FIGURE 2.2.
§2. The Principal Bundle Hidden in a Cartan Geometry 181 By this construction, an effective Cartan geometry uniquely determines a right principal Η bundle π: Ρ —> M. Right Multiplication and Ad As we know, each vector ν G \) uniquely determines a left-invariant vector field V on the Lie group Η whose value at e is v. Such a vector field extends over U x Η as v^ = (0,V). Because the field V is left invariant and the bundle Ρ is made up of identification maps that alter the second factor only by left multiplication, the vector fields v^ on the coordinate patches U x Η fit together to yield a well-defined vector field v^ on Ρ itself. Lemma 2.3. Let R^: Ρ —> Ρ denote right multiplication by к G H. Then (Rk)*(vl) = (Ad(k-l)v)i for all ν e Ϊ). Proof. Since both sides are invariantly defined, we need only check the formula on a coordinate patch of the form U χ Η. In this coordinate patch, right multiplication takes the form Rk = id χ r^, where r^: Η —» Η is just right multiplication on H. Let lk'H—*H denote left multiplication on H. Thus, in our coordinate patch we have (Я*)*И = (id x rfc)*((0,V0) = (0,rfc*V0 = (0,rfc#/fc-i*^) = (Ad(fc-1)v)t, where we use the fact that V is a left-invariant vector field to write V = lk-**V. Ш The Cartan Connection In addition to the bundle Ρ just constructed, we also get a g-valued 1-form ω on P, called the Cartan connection, arising as follows. Given a gauge (17,0), we have the canonical linear isomorphism u;:T(Uth)(U x Я) can-^lcal TU(U) χ Th{H) -+ TU{U) χ ί) - g. (w,y) ■- («,wtf(y)) -+ Ac1(/i-1)0(i;)+u;h(2/) Let us verify that as we vary the gauge, these isomorphisms fit together smoothly to give a 1-form ω on P. We do this by comparing them by a transition function of the form f = (fi,f2):UxH^ UxH. (u,h) ь- (u,fc(u)-1/i) The derivative of this map is the vertical left-hand map in the following diagram.
182 5. Shapes High Fantastical: Cartan Geometries T(u,h)(U x H) > β (ν, у,) , > Adih^Mv) + ωΗ(Λ) i I! where T(uMuT%(U x H) > 9 (v,y2). > Κά(Κ\)θ2(ν) + ωΗ(γ2) Since f2(u,h) — k(u)~lh and f2*(v,yi) — 2/2, it follows from Proposition 3.4.10 ii) and Exercise 3.4.12 that ^(2/2) =ω//(/2*(ν,ί/ι)) = fe{uH){v,y{) = -Adt/r^X^H^ + ^H^i). This fact, together with the formula #2 = Ad(fc-1)^ + к*ин, verifies the arrow conjectured in the diagram above. Properties of the Cartan Connection Proposition 2.4. The Cartan connection ω on Ρ with values in g has the properties (i) for each point ρ G P, the linear map ωρ:Τρ(Ρ) —> g is an isomorphism, (ii) {Rh)*Lj = Ad(h-l)u>, (iii) ω(ν^) — ν for all ν el). Proof, (i) Since dim Ρ — dim Μ + dim Η — dim g/ϊ) + dim \) — dim g, it suffices to show that ωρ:Τρ(Ρ) —► g is a monomorphism for each p. Thus, it suffices to show that for ν G TU(U) and ?/ G Th{H), the equation Ad(/Tx )#(?;) +ия(2/) = ° implies (v,y) = 0. Now Ad(ft"ι)θ(ν) = -uH(y) lies in i), and since i) is Ad(#) stable, this means that θ(ν) e \). Thus ν = 0 by the regularity property for Cartan gauges. But then ин(у) — 0, and hence у — 0 since u;# is injective. (ii) We must verify the commutativity of the following diagram. T(u>h)(UxH) —— > g (y,y) . > Aa(h-x)e(v)^H(y) #** Ad(k~x)\ where ? Adtr1) T(uM{U x H) — > g (v, Rk* y). > AdWtyv) +Щ (#**y) But Ad(^)"1^) +а;я(Дк#2/) = Ad^XAd^"1)^) +uH(y))· (iii) The vector field v^ in a chart is given by (0,F), where V is the left-invariant vector field on Η corresponding to v. Thus, cj(0, V) = Ad(h-1)6(0)+iJH(Vr) = v. ■
§2. The Principal Bundle Hidden in a Cartan Geometry 183 Note that the Cartan connection ω on Ρ gives a canonical parallelization of the space Ρ and is the analog of the Maurer-Cartan form that appears on Ρ — G in the case of a Klein geometry Μ — G/H. Proposition 2.5. Let a:U —> Ρ be any section over U. (i) θ — σ*ω is a Cartan gauge on U compatible with the Cartan geometry on M. (ii) Let Ω be the ^-valued 2-form on Ρ given by Ω — άω + |[α;,α;]. Then (τ*Ω = θ, the curvature computed from the gauge σ*ω. Proof, (i) To see this we note that it is automatically true for any of the sections used to define Ρ and that the others differ from these by gauge transformations. (ii) Θ = άθ + \[θ,θ] = σ*(άω + %[ω,ώ\) = σ*(Ω)· ■ Proposition 2.6. There is a canonical one-to-one correspondence between gauge symmetries of an effective Cartan geometry and bundle automorphisms of the bundle P. Proof. Let b: Ρ —> Ρ be a bundle automorphism and ω the Cartan connection on P. Then b determines a gauge symmetry on the maximal atlas Я in the following fashion. Given a connected open set U С М over which Ρ is trivial, and a section σ: U —» P, we may compose σ with b to obtain another section ba: U —> P. Thus, we obtain two Cartan gauges, ({/, <τ*(ω)) and (U,a*(b*u)). We define the gauge symmetry by sending (υ,σ*(ω)) —► (U,a*(b*u;)). Compatibility condition (i) for gauge symmetries is automatic, and condition (ii) follows also, for if θα =>h θβ, then writing θα = σ*(ω) and θ β — <Τβ(ω), we have σβ = aah and hence οσβ = b(aah) = b(aa)h since b is a bundle map; thus, (6σα)*α; =>h (bσβ)*ω, which is condition (ii). Conversely, suppose that we are given a gauge symmetry b permuting the charts in a maximal atlas for a Cartan geometry. Then we may construct a bundle automorphism of the bundle Ρ as follows. Given a chart (ΙΙα,θα) G -Я = {{Ua,ea) | a G Я}, we have the associated chart (£4(а),#ь(а)) G Ά given by the gauge symmetry. Now Ua = Ub(a) = U, say. These two charts determine two trivializations (which are bundle maps), φα:ΙΙ χ Η —> Ρ and фъ(а)'-и x Η —> P. We define the bundle map over U by ψ^^Ψα1· Moreover, compatibility condition (ii) for gauge symmetries ensures that if (ΙΙβ,θβ) G Я is another chart with Ό β — C/a, then the same gauge transformation h giving θα =>h θβ also identifies their images under the gauge symmetry 9ца) =>h #ь(/3)· This means that ψβ(χ,ΐζ) — фа(х,Ь,(х)к) and Vb(/3)(s»fc) = Фъ(а)(х,Ь(х)к), and hence фц^Фа1 = Фъ^Фр1- Tnus> the bundle maps we have defined over our open sets U fit together to give a uniquely determined bundle map on P.
184 5. Shapes High Fantastical: Cart an Geometries Finally, we leave it as an easy task for the reader to verify that the two correspondences we have described are actually inverse to each other. ■ §3. The Bundle Definition of a Cartan Geometry Motivated either by the discussion above or by the Lie-theoretic properties of Klein geometries mentioned on p. 153, we are led to the following alternate definition of a Cartan geometry. Definition 3.1. A Cartan geometry ξ = {Ρ, ω) on Μ modeled on (g, \)) with group Η consists of the following data: (a) a smooth manifold Μ; (b) a principal right Η bundle Ρ over Μ; (c) a g-valued 1-form a; on Ρ satisfies the following conditions: (i) for each point ρ G P, the linear map ωρ:Τρ(Ρ) —» g is an isomorphism; (ii) (Rh)*u> = Ad(h~l)u for all h e #; (iii) cjpft) = χ for all X e\). By abuse of notation, we also speak of a Cartan geometry M. The g- valued 2-form on Ρ given by Ω = άω -f |[ω,ω] is called the curvature. If 9'· Q -* &/Ь is the canonical projection, then ρ(Ω) is called the torsion. If Ω takes values in the subalgebra f), we say the geometry is torsion free, or without torsion. The geometry is called complete if the form ω is complete, that is, if all the cj-constant vector fields are complete.9 We say a Cartan geometry is effective, primitive or reductive, respectively, if the model geometry is effective, primitive or reductive. ft A geometry in the global sense of Definition 3.1 determines one in the local sense of Definition 1.6, but the definitions are equivalent only when the model is effective. The discussion in §2 culminating in Proposition 2.5 of the previous section shows that every effective Cartan geometry in the original sense corresponds to an effective Cartan geometry in the new sense. Proposition 2.6 shows that the converse is true and that the two correspondences are inverse to each other. We shall henceforth take the present definition as the definitive one. It would be very interesting to have a definition of completeness in terms of Μ somewhat along the lines of Riemannian geometry, namely, something like completeness of geodesies. Cf. [B. Kamte, 1995].
§3. The Bundle Definition of a Cartan Geometry 185 Definition 3.2. Let (Ρχ,ωι) and (Ρ2,ω2) be Cartan geometries on Mi and Mi respectively, modeled on the pair (g, f)) with group H. Let /: Mi —» M2 be an immersion covered by an Η bundle map /: P\ —» P2 with the property that /*α;2 = ωχ. Then / is called a /oca/ isomorphism of geometries, or а /oca/ geometric isomorphism. If in addition / is a diffeomorphism then it is called an isomorphism of geometries, or a geometric1® isomorphism. Φ Exercise 3.3. If (Ρ,ω) is a Cartan geometry and b: Ρ —» Ρ is any bundle automorphism, show that (P, &*(u;)) is also a Cartan geometry. □ Note, in particular, that if ξ — (Ρ,ω) is a Cartan geometry on M, then it is geometrically equivalent to the geometry &*£ = (P, b*u)), where b: Ρ —> Ρ is an?/ bundle map. It turns out that the two Cartan connections will be equal if and only if the b is the identity (see Theorem 3.5 ahead), but it is clear that the geometries must be regarded as "the same" from the base definition of the geometry. The two Cartan connections ω and 0*ω are called gauge equivalent connections. Exercise 3.4. Verify that two effective Cartan geometries are isomorphic in the original sense of Definition 1.7 if and only if their corresponding geometries are isomorphic in the sense of Definition 3.2. □ We have the following uniqueness theorem for the bundle map covering an isomorphism of geometries. Theorem 3.5. Suppose that φ: Μ\ —» M2 is an isomorphism of effective Cartan geometries, and let fj'.Pi —► Pi (j = 1,2) be two Η bundle maps covering φ and satisfying //^2 = uji (j = 1,2). Then f\ = /2. Proof. Setting / = /х_1/2 (where /1_1:P2 —» P\ is the inverse of /1), we obtain an Η bundle map /: P\ —► P\ satisfying /*cji = ωχ and covering the identity map. Thus, it suffices to show that such a map is the identity. We set ω — ω\ and Ρ = Ρχ. Define ψ: Ρ —> Η by f(p) — ρψ(ρ). Let us factor / as f:pA+pxp^PxHJUp, where μ is the right multiplication map. Thus ω — (μ ο (id χ ψ) ο Δ)*ω = A*(idx ?/>)*μ*ω = A*(id χ φ)*{πΙΑά(φ~ι)ω + π^ωΗ] = Αά(ψ~ι)ω + ψ*ωΗ. Cartan uses the term holoedrique instead of geometric.
186 5. Shapes High Fantastical: Cartan Geometries It follows that Αά(φ~ι)ω — ω = —ф*ин- But then, proceeding as in the last paragraph of Proposition 2.1, we may show by induction that φ takes its values in 7VS, for s = 0,1,2, Since No = Я, we see that φ: Ρ -> Ns for s = 0. Assume inductively that φ: Ρ -> Ns. Fix ρ G P. Then the equation says that for ν G ωρ(Τρ(Ρ)) = g we have Αά(φ(ρ)~ι)ν — ν G image^tfV^x) С ns. Thus, ^(p) G iVj+i by definition. But ρ was arbitrary here, so in fact φ: Ρ —» TVj+i· It follows by induction that φ:Ρ —> N^. Since TVoo is a normal subgroup of G that lies in #, it follows that TVoo С if. Since we have assumed that the geometry is effective, ф(р) = 1 for all peP. Ш The following two exercises express obvious facts about the hierarchy of Cartan geometries on bundles intermediate between Ρ and M. Exercise 3.6. Let (Ρ, ω) be a Cartan geometry modeled on (g, f)) with group Η and let Б С Н be a closed subgroup with Lie algebra b. Show that (Ρ, ω) may also be regarded as a Cartan geometry on P/B modeled on (g, b) with group B. □ Exercise 3.7. Let Ρ be an Η bundle over Μ and let В С Н be a closed subgroup. Assume that (P, cj) is a Cartan geometry on P/P modeled on (g, b) with group B. Show that necessary and sufficient conditions on the form ω so that (Ρ, ω) is a Cartan geometry on Μ modeled on (g, f)) with group ϋΓ are that condition (c), parts (ii) and (iii) hold for all elements of Η and f), respectively, and not just those of В and b. □ Exercise 3.8. Resolve the following paradox. We have described above a procedure that yields a one-to-one correspondence: J Equivalence classes of 1 -> J Cartan geometries 1 \ Cartan atlases on Μ J ^7 [ (Ρ,ω)οηΜ J Let (Ρ, ω) be a Cartan geometry. If b: Ρ —> Ρ is any nontrivial bundle automorphism, set η = b*(u). Then, by Theorem 3.5, η φ ω, and by Exercise 3.4, (P, 77) is also a Cartan geometry. Now, given an open set U С Μ and a section σ: U —> Ρ, the forms cj and 77 pull back by σ to yield forms σ*ω and σ*(?7) = σ*(6*(ω)) = (6σ)*α;, which are compatible in the sense of Definition 1.3 since σ and ba are both sections over U (cf. Proposition 2.5). Thus β(Ρ,η) = β(Ρ,ω) and hence (Ρ,η) = αβ(Ρ,η) = αβ(Ρ,ω) — (Ρ, ω). This is in apparent contradiction to the fact that τ)φω. □ As one might expect from the strong form of the invariance of the curvature in the base definition, there is an analogously strong invariance in the bundle definition. We have the following result.
§3. The Bundle Definition of a Cartan Geometry 187 Lemma 3.9. Let (Ρ, ω) be a Cartan geometry on Μ modeled on (g, f)) with group H. Assume φ: Ρ —> Η is a smooth map. Define f:P —> Ρ by f(p) = R*(p)P· Then /*Ω = Αά(ψ(ρ))Ω. Proof. By the calculation in Theorem 3.5, we have f*u = Αά(ψ~ι)ω -f ф*и>н· Thus, by a calculation entirely analogous to the one in Lemma 1.9, we obtain /*Ω = Αά(ψ(ρ))Γί. Μ Corollary 3.10. The curvature form Q(u,v) vanishes whenever и or ν is tangent to the fiber. Proof. We may suppose that u,v G TP(P) are independent and that ν is tangent to the fiber. We may choose arbitrarily any map φ: (Ρ, ρ) —► (#, е) such that ψ*ρ(ν) = —ωρ(ν). (Since v is tangent to the fiber, ψ*ρ(ν) Ε f) = Te(#).) Define /: Ρ -» Ρ by f(q) = q-^{q). By the calculation in Theorem 3.5, and by the lemma, at ρ we have /*cj = Ad(^_1)o; + ψ*ωΗ =ω + ψ*ωΗ and /*Ω = Ω, so that up(f*v) — ωρ(υ) +инф*р{ъ) = ωρ(ν) —ωρ(ν) = 0, that is, f*v = 0. Thus, П(щ ν) = (/*Ω)(ΐ/, ν) = Ω(/*ϊχ, Дν) = Ω(Дu, 0) = 0. ■ Corollary 3.11. The curvature form Ω may be regarded as a 2-form on the pullback of the tangent bundle of Μ to P. Proof. By Corollary 3.10, the curvature may be regarded as a 2-form on the quotient bundle T(P)/ker π* that is canonically isomorphic to π*(Τ(Μ)). ■ We also obtain various foliations on the principal bundle P, as seen in the following exercise. Exercise 3.12.* Let Μ be a manifold and ξ = (Ρ, ω) a Cartan geometry on Μ modeled on (g, i)) with group H. (a) Let V be any vector subspace of the Lie algebra f). Show that u~l{V) is an integrable distribution on Ρ if and only if V is a subalgebra of (b) If ξ is torsion free and V is a vector subspace of g containing f), show that ω~ι(ν) is an integrable distribution on Ρ if and only if V is a subalgebra of g. □ The manifold Ρ may be regarded as some sort of "lumpy Lie group"11 that is homogeneous in the Η direction. Moreover, ω may be regarded 11 Of course, it is not a group.
188 5. Shapes High Fantastical: Cartan Geometries as a "lumpy" version of the Maurer-Cartan form. As we shall see in the following result, ω restricts to the Maurer-Cartan form on the fibers and hence satisfies the structural equation in the fiber directions; but when Ω φ 0, we lose the "rigidity" that would otherwise have been provided by the structural equation in the base directions and that would have as a consequence that, locally, Ρ would be a Lie group with ω its Maurer-Cartan form. (This subject is studied in detail again in §5.) Thus, the curvature measures this loss of rigidity. Lemma 3.13. Each fiber F of the principal bundle Ρ is canonically identified with Η up to left multiplication by some element of H. The Maurer- Cartan form ω η on Η induces a canonical form ωρ on F that agrees with the restriction of the Cartan connection ω on Ρ to F. Proof. By the very definition of a principal bundle, any fiber F of Ρ may be canonically identified with Η up to left multiplication by some element of H. This means that the left-invariant Maurer-Cartan form u;# on Η canonically determines a "Maurer-Cartan" form ω ρ on the fiber F. Again by definition, the vector field v^ on P, for ν £ f), restricts to a vector field on each fiber, which, under the canonical identification of F with Η (always modulo left multiplication), corresponds to a left-invariant vector field on H. Then the condition ω(ν^) —ν for all ν £ f) implies ωρ = ω \ F. Ш Exercise 3.14. Use this result to get a second proof that the curvature form vanishes when restricted to any fiber. □ Tangent Bundle of a Cartan Geometry In §4.5, we saw that the tangent bundle of a Klein geometry G/H can be expressed as a vector bundle associated to the principal bundle Η —> G —> G/H via the representation AdflA:tf-+End(fl/f>), so that T(G/H) и G x# &/Ь- We also saw the related fact that for each element g £ G there is a canonical linear isomorphism (pg:TgH(G/H) —» g/f) such that φ9\ι = Αά{Η~ι)φ9. These relationships continue to hold for Cartan geometries. This is expressed in the next result. Theorem 3.15. Let {Ρ,ω) be a Cartan geometry on Μ modeled on (g, f)) with group H. Then there is a canonical bundle isomorphism T(M) и Рхн g/f). Moreover, for each point ρ £ Ρ with π(ρ) = χ, there is a canonical linear isomorphism φρ:Τχ{Μ) —» g/f) such that φρ^ = Ad{h~l)ipp. Proof. Consider the following diagram. The columns are short, exact sequences and the two upper rows are isomorphisms, so there is a canonical isomorphism across the bottom making the diagram commute.
§3. The Bundle Definition of a Cartan Geometry 189 TP(pH) J ) f, I ^ ω W) —-—> 9 Ψ φ Φ Moreover, if г; G TX(M), we may write ν = π*ρ(ιι) = n*ph(Rh*u) for some и в TP(P). Thus, <PPh(v) = 4>ph{K*ph{Rh*u)) = p{uph(Rh*u)) = p(Ad(h-l)up(u)) = Ad(h-l)p(up(u)) = Aa(h~l)(pp(n*p(u)) = Αά(Η-ι)φρ(ν). It follows that we may define a smooth bundle map q:Pxg-+ T(M). (p,w) ·-+ (π{Ρ),φ-ι{ρ(ιυ)) Note that q(ph, Ad(h~l)w) = (n(ph)^^(p(Ad(h-l)w))) = (π(ρ),(Αά(Η)φρΗ)-ι(ρ(η>)) = (π(ρ),φΡι(ρΜ) = q(p,w). Thus, we get a canonical smooth bundle map q: Ρ x# 8/Ъ —> Τ (Μ). This is a vector bundle equivalence since it is an isomorphism on fibers and induces the identity map on the base M. Ш Corollary 3.16. Let (Ρ,ω) be a Cartan geometry on Μ modeled on (g, ()) with group H. The vector fields X on Μ are in bijective correspondence with functions f:P —» g/f) transforming according to the adjoint representation (i.e., f(ph) — AdQ/t)(h~l)f(p)). The correspondence is given by X^fx = {peP^ φΡ{Χ*(ρ)) e ф}.
190 5. Shapes High Fantastical: Cartan Geometries Proof, fx transforms correctly since fxiph) = ψΜ^(ρΗ)) = АЩ/Г1 νρ(Χπ(ρ)) = Adih-^fxip). Conversely, a function /: Ρ —> g/f) transforming according to the adjoint representation arises from the vector field X given by Xx — ψρ1{ί(ρ)), where ρ G Ρ is any point lying over χ G M. The choice of ρ does not matter since ^(/(pft)) = (Aa(h-^p)-l(Ad(h-l)f(p)) = φ-ι(№)· ■ Definition 3.17. Let (Ρ,ω) be a Cartan geometry on Μ modeled on (g, ()) with group H. A vector bundle Ε —» Ρ is called a geometnc vector bundle if it is given in the form Ε = Ρ x# У for some representation p: Η —> Gl(V). Example 3.18. Theorem 3.15 shows that T(M) is always a geometric bundle. ♦ Proposition 3.19. Let Μ be a connected manifold. Let (Ρ,ω) be a Cartan geometry on Μ modeled on (g, ()) with groups H. Fix χ Ε Μ and fix ρ Ε Ρ lying over x. Then ( Adg/b(h) G Gl+(a/b) for every Μ is orientable <=$ I h G Η such that ph G Ρ lies in the У same path components as ρ G P. Proof. Recall that Μ is orientiable if and only if every loop on Μ is orientation preserving (Proposition 1.1.17). Let λ: (/, 0,1) —» (Μ, χ, χ) be a loop on M. By Proposition 1.1.14, we may choose a partition of 7, 0 = to < t\ < · · · < tk — 1, and a compatible family of charts (Ui,i/>i), 1 < i < &» such that λ([ίΐ_ι,ίΐ]) С С/», 1 < г < к. Let σ: (7,0,1) —» (P,p,hp) be any lift of λ. We consider the linear isomorphisms R" ^ TaW(M) ^ 0/f), t e [ii-i.ti], l < i < k, and a basis for g/ί) so that this composite has positive determinant for t — 0. The compatibility of the family of charts together with the continuity of these maps implies that these maps have positive determinant for all t G I. Then the commutativity of the diagram R»«_!u. TAM)-^g/\j
§3. The Bundle Definition of a Cartan Geometry 191 shows that the loop λ is orientation preserving (i.e., ^fc+χ^ΓΛ nas positive determinant) if and only if Ad(h~l) has positive determinant. Thus, if Μ is orientable, then a path on Ρ joining ρ to ph projects to an orient able loop on M, and hence Ad(h) has positive determinant. Conversely, suppose Ad(h) has positive determinant whenever ρ can be joined to ph by a path. Then if λ is any loop on Μ based at x, we can lift λ to a path joining ρ and ph for some h e H. Since Ad(h) has positive determinant, it follows that λ is orientation preserving. ■ Definition 3.20. A Cartan geometry ξ — {Ρ, ω) on Μ, modeled on (g, f)) with group Я, is a first-order geometry if Ad: Я —» Gl(g/\j) is injective. Otherwise it is a higher- order geometry. & Exercise 3.21.* Let ξ = (Ρ, ω) be a first-order Cartan geometry on M, modeled on (g, fj) with group Я. Fix a basis (ei,..., en) for g/ί). Call the bases of TX(M) of the form (φ~ 1(ё\),..., φ~ 1(ёп)), where ρ Ε Ρ lies over χ, admissible frames. Let Q be the set of admissible frames over Μ with right Я action given by (<Ppl(ei), · · · > <Pp Чёп)) · Л = (^(ei), · ■ ·, ^(ёп))· Show that Q is a principal right Я bundle over Μ and that the map Ρ —> Q sending P^ (^(ei)»···»^1^)) is a bundle isomorphism. □ Curvature Function The curvature form Ω on the principal bundle Ρ determines a certain function К on Ρ called the curvature function. Definition 3.22. The curvature function K:P —> Hom(A2(g/i)),g) is defined by the formula К{р){ХъХ2) = ςΐρ{ω-ιΧι,ω-1Χ2). Lemma 3.23. The curvature function is well defined and satisfies the in- variance property K(ph)(XuX2) = Ad(h-l)(K(p)(Ad(h)XuAd(h)X2)). Proof. First consider ρ fixed and suppose that Yj — Xj + Vj, for j — 1,2, for some V0 e I). Then ίϊ^ω^Υχ,ω-^) = ftp(u-lXuu-lX2) by Corollary 3.10, since uj~1Vj is tangent to the fiber. Thus, К takes values in
192 5. Shapes High Fantastical: Cart an Geometries Hom(A2(g/i)),g). The invariance property follows from those of ω and Ω. Exercise 3.24. Let Μ be a manifold; let (Ρ*, ω*), г — 1, 2, be two geometrically equivalent Cartan geometries on Μ with curvature functions UQ, and let b: Pi —» P2 be a geometrical equivalence. Show that K2(b{p)) = Ki(p). □ As in Exercise 3.28 of Chapter 1, we may interpret the curvature function as a section, the curvature section, of the vector bundle, for ρ = 2, over Μ given by CP(M) = Ρ xH Hom(Ap(g/i)),g). The action of Η on Hom(A*(fl/fO,fl) is given by h · φ = Ad(/i)^(Ap(Ad(/i-1))). Exercise 3.25. Show that the two bundles CP(M) associated to two geometrically equivalent Cartan geometries on Μ are canonically isomorphic and that, for ρ = 2, this isomorphism identifies the two curvature sections. □ Exercise 3.26. Show that a Cartan geometry is torsion free if and only if the curvature function takes values in the subrepresentation Hom(A2(g/i)), 0)cHom(A2(g/0),g). □ Exercise 3.27. Show that K(p){X,Y) = [Χ,Υ] - ωρ([ω-ι(Χ),ω-ι(Υ)}). This identity interprets the curvature function as measuring the difference between the Lie algebra bracket and the bracket of the corresponding vector fields on P. □ Bianchi Identity Within the graded Lie algebra A(P, g) of g-valued exterior differential forms on any manifold P, various identities arise automatically. Here are some of them. Lemma 3.28. Let α, β G A(P,q) be homogeneous12 elements. Then ' deg β is odd, ' deg β is even. [αΛβ,β}\ = {ΐ[[α'β]'β] ;;de^isodd' Proof. The graded Jacobi identity (Exercise 1.5.20(iii)) yields (-ίγα[[β,β},α} + (-1)*[\β,α],β\ + {-1)α\[α,β],β\ = О, where α = deg a and b — deg β. By graded commutativity (Exercise 1.5.20(ii)), this is That is, each lies entirely within some grade of A(P,g).
§3. The Bundle Definition of a Cartan Geometry 193 (_l)ba(-l)2ba+1[a,[A/3]] + (-lf(-l)ba+1[[«^b/3] + (-l)ab[^/3]^]=0, which yields the result. ■ Corollary 3.29. [α, [α, α]] = 0 for every homogeneous element a G A(P,s). Proof. We may clearly assume that deg a is even, so that [α,[α,α]] = 2[[α,α],α] (by (3.28)) = — 2[α, [α, α]] (by graded commutativity). щ The following Bianchi identity is a formal consequence of the above identities and hence is true of the "curvature" associated to any element of degree 1 in Л(Р,д).13 Lemma 3.30 (The Bianchi identity). dVt = [Ω,ω]. Proof. Taking the exterior derivative of Ω = άω -f |[α;,α;], we get (cf. Exercise 1.5.20(i) dft = 0 -f -{[du;,u;] — [ω, do;]} = 2^Ω ~ 2^'^'^ ~ ^'Ω ~ 2^,α;^ = -{[Ω,ω] - [ω,Ω]} (by the corollary 3.29) = [Ω, ω] (by graded commutativity). щ Exercise 3.31. Show that [Ω,Ω] -f [[Ω,ω],α;] = |[Ω, [α;,α;]] by taking the exterior derivative of the Bianchi identity. (In fact, there is a whole sequence of derived identities obtained by successive differentiation of the Bianchi identity.) □ The original Bianchi identity of Riemannian geometry is usually expressed as two identities, called the first and second Bianchi identities. In that case the Lie algebra g is the Lie algebra of rigid motions of Euclidean space which decomposes canonically as a direct sum of a translation part and a rotation part. The two classical identities correspond to the two projections. See Chapter 6, section 2, for more details. More generally, it is true of the curvature associated to any element of degree 1 in any differential graded Lie algebra.
194 5. Shapes High Fantastical: Cart an Geometries Tensors Definition 3.32. Let ξ = {Ρ,ω) be a Cartan geometry on Μ modeled on (g, ()) with group H. Let V be a vector space and ρ: Η —> Gl(V) a representation. A tensor of type (V, p) is a function f:P—>V transforming according to the formula R*hf — p(h~l)f. Φ We have already seen several examples of tensors. For example, the curvature function of a Cartan geometry is a tensor of type (Hom(A2(g/i)), g), Hom(A2(Ad6/l)), Ad6)). Tensors of type (V, p) may of course be interpreted as sections of the vector bundle E(p) = Ρ χ # (V, ρ) (cf. Corollary 3.16 and Exercise 1.3.28, for example). How does a tensor of type (V, p) appear in a gauge? The following definition replies to this question. Definition 3.33. Let (t/,0) be a gauge of a Cartan geometry (Ρ,ω) corresponding to the section a: U —» Ρ (i.e., σ*ω = 0). If /: Ρ —» V is a tensor of type (V, p), then φ — fa: U —» V is called (ifte expression of) the tensor in the gauge (t/, 0). * Lemma 3.34. Lei φ:ϋ —>V be the expression of a tensor of type (V, p) m the gauge (C/,0). Lei 0 =>h 0', ft: C/ —» H, denote a change of gauge. Then р(ЬГ1)ф\ U —> V is the expression of the same tensor in the gauge (C/,0'). Proof. The change of gauge determined by h:U —> Η replaces the section a: U —» Ρ by aft: С/ —» Ρ, so that φ = fa gets replaced by (// = /(aft) = p(h-l)f(a) = p(h-^. Ш Universal Covariant Derivative Suppose V is a vector space and / G j4°(P, V) (i.e., / is a function on Ρ with values in V). Then the universal covariant derivative D is, roughly speaking, just the derivative of / with respect to the ω-constant vector fields.14 More precisely, if X G g, then Dxf = uj-l(X)f, so DX:A°(P,V)->A°(P,V). (3.35) Since this expression is linear in X, we may regard D itself as the adjoint operator D: A°(P, V) -> A°(P, V ® g*) defined by ix.{Df) = Dxf. (Note that ΐχ* is the coefficient homomorphism induced by 14E. Cartan studied this notion in the context of projective and conformal geometry in his papers [E. Cartan, 1934, 1935, and 1937].
§3. The Bundle Definition of a Cartan Geometry 195 tx:V ® β*-* V. For the details of this, see the subsection concerning change of coefficients in §1.5.) The universal covariant derivative is the grandfather of all "geometric" differential operators (i.e., operators with geometric meaning) in a Cartan geometry.15 For example, as we shall see below, the universal co- variant derivative, which exists in any Cartan geometry, gives rise to the usual covariant derivative if the geometry is reductive (cf. Definition 3.42). In Appendix D we show how the classical operators in the plane (i.e., div, curl, and the Cauchy-Riemann operator) arise from it via representation theory. Suppose that f:P —» V is a tensor of type (V,p). It is interesting to inquire after the transformation properties of Dp to see if it also may be interpreted as a tensor of some type. Definition 3.36. For any representation ρ: Η —» Gl(V), we denote by A*(P,(V,p))(=A*(P,p)) = {η: Xq(T(P)) ->ν\ϋ*Ηη = ρ{Η~ι)η, for all h G tf} the associated space of functions (if q — 0) or forms (if q > 0). Aq{P, V, p)) is called the space of q-forms on Ρ transforming according to the representation p. * (We shall deal mostly with the case of functions in our use of this definition.) Exercise 3.37. Show that there is a canonical isomorphism φ: Aq(P,p) w A°(P, p® A9(Ad*)). (Note that an example of this correspondence is given in the case q = 2 by Κ = φ(Ω); cf. Definition 3.22.) □ Lemma 3.38. D: A°(P, p) -» ΑΌ(Ρ, ρ <g> Ad*). Proof. Let / G A°(P, p); then for any ρ e Ρ, ν G g, we have ix4R*h(Df))(p) = LX*((Df)(ph)) = (Dxf)(ph) = «>;№)/■ Now the equation R^u = Aa{h~l)u may be read as uphRh* — Ad(/i_1)u;p for each p, and hence ω~£ — Κ^ω~λAd(h). Thus, 15Cartan's book [E. Cartan, 1938] (English translation, [E. Cartan, 1966]) is devoted to the study of differential operators in Riemannian and Lorentzian geometries from this point of view. For example, he is able to derive the Dirac operator.
196 5. Shapes High Fantastical: Cartan Geometries Lx.(RUDf))b>) = (Дь.КЧАсКВД))/. Since for any vector field У on Ρ we have (Rh.Y)f = Wh.Y) = (fRh)*Y = p{h'l)UY = pih-^Yf, taking Υ = ω"1(Α(1(/ι)Χ), we get LX4R*h(Df))(p) = pih-^-'iAd^X))/ = p(h-l)bKdWX}. Ш Note that even if the representation (V, p) is irreducible, the same need not be true of (V ® β*, ρ0 Ad*). For example, it may be that the latter decomposes as a sum of several representations V ® £* = W^. Θ... Θ Wfc. In that case D may be broken up correspondingly as a sum D = D\-\ \-Dk of several first-order differential operators where the Dj are the projections of D on the various summands. This is the way in which the Cauchy- Riemann operator, div, and curl are obtained in Appendix D. The following result calculates Df in the fiber direction. Lemma 3.39. For X e f) and f e A°(P,p), we have LX*(Df) = -p*(X)f, where p*: i) —» End(F) is the derivative at the identity of the representation p:H->Gl(V). Proof. (Lx.(Df))(p) = u~\X)f = Μω~ι(Χ)) -aL/e-*'*»-! = -p*(X)f(p). p(exp(-tX))f(p) Exercise 3.40. Exercise 3.37 allows us to extend the definition of the universal covariant derivative to forms on Ρ with values in (V, p) to yield a map D:Aq(P,p) —» Л9(Р, ρ 0 g*). Suppose 77 is any 9 form on P. Show Όη — 0 if and only if 77 may be expressed as a linear combination, with constant coefficients, of exterior products of the components of the Cartan connection ω with respect to some basis of g. □ Because of the canonical identification A°(P, p) — A°(M,E), where Ε is the vector bundle Ρ χ я (V, ρ), the universal covariant derivative associated to (V,p) may equivalently be regarded as a linear first-order differential operator D:A°(M,E) -» A°(M,F), where F = Ρ xH (V (g) g*, ρ 0 Ad*). We note that this operator gives the derivatives of a section of Ε not in a direction given by a tangent vector of Μ but in a direction given by a tangent vector of P. Another way to say this is that to describe how a
§3. The Bundle Definition of a Cartan Geometry 197 section changes from point to point it is not enough to give the two nearby points of M; we must give two nearby frames of Μ (infinitesimal meteors in the parlance of §4.6). Exercise 3.41. Show that, if Ρ is connected, then ker(L>: A°(M, E) -» A°(M, F)) = VH where VH = {veV\ p(h)v = ν for all h Ε Η}. □ This exercise shows that, apart from the constant functions, there are no tensors on Μ that are "generalized covariant constant" functions. Finally, we remark that if the representation V0g* decomposes as a sum of representations V ® β* = W\ Θ ··· Θ Wk, then the bundle F decomposes correspondingly as a sum F — F\ Θ ··· Θ i^, where Fi — Ρ Хн Wi. Moreover, the constituent operators in the decomposition D = D\ + --· + Dk may be regarded as first-order linear differential operators Dk: A?(M, E) -» Ap(M, Fi). Covariant Derivative in a Reductive Geometry The covariant derivative D (see Definition 3.47) generalizes the vector gradient in Euclidean space (see Exercise 3.50). However, the existence of a covariant derivative requires that the geometry be reductive.16 Definition 3.42. A Cartan geometry modeled on (g, fj) with group Η is reductive if there is an Η module decomposition g — \) Θ ρ (cf. Definition 4.3.2).17 * Let us assume now that the geometry we are considering is reductive in this sense. Any form with values in g will decompose into an i) component and a p component. This is true in particular for a Cartan connection (ω = ω^ Η- ωρ) and any gauge (0 = вц Η- θρ). In addition, the universal covariant derivative also decomposes as Ό χ = Ό^χ Η- Όρχ. By Lemma 3.39, for X e j) and / e A°(P,p), iXJ(Df) = -p(X)f, namely, D^ = —p. That is, D^ merely tells us how / transforms under 6The reductive geometries have, therefore, a much richer structure than the nonreductive ones. 17This use of the term reductive conflicts with another use of it in the theory of Lie algebras, which calls a Lie algebra reductive if it is the sum of a semisimple ideal and an abelian ideal. The two notions are loosely related in the following sense. Given a Klein pair (g, Ϊ)) for which the representation ad: \) —> βί(β/ϊ)) is injective and \) is reductive in the Lie algebra sense, then (g, Ϊ)) is reductive in the sense of Definition 3.42 (cf., e.g., [J. Humphreys, 1972], pp. 31 and 102, and [W. Fulton and J. Harris, 1991], p. 131).
198 5. Shapes High Fantastical: Cart an Geometries Я, which we already know, so this projection is not very interesting. The component Dp, on the other hand, is very interesting. Definition 3.43. In a reductive geometry the operator Dp is called (the bundle version of) the covariant derivative. A function /: Ρ —► V is called covariant constant or parallel if Dpf = 0. ft Since sections of geometric vector bundles over Μ can be interpreted as functions on P, it follows that this notion of parallel applies also to these sections. We wish now to go into more detail on this point in order to reinterpret the covariant derivative (as well as the notion of parallel) at the level of sections. The key idea is to use the reductive hypothesis again to get a canonical way of lifting a vector field on Μ to a vector field on P. Definition 3.44. The distribution on Ρ given by ω^ =0 is called the horizontal distribution and vectors in it are called horizontal vectors. ft Note that a function /: Ρ —» V is parallel if and only if it is constant in the horizontal directions. We now relate these horizontal directions to the directions in the base M. Lemma 3.45. The projection π: Ρ —> Μ induces an isomorphism between the vector space of horizontal vectors at ρ Ε Ρ and Τπ(ρ)(Μ). Proof. From the exactness of the columns in the following diagram and from its commutativity Tp{pH) J ) f, 1 TP(P) % > β = ί)®Ρ ψ φ ψ (see the proof of Theorem 3.15), the lemma is immediate. ■ Definition 3.46. Let (Ρ, ω) be a reductive Cartan geometry on M. If X is a vector field on M, then the horizontal lift of X, denoted by X, is the (unique) vector field on Μ such that π*ρ(Χ) = Χπ(ρ) for all ρ G Ρ and ub(X) = 0. ft Definition 3.47. Let (Ρ, ω) be a reductive Cartan geometry on Μ, Ε — Ρ χ н (V, ρ), and Χ e Γ(Τ(Μ)). Let φ: Γ (Ε) и А° (Ρ, ρ) be the usual
§3. The Bundle Definition of a Cart an Geometry 199 identification of sections of vector bundles associated to Ρ with functions on Ρ (cf. Exercise 1.3.28). The (base version of the) covariant derivative Όχ:Γ(Ε) -» Γ {Ε) is defined by the equation i/>(Dxf) = Χ(ψ(/)), where / e Γ(Ε). * Proposition 3.48. Let k: Μ -» R, Χ, Υ e Γ(Τ(Μ)), and f,g e T(E); then the covariant derivative has the following properties: (i) Dx(f+g) = Dxf + Dxg; (ii) Dx+Yf = Dx! + Dxf; (iii) DkXf = kDxf; (iv) Dx(kf) = X(k)f + kDxf. Proof, (i) and (ii) are straightforward from the R-linearity of ψ and of the R-bilinearity of the derivation X. For (iii) we note that ж*(к)Х is the horizontal lift of kX, where π: Ρ —> Μ is the canonical projection. Thus, tl>(DkXf) = А,.(к)*Ш)) = *4*)£>χ(ψϋ)) = n*(k)ip(Dxf) = tKkDxf) whence (iii). Finally, t(Dx(kf)) = XMkf)) = Х(п*(к)ф(/)) = Χ(π*(Α)Μ/)+π*(*)ΧΜ/)) = (n*X{k))t{f)+T'{k)t(Dxf) = t{X(k)f) + u(kDxf), which, by the linearity and injectivity of ψ, yields (iv). ■ Although we see some of the properties of D from Proposition 3.48, it is still not clear how to calculate it in terms of a gauge. The following result remedies this. Proposition 3.49. Let ({/, Θ) be a gauge for a reductive Cartan geometry, X be a vector field on U, φ be the expression in the gauge (t/, 0) of a tensor of type (V, p). Then Όχφ = Χ(φ) — р*(вц(Х))ф is the expression in the gauge ([/,Θ) of the covariant derivative of the tensor expressed by φ in the gauge ([/, Θ). Proof. Let a:U —> Pjj = n~l(U) be the section corresponding to the gauge ([/, Θ) and let Φ: Ρ —► V be the tensor of type (V, p) of which φ is
200 5. Shapes High Fantastical: Cartan Geometries the expression. If σ*(Χ) were horizontal, then our job would be easy since then it would be the horizontal lift of X and we would have (Όχφ)(χ) = (ΌχΦ)σ(χ) = Χσ(χ)(Φ) = Φ.(Χσ(χ)) = Φ*(σ*(Χ,)) = (Φ ο σ)*(Χχ) = φ*{Χχ) = Χχ(Φ)- This does indeed yield the formula claimed when σ*(Χ) is horizontal since in that case the expression fy (X) vanishes. The general case will come from making a change of gauge so that σ*(Χ) is horizontal in the new gauge σ'. We have Θ' = Ad(h~l)9 + Л*(о;я)· Applying this to X and taking the \) components yields ffb{X) = ka{h-l)eb(X)+uH(KX). We may choose h: U —► Η so that at a fixed but arbitrarily chosen point χ we have h(x) — e and h*X = —Θ^{Χ). Then, at x, we have Όχφ' = Χ(φ') = Χ{ρ{Η-ι)φ) = Χ{ρ{Η-ι))φ + ρ(Η-1)Χ(φ) = -(ρ(Η)*(Χ))φ + Χ(φ). The left- (and therefore right-) hand side of this equation is ΌχΦ expressed in the gauge ([/, Θ'). By Lemma 3.34, and because h(x) — e, this is the same, at x, as the expression for ΌχΦ in the gauge (U,0). Ш Exercise 3.50. Show that in Euclidean space the covariant derivative of a vector field is just the usual vector gradient. [Hint: Apply Proposition 3.49 using the gauge given by θ(βι) = ί n J G eucn(R), where e» is the standard column vector.] □ For simplicity, we have expressed the covariant derivative as a map Όχ\ Τ (Ε) —» Γ (Ε), where Χ is a vector field on X. However, we note that this operator is local in the sense that the value of Dxf at χ depends only on the value of X at χ and the behavior of / on a neighborhood of χ in M. (In fact, only the zero- and first-order terms of the Taylor series of / at χ play a role in the contribution of / to this formula. Cartan would say that the formula depends only on the values of / in a first-order neighborhood of χ in M.) These and similar facts may be invoked to justify the following definition and exercise. Definition 3.51. A section η G T(E(p)) is called parallel along a curve σ:Ι -» Μ if Όσ^η = 0 for every t el. * Exercise 3.52. Show that a function f:P —► (V,p) is parallel in the sense of Definition 3.43 if and only if the corresponding section η G Y{E(p)) is parallel along every curve in M. □
§3. The Bundle Definition of a Cartan Geometry 201 Special Geometries In general, the values of the curvature form will span, as a vector space, the whole of the Lie algebra g. In special cases, however, this span may not be the whole of g. For example, in the case of a torsion free geometry, the span will lie in f). However, as the following lemma shows, the span cannot be an arbitrary vector subspace of g. Lemma 3.53. Let V С g be the vector subspace spanned by the values of the curvature form Ω. Then V is an Η submodule of g. Proof. It suffices to show that the set of values of Ω is stable under the adjoint action of Я. Let ν = Ωρ(Χρ,Υρ). Then Аа(к~1)у = Ad(h-l)(np(Xp, Yp)) = {R*hilp)(Xp, Yp) = Qph(Rh*Xp, Rh*Yp) = a value of Ω. ■ In particular, if V С f), so that the geometry is torsion free, then V is an ideal in \). Definition 3.54. Let V С д be an Я submodule. A Cartan geometry of g-curvature type V is a geometry whose curvature form takes values in V. Assume now that our geometry is torsion free and the adjoint action of Η on f) is irreducible. In this case, there are no special geometries arising from g curvature-type conditions. Nevertheless, the representation of Η on Hom(A2(g/i)), i)) may have nontrivial submodules V so that one may distinguish various cases according to how К relates to these submodules. The following definition carries the same idea as Definition 3.54 but the context is different. Definition 3.55. Let V С Hom(A2(g/i)), i)) be an Η submodule. A Cartan geometry of curvature type V is a geometry whose curvature function К takes values in V. * By Exercise 3.4.8(b), Ноть(А2(д/{)), ij) С Hom(A2(g/i)), ij) is a submodule and, for Η connected, it is the submodule of Η invariant elements. If the submodule is nontrivial, this leads to a class of special geometries, the constant-curvature geometries, studied in §4. When Η is a compact group, it is known from the representation theory of Lie groups that Hom(A2(g/i)), i)) decomposes as a direct sum of irreducible submodules. In this case C2{M) and its section К will split up correspondingly, and we may speak of the various "component curvatures" associated to K. For example, in the general Riemannian case C2(M) splits into three pieces corresponding to the scalar, the Ricci, and the Weyl curvatures (cf. Table 2.5 on page 236). In the special case when Μ has dimension 4, the Weyl curvature splits up further into a self-dual and an anti-self-dual part.
202 5. Shapes High Fantastical: Cartan Geometries When Η is noncompact, even though Hom(A2(g/f)), i}) may not in general decompose into irreducibles, it may still have nontrivial submodules. For example, the composite mapping Hom(A2(0/f,), {,) « A2(0/f))* ® t, Ы-^ \2{фГ <g> End(fl/{,) и λ2(0/())* ® Φ ® (fl/W* - (fl/W* ® (fl/W* i*Au*(g)v(g)ii;* ι-»· (t*(v)u*-u*(v)t*)®w* is an if module homomorphism, and so its kernel is a submodule. Definition 3.56. The kernel of the composite mapping above is called the normal submodule o/Hom(A2(g/f)), i)). ft Exercise 3.57. Show that if End(g/i}) is given the Η module structure h-φ = Ad(/i)0Ad(/i_1), then the map ad: i) —» End(g/i}) and the canonical map End(g/i)) —» g/f) 0 (β/f))* are Я module maps. Use these facts to verify that the composite mapping above is an Η module map. ft The normal submodule will be useful in defining normal geometries. In some cases—for example, in conformal geometries—"normal" means that К takes values in the normal submodule. In other cases, such as projective geometry, it means that К takes values in a submodule analogous to the normal submodule. In general, however, a normal geometry is defined in a somewhat ad hoc manner so that the Cartan geometry will be uniquely determined by a given set of geometric data that may seem at the outset to have no necessary connection with a Cartan geometry at all but determines one via Cartan's method of equivalence. Several examples of this are given in Chapters 6, 7, and 8 on specific geometries. §4. Development, Geometric Orientation, and Holonomy In this section we study properties related to paths in a Cartan geometry. In particular, we introduce the notion of geometric orientation18 in this context. Its importance for us is in connection with the classification of locally Klein geometries given in Theorem 5.4. Then we discuss issues connected with the development of paths in a Cartan geometry as paths on the model space. In the last part of this section, we apply the notion of development to introduce and to discuss some issues surrounding the holonomy group of a Cartan geometry. Our notion of a geometric orientation differs from the notion of the same name discussed in [E. Cartan, 1941].
§4. Development, Geometric Orientation, and Holonomy 203 Recall the notion of development given in Definition 3.7.4 in connection with the global version of the fundamental theorem. In that case we assumed we had a Lie group G with Lie algebra g, a manifold Ρ equipped with a g-valued 1-form ω, a piecewise smooth path σ: (/, α, b) —» (Ρ, ρ, g), where I — [a, b]. We showed that, given g e G, there is always a unique path σ: (7, α) —> (G, #) such that σ*ωο = cr*cj. This path on G was called the development of ω starting at g. Recall that ad ω takes values in the Lie algebra End(g) of Gl(g) (cf. Exercise 3.4.9). Lemma 4.1. Let λ be a path on Ρ and let (i) λ: (I, a) —> (G, e) be the development of λ via ω, (ii) λ: (I, a) —> (G/(g),e) бе £Ле development of λ via ad ω. Then X(t) = Ad(\(t)) for t e I. Proof. Since, by Proposition 3.1.8, the homomorphism Ad: G —> Gl(g) has the property that Αά*ωοΐ($) = Ad*ea;G, it follows that Αά(\)*ωοι(9) = A*Ad*o;G/(6) - A*Ad*eu;G = Ad*e(A*u;G) = Ad*e(A*u;) = A*Ad*e(u;) = A*ad ω = A*cjg/(6)· Since A(0) =e = Ad(A(0)), it follows that Ad(A) = A. ■ Geometric Orientation Let ξ = (Ρ, ω) be a Cartan geometry on a manifold Μ with model (g, ()) and (not necessarily connected) group #. We approach the notion of a geometric orientation for ξ indirectly through the notion of the geometric orientation-preserving subgroup of H. Definition 4.2. Fix a point ρ e Ρ lying over x. An element h G Η is called geometrically orientation preserving with respect to the base point ρ if there is a path A from ρ Ε Ρ to ph Ε Ρ whose development, via ad ω, yields a path A on Gl(g) joining the identity e to Ad(/i). The set of
204 5. Shapes High Fantastical: Cartan Geometries geometrically orientation-preserving elements of Η is denoted by Hor (cf. Propositions 4.4 and 4.5). * We will need the following result. Lemma 4.3. Let p,q G P, let σ be a path joining ρ to q, and let σ: (/, α) —» (G/(g),e) be its development, via ad ω, on Gl(g). We claim that, for any h G H, the development of Rh,o~ on Gl(g), via ad ω, and starting at the identity, is the composite Ad(Ad(/i)_1) (/,0)-^(GZ(fl),e) — (<ЭД,е). P^Ad(/i)-15rAd(/i) Proof. We calculate (Αά(Αά(ΗΓι)οσ)*ωοιω = σ(Αά(Αά(Η)-ι))*ωοι(9) = σ*(Αά(Αά(ΗΓι))ωοι(9) (by 3.1.8) = Αά(Αά(Η-ι))σ*ωοι{9) = Ad(Ad(/rx))a*(ad ω) (by definition of σ) = a*(Ad(Ad(/i"1))ado;) = a*(ad(Ad(/i-1)a;)) (by 3.3.5(H)) = (ϋΗσ)**ά(ω). By the uniqueness part of the fundamental theorem of calculus (3.5.2), Ad(Ad(h)~l) ο σ is the development of Rh,o~. Ш Proposition 4.4. For Ρ connected, the subset Hor С Η of elements preserving the geometric orientation does not depend on the choice of p. Proof. Fix p,q G P. Let h G Hor preserve the geometric orientation with respect to p. We wish to show that it also preserves the geometric orientation with respect to q. Let λ be a path joining ρ to ph and σ be a path joining ρ to q. Let the developments of λ and σ on Gl(g), via ad ω, and starting at the identity, be denoted by λ: (7, α, b) —► (Gl(g),e,Ad(h)) and σ:(/,α, b) —► (G/(g),e,Ad(r1)), respectively. By Lemma 4.3, the development of Rh,o~, via ad cj, and starting at the identity, ends at Ad(Ad(/i)_1)Ad(0 = Ad(h~llh). Using this fact, the following diagram, and Exercise 3.7.5, we see the path σ_1 • λ • (Rh<?) develops, via ad ug, to a path on Gl(g) joining e to Ad{Tl)Ad{h) · Ad{h~llh) = Ad{h).
§4. Development, Geometric Orientation, and Holonomy 205 Ad(/ Ad(/ Gi(a) This shows that h preserves the geometric orientation with respect to q. Ш Proposition 4.5. For Ρ connected, we have He С Hor < Η. Proof. If h lies in the identity component He of Я", then there is a path h(t) on Η with h(0) = e and h(l) = h. This path then yields a path ph(t) on Ρ joining ρ and ph. Since this path lies on the fiber pH, where ω restricts to the Maurer-Cartan form of H, it follows that ph(t) develops to h(t) on Η and hence to Ad(h(t)) on Gl(g). Thus it joins e to Ad(ft), showing that h G Hor. Next we show that Hor is a subgroup of Η by showing that it is closed under multiplication and inverse. If hi,h,2 £ Hor, then there are paths λι and λ2 joining ρ to phi and ph2, respectively, whose developments λι and λ2 on Gl(g), via ad cj, join the identity to Ad(fti) and Ad(/i2), respectively. Moreover, by Lemma 4.3, Rh2^i develops to Ad(Ad(/i2)_1Ai. Then the path λ2 • (i?^2Ai) joins ρ to ph\h2 and develops, via ad cj, to the path λ2 • Ad(/i2)Ad(Ad(/i2)_1)Ai on Gl(g) joining e to Ad(ftife2). Thus, hih,2 G iior· If /г G Яог, then there is a path λ joining ρ to p/i whose development λ on GI(q) via ad ω, joins the identity to Ad(h). Then λ-1 is a path joining
206 5. Shapes High Fantastical: Cartan Geometries q (= ph) to qh~l (= p), and the development of λ-1, via ad ω, starting at the identity on Gl(g) is (cf. Exercise 3.7.5) Ad(/i)_1A_1, which ends at Ad(h)~l. Thus,/ι"1 eHor. To show that Hor С Η is normal, let h G Hor and к G H. Choose a path λ, joining ρ to ph, so that it develops, via ad ω, to a path λ on Gl(g) joining e to Ad(h). By Lemma 4.3, Rk\, which joins pk to phk = pk(k~lhk), develops, via ad ω, to the composite which ends at Ad(Ad(k)-l)Ad(h) = Ad(k~lhk). Thus, k~lhk G Hor. Ш While in general it seems difficult to describe the geometric orientation- preserving subgroup Hor of a Cartan geometry with more precision than that given in Proposition 4.5, there is an important case in which Hor = He. Proposition 4.6. Let (Ρ,ω) be a Cartan geometry on Μ modeled on (g, i}) with group Η. Let G be a Lie group realizing g and containing Η as a subgroup. Assume that (i) GeC\H = He, and (ii) AdgiG —> Gl(g) is infective. Then Иor = He. Proof. Since Hor D He by Proposition 4.5, it suffices to show the reverse inclusion. Let h G Hor and fix ρ G P. Then there is a path λ joining ρ to ph whose development, via ad cj, yields a path λ on Gl(g) joining the identity e to AdQ(h). We can also develop λ via ω, to yield a path λ on G joining the identity e to an element g G G. By Lemma 4.1, Ad6(A) = λ so, in particular, Ad6(#) = AdQ(h). Thus, g = h, since Ad6 is injective. It follows that h G GeC\H — He and hence Hor С Не. Ш Corollary 4.7. Suppose that Η is a Lie group and p:H —* Gl(V) is an injective homomorphism. Regarding V as an abelian Lie algebra, we set g = f) xp V (i.e., [h,v] = p*(h)v for h el), ν G V). Let (Ρ,ω) be a Cartan geometry on Μ modeled on (β, f)) with group Η. Then Hor — He whenever Ρ is connected. Proof. First note that g is the Lie algebra of the Lie group G = Η xpV. Clearly, Η С G. By the proposition, it suffices to verify conditions (i) and (i) Let (h,v) e Η xpV and (k,w) G i) xp V. We have Ad(ft, v)(k, w) = (ad(ft)fc, p{h)(w - p*{k)v)).
§4. Development, Geometric Orientation, and Holonomy 207 Then ( Ad{h)k = k, (Ad(h) = idb, Ad(h,v) = id0 =» < p(h)w = w, for all (k,w) G g =» < ρ(/ι) = idy, Ιρ(λ)ρ*(Φ = ο, U = o, and so Ad0: G —» G/(g) is injective because ρ is injective. (ii) As a topological space, G is just the topological product of Я and V. Since V is contractible, it follows that Ge — HexV and hence GeilH = He. Definition 4.8. Let ξ = (Ρ, ω) be a Cartan geometry on a manifold Μ with model (g, i}) and (not necessarily connected) group Η. ξ is called geometrically oriented if ϋΓΟΓ = H. The principal covering MOT = P/Hor —» Μ (with group H/Hor) is called the geometric orientation cover of Μ. ζ is called geometrically orientable if there is a reduction of the bundle Ρ —» Μ to the subgroup ϋΓΟΓ. A geometric orientation is the choice of such a reduction. ft Exercise 4.9. Show that a Cartan geometry ξ — (Ρ, ω) on Μ is geometrically orientiable if and only if the principal covering space Mor —> Μ is trivial. Show that the geometric orientation cover of Μ is geometrically oriented. □ Proposition 4.10. Locally Klein geometries are geometrically oriented. Proof. Consider a locally Klein geometry Μ = Γ \ G/H. As a Cartan geometry, this has principal bundle Ρ = Γ \ G and Cartan connection ω = ur\G- Fix the point ρ = Ге e Ρ = T\G. Let h G Η be arbitrary. Since G is connected, we may choose a path A: (7,0,1) —» (С,е,/г). Let its projection to Ρ be λ: (7,0,1) —» (P,p,ph). Since π*α;ρ\σ — ^G, it follows that the development of λ on G via λ*ωΓ\ο is the same as the development of λ on G via ωό, which is just λ itself, which ends at h. Thus, the development of λ on GI(q) via ad ωΓ\ο is the development of λ on Gl(g) via ad uq-, which is λ = Ad(A) ending at Ad(/i). Thus, h G Hor for any h G H. It follows that locally Klein geometries are geometrically oriented. ■ Lemma 4.11. The Cartan geometry ξ = {Ρ, ω) on Mor with model (g,f)) and group Hor is orientable with a canonical orientation. Proof. By the very definition of Hor, the geometry is geometrically orientable and is itself a geometric orientation.
208 5. Shapes High Fantastical: Cart an Geometries Development of Paths on Μ Let ξ = (Ρ, ω) be a Cartan geometry on Μ modelled on the Klein geometry (G, H). Here we are ultimately interested in the development of a path on Μ as a path on G/H. However, we begin by considering again the development of paths on P. If ξ is flat, so that the structural equation holds for ω then the discussion of monodromy given in §7 of Chapter 3 applies directly. In the case of a general Cartan geometry the structural equation fails so the development of a path on Ρ depends on more than just the homotopy class of this path. However, there is a remnant of the structural equation still in force. According to Lemma 3.13, on each fiber pH of Ρ the Cartan connection can be identified with the Maurer-Cartan form of H. This means that the structural equation continues to hold in the fiber directions even in the most general Cartan geometries. The effect of this will be to maintain some homotopy invariance in the fiber direction in the development construction. Our study of this phenomenon will be based on the following observation. Lemma 4.12. Let ξ = (Ρ, ω) be a Cartan geometry modeled on the Klein geometry (G,H). Let σ:(7,α,6) -» (P,p,q) h:I->H be piecewise smooth maps, where I = [a,b]. Then (ah)~ = ah, where the development (ah)~ starts at h{a) and the development a starts at e G G. Proof. First note that (ah)~(a) = h(a) = a(a)h(a), so the two curves on G start at the same place. Let us show that the Darboux derivatives of ah and (ah)~ are the same. For ah we can use Exercise 3.4.12 to calculate the Darboux derivative as {ph)*UG = Ad(/i_1)a*u;G + №ω0. Now consider (ah)~. By the definition of development, we have {ah)~*UG — (ah)*uj. We cannot use Exercise 3.4.12 to calculate (ah)*ω, since ah takes values in P, which is not a Lie group. However, a similar calculation is possible. Factoring ah as we find (ah)*u = (μο(α x h)o Α)*ω = ((a x h) οΔ)*μ*ω = ((σ x h) о A)*(Ad(/Tх)7Гр^ + π*ΗωΗ)
§4. Development, Geometric Orientation, and Holonomy 209 ^ = Ad(h~~l)(np ο (σ x ft) ο Α)*ω + π# ο (σ χ ft) ο Α)*ωΗ = Ad(h-l)a*u + h*uH = Aa(h~l)a*uG + h*uG. Thus, (σΙι)*ωο = (σϊι)~*ωο and hence ah = (ah)~. Ш The first consequence of Lemma 4.12 is that we can extend the notion of development of paths on Ρ to that of development of paths on M. Whereas a path on Ρ develops a path on G, a path on Μ develops a path on a model space G/H. This is studied in Proposition 4.13 and Definition 4.15. Proposition 4.13. Let ξ = (Ρ, ω) be a Cartan geometry on the manifold Μ modeled on the Klein geometry (G, 77). Let σ:(7,α,6) —» (Μ,χ,у), where I = [α, b]. Let σ: (7,α, b) —» (Ρ,ρ, q) be any lift, and let σ: (I, a) —» (G, e) 6e zis development on G. T/ien г£§ image σ — πσ: (/,α) —» {G/H,e) in G/H is a curve that is independent of the choice of the lift σ. Proof. If σ: (7,α,6) —» (P,p,q) is a lift of σ developing to σ: {I, a) —» (G,e), then any other lift has the form σ/i for some map h:I —> Η and by Lemma 4.12, σ/ι develops to ah. Since πσ/ι = πσ, the two developments on G/H are the same. ■ Exercise 4.14.* In the context of Proposition 4.13, consider a reductive model of the form (G,#) = (Я xp V,H), where p:H -» G/(F) is a representation and we regard V as a commutative Lie group. Then G/H = V and fl = f) xp V (i.e., [ft,v] = p{h)v for ft G Ρ), ν e V). Now the curve σ: (7, α, 6) —» (Μ,χ, у) may be lifted to a horizontal curve σ: (7, α, 6) —» (Ρ,ρ, g), namely, a curve such that σ*α; takes values in V.19 Show that the development of σ on V is the curve σ: (7, α) —» (V, 0) satisfying σ*ω = da. □ Since any curve on Μ has a lift to P, the following definition makes sense. Definition 4.15. Let ξ = (Ρ, α;) be a Cartan geometry on the manifold Μ modeled on the Klein geometry (G,H). Let σ: (7, α, b) —» {Μ, χ, у), where 7 = [α, 6]. Then the development of σ on G/H is the projection to G/H of the development on G of any lift to Ρ of σ. * For a concrete interpretation of this notion of development, see §3 in Appendix B. Of course, V is its own Lie algebra.
210 5. Shapes High Fantastical: Cartan Geometries This definition has a certain conceptual power. For example, if we have a notion of straight line20 on G/77, we can define the corresponding notion of a geodesic on Μ by calling a path on Μ a geodesic if it develops to a straight line on G/H. See Definitions 6.2.6 and 8.3.4 for examples in Riemannian and projective geometries. In general, however, there is no notion of a straight line in G/H. There is, however, the following more general notion. Definition 4.16. Let ξ = {Ρ,ω) be a Cartan geometry on Μ modeled on (g, rj) with group H. A generalized circle is a curve on Μ that is the projection of an integral curve of an ω constant vector field on P. Φ Exercise 4.17. (i) Show that the generalized circles in a Klein geometry Μ = G/H are the projections to Μ of the left G translates of the one- parameter subgroups of G. (ii) Show that the generalized circles in the Euclidean plane are the circles and the straight lines. Determine the generalized circles in Euclidean 3-space. □ Holonorny A second consequence of Lemma 4.12 is that we obtain an analog of the notion of the monodromy of a loop on Ρ called the holonorny of a loop on M. In general, only the loops on Μ whose classes lie in Im π*: πι(Ρ,ρ) —> πι (Μ, χ) have well-defined holonomies attached to them. (But see §3 of Appendix A.) Definition 4.18. Let ξ = (Ρ, ω) be a Cartan geometry on Μ modeled on the Klein geometry (G, #). Fix a point ρ e Ρ lying over χ e M. Let λ: (7, dl) —► (M,x) (where 7 = [a, b]) represent an element of Im 7Γ*:7Γι(Ρ,ρ) —► πι (Μ, χ). Let λ:(7,<97) —► (Ρ, ρ) be any lift, and let λ: (7, a) —> (G, e) be its development on G. Then λ(α) £ G is called the holonorny of the loop λ with respect to p. The holonorny group of ξ with respect to ρ is the set Ф(р) С G of all such holonomies. φ This definition is justified in the following exercise. Exercise 4.19. (i) Show that the holonorny of a loop with respect to ρ is independent of the choice of lift. 20In a reductive model G/H where Q = \) Θ p, a straight line is (up to left translation) the image of a one-parameter subgroup whose generator lies in p. In the absence of this property, there is no general notion of a straight line in a homogeneous space. The best one can do in general is speak about generalized circles, that is, the images of arbitrary one-parameter subgroups.
§5. Flat Cartan Geometries and Uniformization 211 (ii) Show that if the holonomy of the loop λ with respect to ρ is g, then its holonomy with respect to ph is h~lgh. In particular, Ф(рИ) — Η~ιΦ{ρ)Ιι. (iii) Show that Ф(р) С G is a subgroup. (iv) Show that if p, q £ Ρ lie over x, у £ Μ, respectively, and σ: (/, 0,1) —» (Ρ,ρ, g), then Ф(?/) = #_1Ф(:с)<7, where # is the endpoint of a development (starting at e £ G) of σ. (ν) Let ρ С Im π*: πι (Ρ, ρ) —» πι (Μ, χ) be a subgroup. Show that the holonomies corresponding to the loops in Μ with classes in ρ form a subgroup Φρ(ρ) С Φ (ρ). (vi) Show that for a flat geometry Φ(ρ) is the monodromy group of (Ρ, α;) of Definition 3.7.9. □ Definition 4.20. The subgroup Фо(р) С Φ (ρ) corresponding to the null- homotopic loops on Μ is called the restricted holonomy group. Φ Exercise 4.21. Show that (i) Φο(ρ) is a connected subgroup of G, (ii) Φο(ρ) is a normal subgroup of Φ(ρ), (iii) Φο(ρ) is trivial for a flat geometry, (iv) there is a canonical epimorphism πχ(Μ,χ) —» Φ(ρ)/Φο(ρ). (This could justly be called the monodromy representation of the Cartan geometry.) □ Example 4.22. One interpretation of the meaning of a torsion free geometry is that infinitesimal loops have no translation part to their holonomy. However, this doesn't necessarily hold true for actual loops. The example of a sphere rolling without slipping or twisting on a plane given in Appendix В models the development, and clearly rolling a sphere once along an equator gives a pure translation as its development. ♦ §5. Flat Cartan Geometries and Uniformization In this section we study in detail the question, when is a connected Cartan geometry a locally Klein geometry? Locally, the only obstruction is the obvious one: the curvature. This is our first result. For the corresponding global statement, there are two reasons why mere flatness is not enough. First, any open subset of a Klein geometry is flat but is not geometrically equivalent
212 5. Shapes High Fantastical: Cart an Geometries to a Klein geometry. The missing ingredient here is completeness. Second, by Proposition 4.10, locally Klein geometries are geometrically orientable. For these reasons the argument in the global case is more involved than the local one and uses the "characterization of Lie groups" given in Chapter 3. Theorem 5.1. Let Μ be a flat, effective Cartan geometry modeled on (g, f)) with group Η. Then each point of Μ has a neighborhood U which is canonically isomorphic as a Cartan geometry to an open subset of the model Klein geometry G/H, where G is any Lie group realizing g. Proof. Fix a point χ £ M. Let (£/, Θ) be a Cartan gauge on Μ, with U connected, simply connected, and containing x. Flatness means the structural equation holds, so by the fundamental theorem of calculus, there is a map /: U —» G, unique up to left translation, satisfying /*(cjg) = Θ. Fix и £ U, and set g = f(u), χ — тг(д). Then we have the following commutative diagram. T8(gH) ΊΤ-* f, η ι Tg(G) „ > 9 J*u/ π, *g proj TU(U) Tx(G/H)^->g/b By the regularity condition on Θ, it follows that the composite proj ви = proj cjG/*n = φ9{πί)*η'·Τη(υ) -» g/i) is an isomorphism. Thus, the map ρ — π/: U —» G/H is an immersion, and hence a local diffeomorphism. Shrinking U if necessary, we get an injective immersion l:U —> V С G/H onto an open set V. Clearly, this is a geometrical isomorphism onto its image since σ: V —» G defined by σ(ι(υ)) = f(v) is a section, over V, of n:G —» G/#, and the gauge σ*(ωο) on G/# pulls back under ι to yield 6»G)) = (σι)*(ω0) = f*{u>G) = 0. Ш We also have the following companion result giving uniqueness. Theorem 5.2. Let U С G/H be a connected and simply connected open subset and let (Ρ,ω) be the Cartan geometry induced on U by this inclusion. Let f: U —» G/H be a local geometric isomorphism. Then there is an element g £ G such that f(x) — gx for all χ £ U. Proof. Let (Pf,ujf) be the Cartan geometry on U induced by /. We have the following commutative diagrams.
§5. Flat Cartan Geometries and Uniformization 213 / Ρ с G Pf cz G ι С 1 π ι f π U с GIH U с GIH Since / is a local geometric isomorphism, there is a bundle map k:P —> P/ covering the identity map on U and such that &*ω/ = ω. Then Tu;G = ω = k*ujf = k*f*uG = {fk)*uG so that by Theorem 3.5.2, I and /fc differ by left translation by an element g £ G. Since к covers the identity, it follows that f(x) = gx for all χ £ U. Geometrically Oriented, Complete, Flat Cartan Geometries Now we pass to a study of geometrically oriented, complete, flat Cartan geometries. Every locally Klein geometry Γ \ G/H is not only flat in its canonical Cartan structure but is complete since the cj-constant vector fields on G are just the left-invariant vector fields which are complete. It is also geometrically oriented by Proposition 4.10. Our aim is to show the converse, namely that every geometrically oriented, complete, flat Cartan geometry arises in this way. The following theorem generalizes results found, for example, in [J.H.C. Whitehead, 1932] and [C. Ehresmann, 1936]. Theorem 5.3. Let ξ = (Ρ, ω) be a complete, flat, connected, geometrically oriented Cartan geometry, with model (g, f)) and group H, on a manifold M. Then there is a connected Lie group G with Lie algebra 9 containing Η as a closed subgroup and a subgroup Г С G such that (Г \ G, ujt\g) is a locally Klein geometry on Г \ G/H and such that ξ is geometrically isomorphic to it. In particular, Μ = Γ \ G/H. Proof. Let π: Go —» Ρ denote the universal cover of P, and fix e £ Go- Let ρ = 7r(e) £ P. Theorem 3.8.7 applies to give us the following two pieces of data: (a) there is a unique Lie group structure on Go with Lie algebra g such that e is the identity and π*(ω) = cjg0; (b) the group of covering transformations of π: Go —> Ρ is a subgroup Г0 С Go acting on Go by left translations. Thus, Ρ = Γ0 \ Go- Using this data, we define К = {JheH^ e K~l(ph) | Ad(fc) = Ad(h)}. We are going to show that
214 5. Shapes High Fantastical: Cartan Geometries (i) К is a closed subgroup of Go (see steps 2 and 3 below), (ii) the map ф: К —> Η defined by n(k) = рф(к) is a covering epimor- phism of Lie groups (step 4), (iii) the kernel of φ is a normal subgroup Ζ of Go and Γο Π Κ = Ζ (step β). In the presence of these three facts, we can set G = Go/Z, Γ = Γο/Ζ. Then Η = K/Z С G is a closed subgroup, Ρ = Γ0 \ G0 = Г \ G, and ω — ur\G- Thus, ξ is geometrically isomorphic to (Γ \ G,ur\G) and, by Lemma 4.3.12, Γ acts as a group of covering transformations on the space G/H with quotient M. Now we proceed to verify (i), (ii), and (iii). Step 1. φ is surjective. Proof of step 1: We show that for every h £ Η there is an element к £ n~l(ph) С Go such that Ad(fc) = Ad(ft). It will follow that к £ К and, since π (к) = ph, ф(к) = h. Since the geometry is geometrically oriented, we may choose a path λ:7,0,1 —» P,p,ph such that it develops on G/(g), via ad(cj), to a path λ: 7,0,1 —» G/(g), 7, Ad(h). The lift of λ to Go joins the identity to some element к on Go by a path whose development on G/(g), via ad(cjG), is, since 7T*(u;) = uq-, once again λ. But by Corollary 3.5.3, this development is just the restriction to the path of the adjoint representation of Go- Thus Ad(ife) = Ad(h). Step 2. К is a union of path components of π~ι(ρΗ) and hence is a closed submanifold of Go- Proof of Step 2: Suppose that к and / lie in the same path component of π~ι(ρΗ). It is sufficient to show that к £ К => I £ К. Write 7r(fc) = ph and π (I) = phs, where h,s £ H. Then there is a path σ on π~ι(ρΗ) joining к to / and covering a path σ on Ρ joining ph and phs. Since к £ K, there is a path λ on Go from e to & covering a path λ from ρ to p/i on Ρ such that Ad(fc) = Ad(ft). Then the path λ • σ joins e to f and projects to a path λ^σοηΡ joining ρ to pfts. This path develops (see the figure below) via ad(cj) to a path joining e to Ad(hs) on Gf(g). Then, as in step 1, Ad(Z) = Ad(fts), and so / G K.
§5. Flat Cartan Geometries and Uniformization 215 Finally, we note that since pH С Ρ is a regular submanifold and π: Go —» Ρ is a covering space, it follows that π~ι(ρΗ) is a regular submanifold of Go· Thus K, being a union of some of the components of π~ι(ρΗ), is also a regular submanifold and hence is closed. Step 3. К is a group and К = {g G G | Rg — Rh for some h G #}, where Rg\ Ρ —► Ρ is induced by Rg: Go —► Go- Proo/ o/ 5iep 3: First we show the inclusion c. Let к G K. Since Ρ = Го \ Go, and the left and right actions commute, it follows that Rk'.G —> G induces a map on P, which we denote by Rk'. Ρ —> P. Now π*№) = Я!ь7г*о; = P^Go = Ad(k-l)uGo = Ad(h-l)uGo = Ad(h-l)n*u = 7r*(Ad(/r x)u;) = π*(Ρ^α;). Since π* is injective, it follows that Щи = R*hu, so that Rk and P^ are equal if they agree at a single point. But we also know that RkP = π(Ρ^ε) = тг(&) = ph = RhP, so that P^ = P^ on P. Now we show the inclusion D. Let # satisfy Rg — Rh for some h e H. On the one hand, Ad(^"1)a;Go = Щж*и) = π" Щи = π*ϋΤΗω = π*Αά(Η~ι)ω = Αά(Η~ι)ω0ο and hence Ad(g) = Ad(ft). On the other hand, Rg = Rh=> R9P = RhP => к(ед) = ph => g G n~l(ph). Thus, g e K. Now we show that К is a group. Clearly, e Ε K. If k £ K, then Rk — Rh for some h G #, so P^-i = P^"1 = P^1 = Р/г-ι and hence k~l G /f. Finally, if fcb fc2 £ #\ then Pfci = Rhl, Ы e Η for г = 1, 2. Thus, Pfclfc2 = Rk2Rk! = Pfc2P/u = P/n/i2 and hence кгк2 G if. 5iep 4- For all elements к е К and # G Go, π(<7&) = п(д)ф(к). Also, φ: К —> Η is & covering epimorphism. In particular, the Lie algebra of К is the subalgebra ()C0. Proof of Step 4: We have n(gk) — π(Ρ^) — Rk^(g) = Rh^(g) = K(g)h. Taking g = e, we see by definition that ф(к) = h and we have verified the formula. Now we apply this formula to show that φ is a homomorphism. Let ki,k2 e K. Then 7r(e)0(fcifc2) = тг(^1^2) = 7r(fci)0(fc2) = *"(e)0(fci)0(fc2). Thus, 0(fci*2) = Ф(к1)ф(к2). Since π: Go —> Ρ is a covering map and 0 is the restriction of π to a union of path components over pH (^ #), it follows that 0 is also a covering map.
216 5. Shapes High Fantastical: Cartan Geometries Step 5. Γο Π Κ = {7 G Го | Ad(7) = e}. Proof of Step 5: 7 G Γ0 => n(*ye) = 7r(e) = p. Thus, Г0 С π-1 (ρ) С π~ι(ρΗ). Hence 7^Г0ПЯ'^7^Го and Ad(7) = е. 5iep tf. Let Ζ be the kernel of φ. Then Ζ = Γο Π Κ and is a normal subgroup of Go- Proof of Step 6: Consider the right action of ζ G Ζ on Γ0 \ Go- This action is of course trivial, but it means that for each g G Go there is an element 7 G Го such that 7*7 = gz. This shows that gzg~l G Г0 for all g G Go- Let N be the subgroup of Go generated by the set {gzg~l \ ζ G Z, g G Go} С Tq. Since this generating set is stable under conjugation by elements of G, it follows that TV is a normal subgroup of G that lies in Tq. We have Ζ С N С Г0. Also, for 2 G Ζ, Ad(^-1) = Ad(i/)Ad(^)Ad((/-1) = Ad(g)Ad(g-1) = e. So by step 5, N С Г0 П К". Hence Ζ С TV С Г0 П К. On the other hand, let g G Γ0 Π Κ. Then ^еГ0=} 7г(е) = ж{д) = 7г(е)ф(д). Thus, 0(#) = е and hence # G ker φ = Z.lt follows that ToDK С Ζ and Ζ С Ν с Γο Π К С Ζ. Thus, Ζ = Ν = Γ0 Π Κ is normal in G0. ■ If we have two locally Klein geometries (Tj\G/H), j = 1,2, for which the subgroups Tj, j = 1,2, are conjugate in G, say cTic~l — Γ2, then there is a geometric isomorphism relating these geometries induced by Lc: G —» G. The following result shows that this is the only geometric isomorphism between two such locally Klein geometries. Proposition 5.4 (Geometric rgidity). If (Tj \ G,H), j = 1,2, are two locally Klein geometries that are geometrically isomorphic in their canonical structures as Cartan geometries modeled on the Klein geometry (G,H), then the subgroups Γι and Γ2 are conjugate in G by some element с G G and the geometric isomorphism is induced by the left translation Lc: G —> G. Proof. Let π: G —► G be the universal covering group, and consider the following commutative diagram 1" G=>r, I7* »r2\G «"OD-ficG / l* r.cG |π' E\G /
§6. Cartan Space Forms 217 where / is the Η bundle map satisfying /*^r2\G — ωΓι\σ and / is any lift of / to the universal covers. Since ^r2\G and ^Ti\g both lift to Uq on G, and / is just / locally, it follows that /*ω£ = ω& and hence by Theorem 3.5.2, / is just left multiplication by some element с G G. For all 7i G Γι, we have 7Γ2 ° 7Γ о /(7i) = / о m ο π(7ι) = f ο πι о 7г(е) = π2 ο π о /(e), and so it follows that for all 71 G Γι, there is a 72 G f2 such that /(71) = 72/(e). Recalling that / is left multiplication yields cr/i — 72c or οη\θ~ι G f2. Thus, cTic~l С Г2. The same argument applied to f~l yields the reverse inclusion so that cY\c~l = Г2. Since the kernel of n:G —» G is a central subgroup Ζ and Tj/Z = Tj, it follows that cTic~l = Γ2, where С = 7г(с). ■ Moduli Space of Complete, Flat Geometries Suppose that Г С G is a subgroup. It acts as a group of covering transformations on G/H if and only if the following two conditions hold. (i) (Free action) For every g G Г, д~1Гд Π Η = e. (ii) (Proper action) For every compact set К С G, {7 G Г | 7/f Π /Γ# ^ 0} is finite. In light of Theorems 5.3 and Proposition 5.4, we see that the "moduli space" of geometric isomorphism classes of complete, flat Cartan geometries modeled on a given Klein geometry (G, H) may be identified with the set of G conjugacy classes of subgroups Г С G satisfying (i) and (ii). §6. Cartan Space Forms In this section we introduce and study Cartan geometries of constant curvature. The possibility for the existence of nontrivial (i.e., nonflat) examples of these is governed by the Η module Hom#(A2(g/f)), fy)? which depends only on the model. If this module vanishes, there are no nontrivial examples (cf. Exercise 6.7). A Cartan form is a complete, torsion free, geometrically oriented, constant-curvature geometry, and our main result is that these geometries are all locally Klein geometries of the form Γ \ Gf /H, where the Lie algebra of G' may not be the model algebra. Mutation The appearance of G' in the preceding description is the result of model mutation. Mutation replaces a geometry modeled on (g, f)) with group Η
218 5. Shapes High Fantastical: Cart an Geometries by one modeled on (g', f)) with group H. We formulate the precise relation we require for mutation between models in the following definition which is essentially the same as Definition 4.3.8. Definition 6.1. Let (g, f)) and (g',l)) be two models with group H. A mutation map is an Ad(#) module isomorphism λ: g —» g' satisfying (i) λ | \) = idb, (ii) for all u, ν e g, [X{u),X{v)\ = X([u,v]) mod i). We say the model (g', f)) with group Я is a mutant of the model (g, f)) with group Η. <& We mention that in [K. De Paepe, 1996] it is shown that, for primitive models of higher order, mutations are always trivial in the sense that the map λ is always a Lie algebra isomorphism. Thus, in the primitive case, mutation is entirely a phenomenon of first-order geometries. Example 6.2. Take Η = 50„(R), \) = so„(R). Now define g", g°, g+ as follows: 0 = tigR",46fi These are all mutants of each other, since we may define a — 0+ sending (^ A)»[v A and go^g- sending (0 °)^(° *Q. It is easily verified that these are mutation maps. ♦ Mutation of models gives rise to mutation of geometries. This is described in the following result. Proposition 6.3. Let Μ be a manifold and let ξ — (Ρ, ω) be a Cartan geometry on Μ modeled on (g, f)) with group Η. Let X: g —* g' be a mutation of models, and set u/ = Χω. Then £' = (Ρ, ω') is a Cartan geometry on Μ modeled on (g',f)) with group H. Moreover, (i) the curvatures of these geometries are related by Ω' = λΩΗ-|([α/,α/] — λ[ω,ω]),
§6. Cartan Space Forms 219 (ii) the geometry ξ' is complete if and only if the geometry ξ is complete. Under the additional hypothesis that ξ is torsion free, we have (iii) ξ' is torsion free, (iv) λ induces a bisection leg i) I is a subalgebra И) Ь С I ϊ Γ. , I i) [ is a subalgebra 1 /~VC0|ii)i,c[ /· Proof. Let us first verify that u/ is a Cartan connection. (We refer to Definition 3.1, page 184.) Properties (a) and (b) are clear. (c) (i) Since λ: g —» g' is a linear isomorphism, it follows that u/ — Χω: ΤΡ(Ρ) —» 02 is a linear isomorphism for each ρ Ε P. (с) (ii) Using the fact that λ: g —► g' is an Η module isomorphism, we see that (Rh)*u)' = (Rhy\u = X(Rh)*uj = XAd(h-l)uj = Ad(h-l)u'. (c) (iii) If X e i), let X^ be the corresponding vector field on Ρ so that ω(ΧΪ) = X. Since λ | f) = idb, it follows that X* = A(X)t. Thus, cj'(A(X)t) = ω'{Χ*) = Aij(Xt) = A(X). (i) We calculate the curvature of the mutated geometry: Ω' = βω' + )-{ω',ω'] = ά{Χω) + \[ω',ω'\ = Χ{άω)+1-[ω',ω'\ A (du + ^[ω,ω]\ + \{{ω',ω'} - \[ω,ω\) = ΑΩ+±([ϋ/ν]-λΚϋ>]). (ii) The equivalence of the completeness of the geometries ξι and ξ2 is a consequence of the fact that the ω ι -constant vector fields are the same as the u;2-constant fields. (iii) By the definition of mutation, we have |([u/,u/] — λ [ω, ω]) G ϊ). Thus, Ω e \) & ΧΩ e Ь & Ω' e ί>. (iv) Suppose that i) С l С g for some Lie subalgebra I different from \) and g. Then, by Exercise 3.12(a), ω-1 (I) is an integrable distribution on P. But then {ω')-\Χ{\)) = {Χω)-\Χ(\)) = ω-\\),
220 5. Shapes High Fantastical: Cartan Geometries so that the same integrable distribution may be described as (ω )_1(^(0)· Then () С А([) С g', and by Exercise 3.12(b) we see that λ (I) is a subalgebra. Since λ: g —> g' is an isomorphism, it follows that it induces an injective mapping ilCfl i) I is a subalgebra Ί J , ii)f>ci JC\0 i) I is a subalgebra ii) (, С I Finally, since λ l: g' —» g is also a model mutation, this inclusion must be a bijection. ■ Note that in mutating the model geometry, no information is lost. We may read the mutation data "backwards" and so recover the original geometry. One may say that mutant geometries are the same geometry with different presentations. The interest in mutation lies in the possibility of changing, and perhaps simplifying, the curvature. Mutation for Reductive Models In the case of a reductive model, the notion of mutation takes on a special significance, which can already be seen in Example 6.2 involving reductive models. Recall that a Cartan geometry with model pair (g, f)) is reductive if there is an Η module decomposition g = f) 0 p. Lemma 6.4. Let (g, f)) be a reductive model pair, with group H. Write the Η module decomposition as g = f) Θ p. Then there is, up to isomorphism, a unique mutant (g', f)) with g' = f) Θ ρ' and [ρ', ρ'] = 0. Proof. Existence. Set ρ' = ρ as an \) (and H) module but with multiplication on p' given by [ρ',ρ'] = 0. It is then easy to verify that g' = \) 0 p' is a Lie algebra and that the canonical map g —» g' is a mutation. Uniqueness. This part is easy. ■ Since, in the reductive case, we can always find a mutation such that [p»p] = 0? and since we lose no information by passing to a mutation, it follows that for reductive Cartan geometries we exclude nothing if we assume that [p,p] = 0. This allows us to completely forget about the larger Lie algebra g and retain only the subalgebra f) and the Η module p. If, in addition to being reductive, the geometry is of first order (i.e., Ad: Я —» G/(g/i)) is injective, cf. Definition 3.20), we may use the isomorphism φρ:Τ(Μ) —» g/ί) и ρ to reinterpret the bundle Ρ as a bundle of frames (cf. Exercise 3.21) with group Η С G/n(R), and we can forget about ρ too. This leads to the notion of an Ehresmann connection (see also Appendix A).
§6. Cartan Space Forms 221 Uniformization of Constant-Curvature Geometries Constant-curvature Cartan geometries are generalizations of the classical Riemannian space forms, namely, geometries of constant sectional curvature that are, locally, Euclidean space, the round spheres, and hyperbolic space. The generalization we consider includes, for example, products of the classical space forms. The main point is that every constant-curvature Cartan geometry is locally a mutation of a flat Cartan geometry. In particular, each of the three classical types of Riemannian space forms is locally a mutation of each of the other two. Definition 6.5. Let Μ be a connected manifold and let ξ — (Ρ, ω) be a Cartan geometry on Μ modeled on (g, ()) with group H. Let Ω be the curvature 2-form. We say that ξ has constant curvature if Ωρ(Χρ,Υρ) is independent of ρ Ε Ρ whenever the vector fields X and Υ are cj-constant vector fields. A Cartan space form is a complete, torsion free, geometrically oriented Cartan geometry of constant curvature. * The constant-curvature condition may also be expressed by saying that the curvature function Κ: Ρ —» Hom(A2(g/f)), f)) is constant, since ΓΩρ(ω-1(«),ω-1(«)),| Κ (ρ) = constant <=> < is independent of ρ > [ for all ΐ/, υ G g J S (Ρ,ω) has constant curvature. This may be restated a bit more computationally as follows. Lemma 6.6. Let Μ be a connected manifold and let ξ = (Ρ, ω) be a torsion free Cartan geometry on Μ modeled on (g, fj) with group Η and curvature 2-form Ω. Let {e/} be a basis of\) and let {e;} complete this to a basis of 0. Let θι denote the erfh component of ω with respect to this basis. Then we may write Ω = Σ^/α^/^ί Λ OjCi, where α^ι = —а^ц: Ρ —» R and (Ρ,ω) has constant curvatures the functions a^i are all constant. Proof. It is clear (cf. Corollary 3.10) that (Ρ,ω) has constant curvature <=> Qp(uj~les,u~let) is independent of ρ for all 1 < 5, t < n. But Qp(u~les,u~let) = Σ^Ια^Ιθί A9j(u~les,uj~let)eI — %ijiaiji(6iS6jt - 6it6js)ei = Σ/(αβί/ - atsi)ei = 2EastieIl which finishes the proof. ■
222 5. Shapes High Fantastical: Cartan Geometries The following exercise shows that the existence of nontrivial examples of Cartan space forms modeled on the Klein pair (g, ()) depends on the nontriviality of Hom#(A2(g/i)), pj). Exercise 6.7. Suppose that Hom#(A2(g/i)), pj) is trivial and that К is constant along each fiber of P. Deduce that the geometry is flat (cf. Exercise 3.4.8). □ The fundamental fact about the constant-curvature Cartan geometries is that, at least locally, they are all mutations of flat geometries. The first step toward proving this is to find the appropriate mutation of the Lie algebra 0· Proposition 6.8. Let К G Hom(A2(g), ()) be the {value of the) curvature function of a torsion free, constant-curvature Cartan geometry, so that K(q, i)) = 0. Let g' be g equipped with the multiplication given by [tx, v]f = [u, v] — K(u, v). Then g' is a Lie algebra, and for all u, ν G g and heH, [Ad(h)u,Ad(h)v]' = Ad(h)[u,v]'. Proof. The bilinearity and skew symmetry of the bracket [ , ]' are obvious, but we need to check the Jacobi identity, [[tz, τ^',κ;]' -f [[г/;,?/]',^]' -f [[ν,ΐϋ]',^]' = 0. Since [и,ν]' = [и, ν] — К (и, υ), this reads, in terms of [ , ], as [[u, v] — K(u, v), w] — K([u, v] — K(u, v), w) + (cyclic permutations of u, v, and w) = 0. (6.9) Now the bracket [ , ] already satisfies the Jacobi identity. The terms K(K(u,v),w) are of the form jRT(f),g) and so vanish by Corollary 3.10. Thus, the Jacobi identity for [, ]; is equivalent to the identity — [K(u, v), w] — K{[u, v],w) + (cyclic permutations of u, v, and w) = 0. (6.10) We show this is a consequence of the Bianchi identity άΩ, = [Ω, ω] as follows. Let Xq, X1? X2 be three u;-constant vector fields on P. Then we may explicitly calculate the two sides of the Bianchi identity (ΚΙ(Χο,Χι,Χ2) = [Ω,ω](Χ0,Χι,Χ2). By Exercise 1.5.16, the left-hand side is 2 ^(-1уХ^П(Х0,...,Хи...,Х2)) 2 = 0 0<i<j<2
§6. Cartan Space Forms 223 The first sum vanishes by the constancy of Ω on cj-constant vector fields, and the second sum is simply -Ω([Χο,Χι],Χ2) + η([Χο,Χ2],Χι)-Ω{[Χι,Χ2],Χο). The right-hand side is £ (-ΐΠΩ(Χσ(0),Χσ(1)),ω(Χσ(2))] (2,1) shufffles σ = [Ω(Χ0,Χι),ω(Χ2)} - [Ω(Χ0,Χ2)ΜΧι)] + №Д2)^№)]· Thus the Bianchi identity reduces to Ω([Χ0,Χι],Χ2) + [Ω(Χο,Χι),ω(Χ2)] + (cyclic permutations of 0, 1, and 2) = 0. (6-H) Set ω(Χο) = и, ω (Χ ι) = ν, ω(Χ2) = w. Calculating again, we have ΜΧο,Χι) = Χο(ω(Χι)) - Χι МВД - ω([Χ0,Χι] = 0-0-ω{[Χο,Χι]), and so Κ{η,υ) = Ω{Χ0,Χι) = ΜΧο,Χι) + [ω{Χο)ΜΧι)] = -ω([Χο,Χι]) + {η,ν}. Thus, Ω([Χ0, Χι], Χ2) = Κ(ω([Χο, Χι]), ω{Χ2)) = К([щ ν] -К(щ v),w) = K([u,v],w). This identity allows us to rewrite Eq. (6.11) in terms of K, yielding K([u, v],w) + [K(u, v),w] + (cyclic permutations of u, v, and w) = 0, which is exactly the identity in Eq. (6.10) required of [ , ]' for it to satisfy the Jacobi identity. Finally, we show [Ad(h)u, Ad(h)v]' = Ad(ft)[t/,v]'. Recall Lemma 3.23, which says that K(Ad(h)u,Ad(h)v) = Ad(h)K(u,v) Thus [Ad(h)u,Ad(h)v]' = [Ad{h)u,Ad{h)v] - К(Ad(h)u, Ad(h)v) = Ad(h)[u,v] - Ad(h)K(u,v) = Ad(h)[u,v]'. Ш The following result describes all the Cartan space forms with a connected model group H.
224 5. Shapes High Fantastical: Cartan Geometries Theorem 6.12. Let Μ be a connected manifold and let ξ = (Ρ, α;) be a Cartan space form on Μ modeled on (g, ()) with group H. Suppose one of the following conditions holds.21 (i) Η is connected. (ii) Μ is simply connected. Then ξ is a mutation of a locally Klein geometry Γ \ Gf/H, where G' has Lie algebra g' as described in Proposition 6.8. Proof. The "identity" map l: g —» g' is a linear isomorphism. Since Κ§·> δ) — 0? the restriction of ι to \) is an isomorphism of Lie algebras. This same identity also ensures that [u,v]' — [u,v] whenever и G f). It follows that the identity map l: g —» g' commutes with the adjoint action of the identity component He С Η. By model mutation we may regard Μ as equipped with the geometry of vanishing curvature modeled on (g',f)) with group H. Either of conditions (i) or (ii) implies that the geometry is geometrically oriented. Then the classification of complete, flat geometries, Theorem 5.3, applies to tell us that ξ is locally Klein and of the form Γ \ G'/H. Ш Exercise 6.13. (a) Suppose that (P,u>) is a constant-curvature Cartan geometry, and let ([/, Θ) be a compatible gauge with curvature Θ. Let e;, 1 < г < η, be a fixed basis for g/f) and let ei{x) G TX(U) be the corresponding smooth vector fields on U defined by e* = 9(ei(x)), 1 < i < n. Show that (i) for each 1 < г, j < n, Qx{ei(x),ej{x)) is independent of χ G {/; (ii) for any h e Η and any χ e U, Αά(Η)θχ(θ-ι(ιι),θ-ι(ν)) = 0x(^-1(Ad(/i),^),^-1(Ad(^)). [Hint for (ii): Relate Ωρ(ω~ι(η),ω~ι(ν)) and ^ph(^~h(u),u~^(v))]. (b) Show that when the adjoint representation Ad: Η —» End(g/i)) is faithful, then the converse of (a) holds. That is, in this case, show that if a Cartan geometry (Ρ,ω) is covered by gauges satisfying (i) and (ii), then it must have constant curvature. [Hint: Show first that if (i) and (ii) hold for one gauge and one basis {e;}, they hold for all equivalent gauges and all bases.] □ I do not know whether or not the property of being geometrically orientable is unchanged under mutation. If it is, then Theorem 6.12 is true without the necessity of these conditions.
§7. Symmetric Spaces 225 §7. Symmetric Spaces One of the classical definitions of a Riemannian locally symmetric space is a Riemannian geometry in which the curvature is covariant constant. This definition continues to make sense for any reductive Cartan geometry using the covariant derivative of Definition 3.43. As usual, for reductive geometries we assume that g = \) Θ ρ is a fixed Η module decomposition. Definition 7.1. Let Μ be a reductive Cartan geometry modeled on (g, f)) with group Η. Μ is called a locally symmetric Cartan geometry if the curvature function К satisfies DPK = 0. * The following result shows that the locally symmetric spaces are "geometries extended from flat geometries." Proposition 7.2. Let Μ be a locally symmetric Cartan geometry (Ρ, ω) modeled on (g, f)) with group Η. Then there is a reduction of the bundle Ρ to a connected principal bundle P\ with group H\ С Η such that (Pi, ω\ = ω \ P\) is a constant-curvature Cartan geometry modeled on the Klein pair (f)i 0p,f)i) with group H\. Moreover, if (Ρ,ω) is complete, then so is (Ρι,ωι). Proof. Fix po e P, let P0 = {p e Ρ \ K(p) = K(p0)}, and let H0 = {h G Η | h · K(po) = K(po)}. Then Щ is a closed subgroup of H. Let f)o be its Lie algebra. Since К is constant along horizontal paths in Ρ (i.e., paths tangent to the distribution u;_1(p)) and horizontal paths can join any point to any fiber, the function К takes on the value K(po) on each fiber. Since K(ph) = h~l · K(p), К is equivariant. Thus, by Proposition 4.2.14, Hq —► Po —► Μ is a principal bundle. The fact that К is constant along paths in Ρ tangent to the distribution ω~ι(ρ) implies that, for each ρ e P, we have ωρ(Τρ(Ρο)) D p. Since the fiber of Po is Я0, it follows that ωρ(Τρ(Ρο)) D f)o· The equality dim Po = dim Μ + dim Щ implies ωρ(Τρ(Ρ0)) = ί)ο θ p. Thus, ω0 = ω \ P0 is an (ί)ο θ p)-valued 1-form on Po. Since the form ω:Τρ(Ρ) —► g satisfies conditions (i), (ii), and (iii) of Definition 3.1, it follows that ωο does as well, and so it is a Cartan connection on Po. Let us calculate the curvature of this geometry. It is Ω0 = άω0 + - [ω0, ωο] = ί άω + - [ω, ω] ) \ Ρ0 = Ω | Ρ0. Since the curvature function Ко = К \ Po = constant, the geometry (Po,u;o) has constant curvature. Finally, fix a component Pi of P0 and an element pi G Pi, and set Hi — {h e Η I pih e Pi}. By Exercise 1.3.23, Ηχ has codimension zero in Η and Hi —► Px —» Μ is a principal Hi bundle. Setting ωχ = ωο \ Pi, we see that (Pi,u;i) is a connected, constant-curvature Cartan geometry on Μ modeled on (ί)ιθρ,ί)ι) with group H\.
226 5. Shapes High Fantastical: Cartan Geometries The completeness of Ρ implies the completeness of Pq since the ωο~ constant vector fields on Pq are restrictions of cj-constant vector fields on P. Ш Corollary 7.3. Let ξ = (Ρ, ω) be a complete, reductive, locally symmetric Cartan geometry on Μ modeled on (g, f)) with group H. If Μ is connected and simply connected, then it has the form Gi/Ηχ, where H\ С Но is a closed subgroup with Lie algebra f)i and G\ is a Lie group with Lie algebra f)o Θ Ρ, with bracket given by [u, v\ = [u, v] — Kq(u, v). Proof. The long, exact homotopy sequence . . . -> 7Ti(M) - 7Γο(#ΐ) - ΤΓο(Ρΐ) - ■ · ■ О О shows that Η χ is connected. The corollary follows by applying Theorem 6.12 to the conclusion of Proposition 7.2. ■
6 Riemannian Geometry A Riemannian geometry consists of a smooth manifold Μ together with a Riemannian metric, that is, a smooth function qM'-T(M) —► R, which restricts to a positive, definite, quadratic form on each tangent space.1 Euclidean geometry of dimension n, denoted by En, consists of the pair (Rn, (—,—)), where (—,—) is the usual inner product on Rn. It may be regarded as the simplest example of a Riemannian geometry by defining q:T(M) = RnxRn-^R by q(x,v) = (υ,υ). This example makes it clear how Riemannian geometry is a generalization of Euclidean geometry. A deeper study of Riemannian geometry shows that more structure can be canonically associated to a manifold with a Riemannian metric, including the bundle of orthonormal frames along with its tautological 1-forms and the Levi-Civita connection. This construction is reviewed in §3. The extra canonical structure just mentioned determines a torsion free Cartan geometry on Μ modeled on Euclidean space. Conversely, any Cartan geometry on Μ modeled on Euclidean space determines, up to a constant scalar factor, a Riemannian metric on M. These facts provide a more profound justification for regarding a Riemannian manifold as a generalization of Euclidean space and show the equivalence between Riemann's original idea and Cartan's version of the same thing. 1Riemann actually had something more general in mind. Cf. [S.-S. Chern, 1996].
228 6. Riemannian Geometry In §1 we study, in an ad hoc manner, the representation theory associated to the model Euclidean space. In §2 we define the Cartan geometries modeled on Euclidean space, giving a brief picture of the corresponding special geometries and ending with a discussion of the geodesies. In §3 we show how these geometries "geometrize" Riemannian metrics as described in the preceding paragraph. In §4 we study some of the special geometries, the Riemannian space forms. In §5 we pass to a study of subgeometries, second fundamental form, Ehresmann connection on the normal bundle, and a complete set of invariants for a submanifold. In §6 we apply these notions to introduce isoparametric submanifolds of Riemannian space form. We end that section by proving Cartan's formula relating the principal curvatures of an isoparametric hypersurface in a space form. §1. The Model Euclidean Space There are really two versions of Euclidean geometry, the oriented and the unoriented models. We describe the unoriented version here; the oriented one is obtained from it by replacing On(R) by 50n(R). Let G = Eucn(R) be the group of rigid motions of Euclidean η space. Let Η = On(R) С G denote the subgroup fixing the origin. In more detail, G={(1 ")eMn+1(R|i€On(R)J, Я={(о Ji)eMn+1(R)M€On(R)J, with Lie algebras 0 = eucn(R) and f) = o„(R) given by 0,veRn\, Definition 1.1. The group Eucn(R.) = G is called the Euclidean group in dimension n. The pair (G, Д"), with Η « On(R), described above is called the Euclidean model dimension n. ft The model geometry is reductive since the subalgebra i) С 0 has an Ad(#) invariant complement P = {(°v o)eMn+1(R)|W€R"}, and so 0 = \) 0 ρ is an Η module decomposition, and the adjoint action of Η on ρ « g/ί) is given by ad(^)i; = Av. Let e» e ρ denote the standard 9 = {{°v °А)€Мп+1(В.)\А + А* =
§1. The Model Euclidean Space 229 basis and let e^j G i) denote the unique elements satisfying ad(e^)efc = SjkCi — fiik^j such that ad(eP7) corresponds to ep 0 e* — eq 0 e* under the canonical isomorphism End(g/fj) w (fl/f)) 0 (fl/f))*· Note that {e^ | г < j} is the standard basis for F). The structure of the Η modules rj and Hom(A2(g/i)), fj) determine the possible types of the (torsion free) special Cartan geometries modeled on Euclidean space. The remainder of this section is devoted to the study of these models. Exercise 1.2. Show that i) is a simple Lie algebra for η φ 4, and that for η = 4, \) w 5θ(3) Θ 5θ(3). [ffmi: For η = 4, consider the matrices / 0 1 0 0\ -10 0 0 | 0 0 0 1 \ 0 0-10/ /0010 " 0 0 0-1 -10 0 0 \ 0 1 0 0 0 0 \-l 0 0 Γ 0 1 0 -10 0 о о о, and /0 1 -1 0 0 0 0 0 0 0 0 -1 V ο ο ι ο 0 0 V -1 0 о о \-l 0 0 0, It follows from this exercise that for η φ 4 there are no special types of torsion free Cartan geometry on Μ modeled on Euclidean space arising from the ideals of h. However, there is still the possibility of special types of geometries arising from the decomposition of the Η module Hom(A2(g/h), h). Note that this module vanishes for η = 1. Definition 1.3. The Ricci homomorphism is the composite mapping Ricci:Hom(A2(g/h),f)) A2(0/h)* ® 0 id<g)ad χ2 \2(&/Ы ® End(0/h) -^ (β/ίο* ® Ы/Ы, where φ{ν\ A v% 0 ψ) = v\ 0 (v^ ο φ) — v% 0 (v\ ο φ). i Proposition 1.4. Let η > 2. (i) Ricci is an Η module homomorphism satisfying Ricci(e* Λ e* 0 epq) = bjVe\ 0 e* - bjqe\ 0 e* - <5ipe* 0 e* + ^7e* 0 e; 50 that, in particular, (a) w/ien гфкф j, Ricci(e* Л e£ 0 e^) = e* 0 e* + <5^е£ 0 e£.
230 6. Riemannian Geometry Moreover (b) for η — 2, Ricci is injective. (c) for η > 2, Ricci is surjective. (ii) Let Hom(A2(g/i)), \))в С Hom(A2(g/i)), ()) denote the subspace of elements satisfying the "first Bianchi identity" given by: φ G Hom(A2(g/i)), \))B if and only if Σσ6Λ3[^(^(σ(ι)Λ^σ(2)),^σ(3)] =0, for alluuu2,u3 e g/i) (where As is the alternating group of order 3). Then Hom(A2(g/F)), \))B is an Η submodule o/Hom(A2(g/i)), \)). Moreover (a) let φ = Σα^ρςβ* Л е* 0 epq e Hom(A2(g/i)), f)) where ijpq ^ skew symmetric in the first two and last two indices. Then φ e Hom(A2(g/f)),f))B & aijkq+ajkiq+akijq = 0 for alli,j,k,q (b) for η = 2, Hom(A2(g/i)), Ь)в = Hom(A2(g/i)), f,) (c) for η > 2, Жссг(Нот(А2(д/()), \))в) = S2($/\)Y (the symmetric elements in (g/i))* 0 (fl/f))*)· (Hi) Let η > 2 and set bij = ^к^<п(ег* л e£ ® efcj + ej л e£ ® e^), 1 < г, j < гг. Then Ricci(bij) = (n- 2)(e* 0 e* + e* 0 e*) + 26ijEke*k 0 e£ T/ie elements b^ are linearly independent, and the vector space they span is an Η submodule Rn С Hom(A2(g/i)), \))в- The Ricci homo- morphism induces an isomorphism Rn —► 5'2(g/())* and a canonical Η module decomposition into irreducible pieces Яот(Х2(д/Ъ), t))B ъ Rn®Wn where Wn = ker Ricci \ Rn. (iv) The submodule Horri{)(A2(g/()), ()) С Hom(A2(g/i)), F))b /ias dimension 1 and generator E^e* Λ e£ 0 e^ = ^Σ^η and its image in S2(%/\))* is ^(3n-4)Efce*0e£. (v) There is an Η module decomposition Rn « Rqu 0 Horri{)(A2(g/()), ()) corresponding under the Ricci homomorphism to the decomposition S2(g/i))* » (Ех<*<пе£ 0 ej> Θ {г/; e 52(g/())* | TYace τι; = 0}.
§1. The Model Euclidean Space 231 Proof, (i) We leave it to the reader to check that each of the maps composing Ricci is an Η module homomorphism. Next, note that under Ricci, the element e\ A e*j 0 epq is mapped according to e* Ле·^) epq не*Ле·® ad(ep<7) ■-» e\ Ле·® Et(bqtep - bpieq)e\ н-> Е^е*(^ер - bvteq)c[ 0 c\ - T,te\[bqtev - bvteq)e] 0 e\ = bjve\ 0 e* - bjqe\ 0 e* - ^peJ 0 e* + biqe] 0 e* and in particular, (a) for i φ j = ρ φ q, Ricci{el A e1j 0 ejq) = e* 0 e* + biqc) 0 e*. (b) For η = 2, dim A2(g/i)) = dim i) = 1 so that e\ A e2 0 e2i is clearly a basis for Hom(A2(g/i)), ()) and, by the formula (a), Ricci(el A e2 0 e2i) = el 0 e\ + e* 0 e* φ 0. (c) On the other hand, for η > 2, it is clear that Ricci is surjective from the formula for Ricci(e^ A e£ 0 e^·) given in (a). (ii) The condition that φ G Hom(A2(g/l)), ί))β implies that for all ^1,^2,^3 eg/f), Σσ€Α3Α(1(Λ)^(Α(1(Λ-1)ι/σ(1) Λ Ad(h-l)vff(2)),Ad(h-l)vff(3)] = 0 which is Σσ€Λ3[Αά(Λ)(^(Α(1(Λ-1)ι/σ(2))),ι/σ(3)] = 0 and so ΣσβΑ3[(Αά(Η)φ)(νσ{ι) Ai/ff(2)),^(3)] = 0. Thus Ad(ft)<p G Hom(A2(g/i)), fj)# and so the latter is an Η submodule of Hom(A2(g/i)),i)). (a) The condition that φ G Hom(A2(g/l), F))b may clearly be written as ΖσβΑ3[φ&σ(1) Aei<7(2)),ei<7(3)] =0, for 3l\iui2,i3 which, since φ(βίσ{ι) A et-ff(2)) = 2EP7(ai<7(1)i<7(2)P7)epg, becomes U = ^^a6A3,P7a«cT(i)i<T(2)P7Lepq'e*c7(3)J = ~^^^еА3,рО'г(7(1)га(2)гсг(3)р^р· Since the ep are independent, this is equivalent to (ii)(a). (b) Let us fix г ф к Φ j and note that by (ii)(a) 1 ι -е^Ле^0е^+Ое^Ле^0е^—-е^Ле*0е^ = ejAe^e^ G Hom(A2(g/f)), f>)B. Thus, for η = 2, the basis e^ Λ e£ 0 e2i for Hom(A2(g/i)), ()) lies in Hom(A2(g/f)),i)) which verifies (ii)(b). (c) On the other hand, for η > 2 the formula in (i) verifies (H)(b). (iii) Again we leave it to the reader to verify that Rn is an Η submodule. Using the formula in (i)(a) we calculate
232 6. Riemannian Geometry Ricci(bij) = Efc7£ijjfficd(e* Λ e% 0 ещ + e* Л е\ 0 еы) = EMiJ(e* 0 e* + ej 0 e* + 26це% 0 e£) = Г Σ*^,,(β? 0 ej + e* 0 e*) if i φ j \2Е^(е^егЧе^е^) if г = j = f (n - 2)(e* 0 e* + e* 0 e*) if г ^ j \ 2(n - l)e* 0 e* + 2Efc^eJ 0 e* if i = j f (n - 2)(e* 0 e* + e* 0 e*) if % φ j { 2(n - 2)e* 0 e* + 2Efce£ 0 e£ if i = j = (n- 2)(e* 0 e) + e* 0 e*) + 26ijEke*k 0 e£. It is clear from this formula that the elements Ricci(bij), г <i,j <n are linearly independent and span the symmetric submodule 52(g/l))* С (g/F))*0 (g/ί))*. It follows immediately that the Ricci homomorphism induces an isomorphism Rn —» 52(g/())* and that Hom(A2(g/i)), ί))β « Rn 0 Wn is an isomorphism of if modules. (iv) We refer to Exercise 3.4.8(c) for the proof that Homl)(A2(g/()), fj) has dimension one with the given generator. It is clear that Σ^ε* Λ e\ 0 e^ = ^Σ^ and, using the formula in (Hi), it is easy to determine its image in 52(g/i))* is as stated. (v) The decomposition 52(g/())*^(E1<fc<ne^0e^) Θ {w G S2(g/f))* | w = Tnjdije* 0 e* with Σ^α = 0} is just the decomposition of η χ η symmetric matrices as a sum of multiples of the identity matrix and trace zero matrices. This decomposition is clearly stable under conjugation by any orthogonal matrix. The first summand is irreducible because its dimension is one. For the irreducibil- ity of the second summation, see Exercise 1.6. Since the Ricci homomorphism restricts to an isomorphism Rn и S2(fl/f))* that, by (iv), identifies Homl)(A2(g/()), ()) with (Ei<^<ne^ 0 e£), we may define Ron to be the sub- module of Rn corresponding to the "trace zero" submodule of S2(g/i))*. Then the decomposition Rn « Homl)(A2(g/()), ()) 0 Ron is clear. ■ Definition 1.5. The submodules of Hom(A2(g/i)), ()) described in Proposition 1.4 have the following names: (i) Нот(А2(д/()),())в is the Bianchi submodule. (ii) Rn is the Ricci submodule. (iii) Wn is the Weyl submodule. (iv) Ron is the traceless Ricci submodule. (v) Homl)(A2(g/()),()) is the scalar submodule. cS
§1. The Model Euclidean Space 233 Exercise 1.6. Let 50n(R) act on the space of trace-free symmetric η x η matrices Symn(R)o by conjugation. Show that this representation is irreducible. [Hint: Suppose there is a nonzero subrepresentation V С Symn(R)o. Each nonzero element of V is conjugate to a diagonal matrix of the form diag(a, /?,...), where α φ β and neither is zero. Then this diagonal matrix is in turn conjugate to the same matrix with a and β interchanged. The difference, which also lies in V, is a multiple of a matrix of the form diag(l, —1,0,... ,0). It then easily follows that V = Symn(R)0.] □ It is known that as an On(R) module, Wn is irreducible (cf. [I.M. Singer and J.A. Thorpe, 1969]). Exercise 1.7. Show that, as an 50n(R) module, Wn can decompose into at most two submodules and, if it does so, these submodules are isomorphic and will be interchanged by the action of any orientation-reversing element ofOn(R). □ Note that when η = 4, W4 = W+ ® W~ as an 504(R) module (cf. [I.M. Singer and J.A. Thorpe, 1969]). Wn is irreducible as an 50n(R) module for η φ A. The existence of the decomposition W4 = W+ 0 W~ in dimension 4 is at the root of Donaldson's important work relating differential geometry to the structure of smooth four-dimensional manifolds (cf. [S. Donaldson and P.B. Kronheimer, 1990]). In summary, the decomposition into three irreducible On(R) modules Hom(A2(g/i)), f,)B = Hom^g/i)), t>) Θ Rn Θ Wn provides a corresponding decomposition of L G Hom(A2(g/()),())в as L = s(L) + ro(L) + W(L), where the notation is chosen so that s, ro, and W will correspond to the scalar, traceless Ricci, and Weyl curvatures respectively. We end this section with the following criterion for the equality of elements of Hom(A2(g/i)), ί))β. Proposition 1.8. (i) φ G #ora(A2(p), ())в satisfies {αά(φ(Χ Λ Y))Z, W) = {αά(φ(Ζ Λ W))X, Υ) for all X, Y,Z,W e p. (ii) Let φ\,φ<ι G Яот(А2(р), ί))β. Then ^ ί (ftd(^i(vi Λν2))νι,ν2> = (ad(<p2(vi Λν2))νι,ν2) ψ1 {for allvi,v2 G p. Proof, (i) This is a consequence of the Bianchi identity. Adding the equations
234 6. Riemannian Geometry (*ά(φ{Χ A Y))Z, W) + (ad(^(r Λ Z))X, W) + (ad(p(Z Λ Χ))Υ, W)=0 -((ad(<p{W A X))Y, Z) + (ad(^(X Λ y))W, Z) + (ad(^(F Λ W))X, Z)) = 0 -((ad(p(Z Λ W))X, Y) + (ad(p(W Λ X))Z, Г) + (ad(^(X Λ Z))W, Y)) = 0 (ad(<p(Y Λ Z))W, X) + (ad(<p(Z Λ W)) У, X) + (ad(<p( W Λ y))Z, X) = 0 and using the skew symmetry of {οά(φ(Χ Α Υ))Ζ, W) in Χ, Υ and Z, W yields 2(ad(<p(X Λ Κ))Ζ, W) - 2(*ά(φ(Ζ A W))X, Y) = 0 for all X, Y,Z,W e p. (ii) It suffices to prove this result for the case φ\ (= <p say) and <p2 = 0. Since => is automatic, we prove <=. Suppose {ad(ip(vi Av2))vi,v2) = 0 for all νι,ν2 Ε p. (*) Then (ad((^(vi Λ (v2 + ι;4)))^ι, ^2 + ν a) = 0 for all vb v2, г;4 е p. So by (*), (ad((^(vi Λν2))νι,ν4) + (ad(<p(vi Л^,^) = 0 for all vi,v2,V4 £ Ρ or, by (i), (ad(<p(vi Λ ν2))νι, V4) = 0 f°r all vi, v2, ^4 e p. (**) Thus, (ad(^((vi+V3)Av2))(vi+V3),V4> = 0 for all vi,v2,v3,v4 Ε p. So by (**), (&ά(φ(νι Av2))v3,v4) + (ad(<p(v3 Λ^))^,^) = 0 for all vi,v2,v3,v4 Ε ρ and so, by the Bianchi identity, {a,d(ip(vi A v2))v3,v4) = 0 for all vi,v2,V3, υ4 Ε p. Thus, 99 = 0. ■ §2. Euclidean and Riemannian Geometry Definition 2.1. Let Μ be a smooth manifold. A Euclidean geometry on Μ is a Cartan geometry on Μ modeled on Euclidean space. A Riemannian geometry2 on Μ is a torsion free Euclidean geometry. * We continue the notation at the beginning of §1. Let us assume that (Ρ, ω) is a Euclidean geometry on Μ with model (g, fj) and group H. Since the Lie algebra g decomposes as an Η module as g = ί) Θ p, the Cartan connection ω: Τ (Ρ) —>> g and the curvature Ω: A2(P) —► g decompose correspondingly as ω — ω^ 0 ωρ and Ω = Ω^ 0 Ωρ. These may also be expressed as 2Of course, this is not the usual definition of a Riemannian geometry, which is a manifold together with a Riemnanian metric on it. The equivalence (up to scale) of these definitions will be shown in §3.
§2. Euclidean and Riemannian Geometry 235 (° °\ ν" ^γ^ ί ° o^ Ω = (ωρ Ц) = Σ *« + Σ ЪЧ, or Ω = ( ° ° ) . 4 ' 1<г<п 1<г<.7<п ч ' Definition 2.2. In the case of a Riemannian geometry, the form ω^ is called the Levi-Civita connection. The form ωρ has been called the solder form. * Exercise 2.3. Verify that the Levi-Civita connection on the principal bundle Ρ satisfies the following properties: (i) R*hub = Aa(h-l)ub; (ii) шь(Х*) = X for every lei). □ The properties of ω^ given in this exercise were adopted by Ehresmann as the definition of what is today called an Ehresmann connection on the principal bundle Ρ with any group as fiber. Definition 2.4. Let Η —» Ρ —» Μ be an arbitrary principal bundle over Μ and let i) be the Lie algebra of H. An Ehresmann connection of Ρ is an (^-valued 1-form 7 satisfying the conditions (i) R*hl = Ad(/i-x)7, and (ii) 7(Xt) = χ for every x6(,, The curvature of the Ehresmann connection 7 is the two form d^-\-\ [7,7]. Ehresmann connections are studied in [S. Kobayashi and K. Nomizu, 1963] and enjoy wide usage since in important cases—Riemannian geometry, for example—all of the geometry may be simply expressed in terms of an Ehresmann connection. See Appendix A for a discussion of Ehresmann connections and their relationship with Cartan connections. Special Geometries We are going to study the most important case, the Riemannian geometries3 on M. Thus, we assume that Ωρ = 0 or, equivalently, Ω; = 0, 1 < i < n. 3The more general Euclidean geometries are apparently not very important. The Lorentzian analog of the extreme case of a Euclidean geometry on a manifold Μ with Ω^ = 0 was investigated by Einstein from 1929-1932 in the hope that a
236 6. Riemannian Geometry The Bianchi identity is dVt = [Ω,ω] or 0 0 0 <mb 0 0 О Пь 0 ω» 0 0 ΩΗ Λί 0 This identity decomposes into (first Bianchi identity) 0 = Ω^ Λ ωρ or 0 = V · Ω^· Λ cjj, (second Bianchi identity) аЩ = [Ω^,α;^], or dttij = J2k№ik A ukj - Uik Λ Qkj). The curvature function K:P -+ Hom(A2(p), ()) is defined by Κ(ρ)(υι,υ2) = ^ίρ(ω~ι(νι),ω~ι(ν2)) for vi,v2 Ε p. Writing Ω^ = ΣΜα^^ω^; Λ ω ι (which we may do since Ω is semibasic), the first Bianchi identity becomes 0 = ^jkiaijki^k Α ωι A Uj, or aijki + Q>ikij +aujk — 0 f°r aH h 3·* ^» '· Thus К takes values in Hom(A2(p), ί))β. It follows from the results of § 1 that in a Riemannian geometry the curvature function decomposes as K = s(K)+r0(K) + W(K). Table 2.5 lists the usual terminology. TABLE 2.5 s(K) scalar curvature МЮ traceless Ricci curvature s(K) + r0(K) Ricci curvature W(K) Weyl curvature W(K) = r0(K) = 0 constant curvature if η > 2 (cf. §4) W(K) = 0 conformally flat (cf. Theorem 7.3.9) r0(K) = 0 Einstein manifold Geodesies The notion of a straight line in Euclidean space immediately extends to the notion of a geodesic in a Riemannian manifold as follows. Definition 2.6. A geodesic in a Riemannian manifold Μ is a curve on Μ whose development in Euclidean space is a straight line. & unified field theory might be based on it [E. Cartan and A. Einstein, 1979]. Since the holonomy of such a geometry is purely translational, and the translations form a normal subgroup of all the group of all motions, there is an absolute parallelism on Μ in the sense that there is a canonical trivialization of the tangent bundle of Μ inducing isometries between the tangent spaces at any two points of M.
§3. The Equivalence Problem for Riemannian Metrics 237 Note that we speak here of an unparametrized geodesic. One may of course choose the arclength parameterization for a geodesic, which is also the parameterization by the arclength of the development. Proposition 2.7. Let ([/, 9) be a Riemannian gauge on M. Let I denote the interval (a, b). Then the geodesies c: I —> Μ are the solutions of the ODE. в^ду + Р^с) Θη(ό)· + Ρη(ό) \-ηίληίλ —тттл— = ■ ■ · = —w-пл—> where pi(v) = 2^ θ*Λυ)θΛυ)· 0i (с) МО ^ Proof. We must show that the given differential equation is a necessary and sufficient condition for the regular curve c(t) to develop to a straight line in the model space Rn. Consider the associated function 9(c): I —► g, which develops to give a curve on G c:I,a —► G, e satisfying uc(c(t)) = 9(c(t)). Now the projection of this curve to the model space Rn is c(£)eo, where eo G Rn+1. Then c(t)eo lies in a straight-line segment in Rn if and only if c(t)e0 and c(t)e0 are linearly dependent. Since c(t) is regular, it is a geodesic if and only if there is a function \{t) such that ceo = ^ceo or, equivalently, c~lceo = Ac-1 ceo- But we have c~lc = ug(c) — 9(c). Since (c_1) = —c~lcc~l, we also have 9(c)' = -c-l'cc-l'c+c-le= -9(c)2 + c~lc. With these identities, the equation for geodesies becomes 9(c) eo H- 9(c)2eo — X9(c)eo. If the gauge is written as \ 9i θα the geodesic equation may be expressed as 9i(c) — λ for г — 1,... ,n. Exercise 2.8. Show that the equations for a geodesic parametrized by arclength are 9i(c)' -f Pi(c) = 0, 1 < г < п. □ §3. The Equivalence Problem for Riemannian Metrics We begin with the classical definition of a Riemannian metric.
238 6. Riemannian Geometry Definition 3.1. A Riemannian metric on a manifold Μ is a smooth function qM'-T(M) —► R whose restrictions qx:Tx(M) —► R are all nondegen- erate quadratic forms. φ What is the relation between this notion and that of a Riemannian geometry as given in Definition 2.1? It turns out that these two notions are, up to a constant scale factor, the same. The following proposition shows that a Euclidean geometry on a manifold always determines a Riemannian metric up to scale. The converse will be treated in Theorem 3.4. Taken together, these results "geometrize" or "solve the equivalence problem" for Riemannian metrics in the sense that they show that two manifolds equipped with Riemannian metrics are isometric up to scale if and only if the associated Riemannian geometries are geometrically isomorphic. Proposition 3.2. A Riemannian geometry on Μ determines a Riemannian metric on Μ up to constant scale factor. Proof. As we noted in §1, the adjoint action of Η = SOn(R) on g induces the standard action on g/() « Rn given by ad^)^ = Av. This action preserves the standard quadratic form q on g/ί), and in fact q is, up to scale, the only quadratic form on g/ί) preserved by H. We may use the isomorphisms φρ:Τχ(Μ) —► g/ί) (where ρ G π~ι(χ)) to transport q to a quadratic form qp on TX(M) defined by qp(v) = q^p(y))· Since φν^ — Αά(Η~ι)φρ, it follows that QPh(v) = q(<pPh(v)) = q(Ad(h-l)(pp(v)) = ς(φΡ(υ)) = qp{v). Thus, the quadratic form qp on TX(M) is independent of the choice of ρ G π_1(χ), and we have found a canonical (up-to-scale) Riemannian metric <7m on M. To see that qM is smooth, consider the following diagrams. ωΌ ω π*ρ\ π* Ψ φ Ψ Ψ Ψ £\ J* £\ i* R R FIGURES 3.3(a) and (b) The diagram on the left commutes by the definition of qx and of φρ, and it implies the commutativity of the diagram on the right. However, the upper composite T(P) —► R in the diagram on the right is clearly smooth, and since π* is a submersion, it follows that qM is smooth. ■
§3. The Equivalence Problem for Riemannian Metrics 239 The proof of the converse of Proposition 3.2 depends on the following lemma, which will also be useful in other contexts. Cartan's Lemma 3.4. (a) Let V be an n-dimensional vector space, and let Oi G V*, г\ — 1,..., n, be a basis. Let μι G \2{V*), г = 1,..., η, be arbitrary. Then there exists a unique collection of elements θ^ G V*, i,j — 1,... ,rc, satisfying (i) θίά + θόί = 0, (ii) μϊ + Σάθίά Λ 0, = 0. (b) Suppose that 0^ G Al(U), where U is some open set in Rn? and that μί G A2(U). Then the forms θ^ guaranteed by the lemma are smooth, that is,9ijeAl{U). Proof, (a) Existence. Write μι = J] · kAijk@j A 0^, where we may assume Aijk H- Aikj = 0, so the As are uniquely determined. Set % = - 2^№iik + Aikj ~ Akji)0k- к Then (i) 6ij + Oji = —Y>k(Ajik + Aikj — Akji + A^k + Ajki - Akij)0k — 0, (ii) TtjOij A 9j = —EjkAikj9k A 9j — Ejk(Ajik — Akji)0k A 9j — ~μ% - ^j<k(Ajik - Akji - Akij + Ajki)0k A 0j = -μι- Uniqueness. If there were two sets of 1-forms satisfying (i) and (ii), their differences 7^ would satisfy (i) and (ii) with μί = 0. Writing 7ij = ЕкВцк0к, (i) implies Bijk + Bjik = 0 and (ii) implies 0 = Yt^ijAOj = YtjkBijkOk Λ 0j = Ej<fc(i^jfc - Bikj)9k A 9j, so Bijk = Bikj. But skew symmetry in the first two indices together with symmetry in the last two indices imply that all the Bs vanish. (b) This is automatic from the formula for 0^· appearing in the proof of (a). ■ We apply this lemma to obtain the next result. Theorem 3.5. Let (M,g) be a smooth manifold equipped with a Riemannian metric g. There is exactly one torsion free Cartan geometry on Μ whose associated metric is, up to scale, g. Proof. Let e(x) = (ei(x),e2(x),... ,en(x)) be any choice of orthonormal frame field on an open set U С Μ, and let 0i,..., 0n G Al(U) be the dual 1-forms. Lemma 3.4 assures us that there are unique forms (θ^) on U such that d(0i) + (9ij) A (0;) = 0. We set
240 6. Riemannian Geometry 0 Лелчад. Thus, θ is a g-valued 1-form on [/, and since e(x) is a frame, 0 has the property that the composite tx(U) -^ 9 ^ 9/0 is an isomorphism. Thus, θ is a Cartan gauge. It is clear that the Riemannian metric induced on U from the standard Euclidean metric on g/fj is the original metric g \ U. The whole of Μ is covered by such Cartan gauges, and it remains only to show that a different choice of orthonormal frame fields yields an equivalent gauge. Suppose that f(x) — (/i(x), /20*0, · · · »/n(^)) is another orthonormal frame field with dual 1-forms ψχ,... ,ψη G Al(U). Now the two frame fields must be related by an orthogonal transformation h:U —► O(rc), so that fi(x) = Ejhijej, and ^i(x) = EjhijOj. But under the change of gauge h:U —> 0(n), we have fl^Adih-^fl + h'W 1 0V° "V1 'Λ + ί1 ° 0 h~l J \0i *у vo лу \o • 0 0\_/0 0 h-l(9i) *) ~ \фг • and since this new gauge must also be torsion free, by the Cartan lemma it must be the unique torsion free gauge corresponding to the coframe ^ь · · ·»Ψη· Thus, θ and ψ are gauge equivalent. In this way we have constructed a Riemannian geometry on Μ whose associated Riemannian metric is, up to scale, g. Moreover, the construction shows that any Riemannian gauge whose associated metric is g will be gauge equivalent to the ones constructed above. ■ Exercise 3.6. Show that the bundle Ρ associated to the Cartan geometry described in the proof of Theorem 3.5 may be identified, with the aid of the isomorphisms φρ:Τχ(Μ) —► ρ, with the On(R) bundle of orthonormal frames on M. If Ρ is regarded as the bundle of orthonormal frames on M, show that the solder form ωρ is determined by Ρ alone. (For this reason, the analog of ωρ on the bundle of orthonormal frames is sometimes referred to as the tautological form.) Cf. also Exercise 5.3.21 and §2 of Appendix A. □ Exercise 3.7. Let (U,x) be a coordinate neighborhood in a Riemannian geometry, and suppose that in this coordinate system g— 2_. gijdxi ®dxj. l<i,j<n
§4. Riemannian Space Forms 241 Choosing an orthonormal frame field (ei, e2,..., en) on U determines the dual coframe field (0χ, #2, · · ·, θη). Let e = 0p + 0beAl(U,t), where θ' = {1 ο) and *» = (ο *°)' be the corresponding unique torsion free gauge on U guaranteed by Theorem 3.5. Recall that the covariant derivative DyX is determined by ΘΡ(ΌΧΥ) = ΧΧ(ΘΡ(Υ)) + [вь(Х),вр(У)] (cf. Proposition 5.3.49). (a) Show that Dxei = Y,l<k<nOki{X)ek. (b) Show that X{g{Y, Z)) = g(DxY, Z)+g(Y, DXZ). [Hint: Use the fact that the linear map 6P:TU(U) —^ ρ is an isometry to write g(DxY,Z) = Χ(ΘΡ(Υ)) ■ ΘΡ(Ζ) + [вь(Х),вр(¥)) ■ θν(Ζ), and then use the skew symmetry of вц(Х).] (c) Writing Ddidj = Ei<fc<n7^fc, where ds = d/dxs, 1 < s < n, calculate дкд(дг, dj) in terms of the 7s using the formula in (b). Show that l<k<n where g^^ = digjk, and so forth. □ §4. Riemannian Space Forms In this section we study the special geometries for which W(K) = ro(K) = 0. Our first aim is to justify the terminology in Table 2.5, referring to these geometries as having constant sectional curvature. In these geometries the curvature function takes values in Hom^A^g/i)), fj). By Exercise 3.4.6(c), this is a one-dimensional vector space, so the curvature must have the form Thus, the curvature function is K(p) = ^ c(p)e* Л e* 0 e^, where с = c(p) is some function on P. The case of η = dim g/F) = 2 is somewhat special since Hom(A2(g/i)),i)) = Homl)(A2(0/()), fj), so it is no condition at all to
242 6. Riemannian Geometry say the curvature takes values in this submodule. However, if we avoid this case, we can prove that the curvature function is constant as follows. Lemma 4.1. Suppose the curvature function for a Riemannian geometry takes values in Hom0(A2(g/f)), ()). Ifn>3, then the curvature function is constant. Proof. This is an application of the Bianchi identity dft — [Ω,ω], which in this case reads Hi'0 ° \0 ο(θ{Αθό) 0 0 \ / 0 0 О с&Лва))'\М (%) 0 0 0 c^k(OiAOkA0kj-0ikA0kA0j) Since the geometry is torsion free, d9i — — Y^9k A 9ki, so the Bianchi identity reads, for each z,j, d[cQi A 0j) = c(-0i Α άθό + dOi Α θά) = βά(θι Α θά). Hence (dc) Α θι A Oj = 0 for all г, j and so, if η > 3, then dc = 0 and с is constant. ■ A Riemannian space form is a complete Riemannian geometry with constant sectional curvature. (We take here the definition of completeness described in Definition 5.3.1. This implies the completeness of geodesies, which is the classical notion of completeness.) By Proposition 4.5, a Riemannian space form is a Cartan space form whose model is Euclidean space (cf. §4 of Chapter 5). Of course this case is simpler than the general one in that the coefficients of the curvature function are not only constant (cf. Lemma 5.6.7), they are all equal. For η > 2 this follows from Lemma 4.1, while for η = 2 there is only one coefficient. We are going to describe the universal cover of a Riemannian space form. Proposition 4.2. Let Mn, η > 2, be a Riemannian space form of curvature c. Let Μ be the universal cover of M. (i) lfc = 0, M = Eucn(R)/On(R) (Euclidean n-space) (ii) Ifc>0,M = On+i(R)/On(R) (the n-sphere) (iii) lfc<0, M = Oi>n(R)/On(R) (hyperbolic n-space). Proof. Since the property of having a constant curvature function is local, the universal cover Μ also has constant curvature. Moreover the property of completeness for Μ passes to completeness for the universal cover. Thus Μ is also a Riemannian space form. If с — 0 then Μ is flat and hence it
§4. Riemannian Space Forms 243 is the model space Eucn(R)/On(R). Suppose that с ф 0. Then we have model mutations eucn(R) -+ on+i(R) if с > 0 and eucn(R) -+ Oi>n(R) if с < 0 /0 0\ /0 -cvl\ /0 0\ /0 ctA \v A J ^ \cv A J \v A J ^ \cv A J By Proposition 5.6.4(i) and (ii) both of these mutated geometries are flat and complete. Since they are also simply connected they are geometrically orientable (Exercise 5.4.9) and so, by Theorem 5.5.4, they are as described in (ii) and (iii) above. ■ Sectional Curvature In this subsection we relate the previous results of this section to the classical notion of a manifold with constant sectional curvature. Exercise 4.3. Let Μ be a Riemannian manifold and let V С ТХ(М) be a 2-plane. Choose (i) an orthonormal basis e, / Ε V, (ii) a point ρ Ε Ρ lying over x, (iii) lifts e, / G TP(P) lying over e and /. Show that the number R(e,f) — (а,а(К(р)(ир(ё),ир(/)))ир(ё),ир(/)) depends only on the (unoriented) 2-plane V. In particular, it is independent of the choices in (i), (ii), and (iii). □ Let Gr2(T(M)) denote the Grassman bundle of oriented 2-planes in the tangent bundle of M. Exercise 4.3 justifies the following definition. Definition 4.4. Sectional curvature is the function R: Gr2(T(M)) —» R defined in Exercise 4.3. A Riemannian geometry is said to have constant sectional curvature с at χ G Μ if R(V) — с for every 2-plane V С ТХ{М). А Riemannian geometry has constant sectional curvature с if R is the constant function with value с on Gr2(T(M)). Φ Proposition 4.5. The following are equivalent. (i) A Riemannian geometry has constant sectional curvature с at χ G M. (ii) K(p) — cY^i<Cj e* Λ e* 0 e^· (c G R) on the fiber over x. Proof. Let V = cY/i<je*Ae*®eij = ^с^е*Ле*®е^· G Hom^A2^/!)), \))b- Then for orthonormal vectors υ\ — Σ^α^, ν^ — Σι biei, we have
244 6. Riemannian Geometry (ad(<p(vi Av2))v2,vi) = *}Г akbiasbt(ad((p(ek Л е{))еи es) klst = 2° Σ akbiasbt{8Ld(e* A e*(ek A e{) ® еы)еи es) i jklst = 9C Σ a^iash(SikSji - 6u6jk)(ad(eki)et,e3) г j klst = 2C Σ akbiasht(6ikuji - Sii6jk){6itek - 6ktei,es) г j klst = oc Σ a^iasbtSik6ji6it6ks - -c 22 akbiasbt6ikbji6ktks ijklst ikjlst ~2C z2 akbiasbtbubjkbitbks + -c ^J ^kbiasbt8ii8jk8kt8is ijklst ijklst = -c2_^a>ibjaihj — -cj^aibjdjbi — -cj^ajbidjbi + -c^^djbiOibj ij ij ij ij ~ C /_^ °"i bj "~ c / J O'ibiCLjbj ij ij = c Σ a2i Σ tf -c Σαί bi Σ аэ ьэ i j i j — C. Thus (i) is equivalent to the statement: (аа(К{р)(шр(ё),шр(/)))шр(ё),шр(/)) = с for any orthonormal vectors e, / G u;"1^). This, in turn, is equivalent to the statement: (a,d(K(p)(vi Λ ^2)^1,^2) = с for every pair of orthonormal vectors v\,V2 G p. But by Proposition 1.8(ii), ad(ii(p)) is determined by this formula, so we must have ad(ii(p)) = αφ. ■ §5. Subgeometry of a Riemannian Geometry- Suppose we are given a Riemannian manifold ATn+r and an immersion /: Mn —» 7Vn+r with normal bundle ν (cf. Exercise 5.7). Our main aim in this section is to describe (Ρ/,α;/), the "locally ambient geometry of TV along /" (Definition 5.5). This is a geometric entity that contains all the geometry of the map / and in which the role of the ambient manifold TV is greatly reduced. It is an example of "a locally ambient Riemannian geometry (Ρ, ω) of codimension r on M" described in Definition 5.2. In
§5. Subgeometry of a Riemannian Geometry 245 Proposition 5.8 we show that when N is the Euclidean space Rn+r the map / is determined up to an isometry of Rn+r by the locally ambient geometry of TV along /. In Proposition 5.9, 5.12, and 5.14 we show how to fracture a locally ambient Riemannian geometry on Μ into three pieces consisting of (i) the induced Riemannian geometry on Μ (the tangential part of (ВД), (ii) an Ehresmann connection 7 on the normal bundle ν (the normed part of(i»), (iii) a ^-valued form В (the second fundamental form) on T(M) (the glueing data). The remainder of the section studies the principal curvatures associated to the second fundamental form when the normal bundle is relatively flat (Definition 5.23). Model Algebra We begin with some notation describing the algebra for the model consisting of Euclidean (n + r)-dimensional space and the standard n-dimensional subspace. The corresponding Lie algebras are b = where the diagonal blocks are 1 χ 1, η χ η, and r x r. Of course, we must bear in mind that not all possible entries are allowed. Any matrix in g must have the form '0 0 0 A -Bl В D where A and D are skew-symmetric matrices. As in §1 we have the On+r(R) module decomposition g = \) Θ p. We also note the On(R) χ Or(R) decompositions α = b Θ t and ρ = 10 u, where u = (In particular, we may identify t with α/b and u with g/(fj + a).)
246 6. Riemannian Geometry Geometry of N Localized Along f Recall that we are studying an immersion f:M—>N into a Riemannian manifold. The description of the geometry of TV localized along / is based on the tangent reduction PT along /. This is a reduction of the pullback bundle f*(P) given by Pr = {(*,!>) 6 /* (P) | M/*(T*(M)) = t}. PT sits in the following commutative diagram Ρτ с fiP) f ->P id ' / Μ = Μ > Ν It is a principal bundle over Μ with group HT = 0(ή) χ 0(r) whose Lie algebra is f /0 0 0' Ьт = {[о * о (До о *. If we set ωΤ = /*(ω) | PT, then since t = φρ(ΜΤ(Μ))) = ^(Л(тг,(Тр(Рг)))) = wp(£(T(PT)))mod f) = /*(u;)(rp(PT))mod f), it follows that ωτ mod t) takes values in t and hence has the form '0 0 0 ωτ θ a -pt 0 β 7 Moreover, the restriction of the curvature form to Pr has the shape Ω,τ=άωτ -f ωτ Α ωτ /0 0 0 = \ άθ + αΛθ άα + αΛα-βιΛβ • \ 0 άβ + βΛα + ^Λβ άΊ + ΊΛΊ-βΛβ*/ (5.1) This leads us to make the following definition, which we will relate to (Ρτ,ωτ) in Lemma 5.4. Definition 5.2. Let Μ be a smooth manifold of dimension n. A locally ambient Riemannian geometry on Μ of codimension r is a pair (Ρ, ω) where
§5. Subgeometry of a Riemannian Geometry 247 (i) Ρ is a principal 0(n) χ 0(r) bundle over M, and (ii) ω is an (ί) Θ t)-valued form on Ρ satisfying the following conditions: (a) ω:Τ(Ρτ) —» ί) 0 t is injective at each point, and its composite with the canonical projection i) 01 —» i)r 01 is surjective at each point; (b) R*hu = Aa(h~l)u for all h e #r; (c) ωτ(Χ*) = X for each X e i)r. (Ρ, ω) is called torsion free if, in the notation of Eq. (5.1), άθ -f α Α θ — 0. Two locally ambient geometries (Ρχ,ωχ) and (Рг,^) on Μ are said to be equivalent if there is an Or(R) χ Or(R) bundle map 6: Pi —> P2 such that 6*(ω2)=ωι. * Definition 5.3. The normal bundle associated to the locally ambient Riemannian geometry (Ρ, ω) is the r-dimensional vector bundle ν = Ρ Xon(R)xor(R) Rr where the action of On(R) χ Or(R) on Rr is the standard second factor action. It has a canonical (up to constant scale) metric qv\ ν —» R on it. * The point of Definition 5.2 is that it describes the data associated to the immersion /: Μ —» TV we have been discussing. Lemma 5.4. (Ρτ,ωτ) is a locally ambient geometry on Μ of codimension r. Proof. We verify the conditions of Definition 5.2. (i) is obvious, as is the fact that ωτ is an (i) 0 t)-valued form on Pr. Condition (ii) (b) is an automatic consequence of the formula on Ρ that Щи — Ad(h~l)cj for all h G H. Condition (ii) (c) comes from the fact that the inclusion Pr С /*(Р) is equivariant with respect to the two actions of HT; hence, for lGf)r, the vector field X^ on /*(P) restricts to the vector field X^ on Pr. The injec- tivity of condition (ii) (a) comes from the facts that ωτ = /*(ω) | Pr, / is an immersion, and ω is injective. The rest of (ii) (a) comes from combining (ii) (c) with the fact, shown just before Definition 5.2, that ωτ(Τρ(Ρτ)) — t mod i) for all ρ G Pr. ■ We signal the importance of this example of a locally ambient geometry by giving the following definition. Definition 5.5. Let TV be a Riemannian manifold and let /: Μ —» TV be an immersion. The locally ambient geometry of N along f (or along Μ if / is an inclusion) is the pair (Pr,u;r). ft
248 6. Riemannian Geometry Exercise 5.6. Let φ: Ν —> N is an isometry of the Riemannian manifold TV. If /: Μ —» TV is an immersion, show that the locally ambient geometry of TV along / is equivalent to the locally ambient geometry of TV along φ/. □ Exercise 5.7. Show that the normal bundle of an immersion /: Μ —» TV given by ν = {(x,v) e MxT(N) \ ν e Μ, ν e /^(T^M))-1}, is canonically isomorphic to the normal bundle, in the sense of Definition 5.3, of the locally ambient geometry (Ρτ,ωτ) associated to / as in Lemma 5.4. □ Now we can justify the terminology "ambient geometry localized along /" by showing that, at least in the case of an ambient Euclidean space, it does indeed encapsulate all the geometry of /. Proposition 5.8. (i) Let /: Mn —» Rn+r (Euclidean space) be an immersion. If (Ρτ,ωτ) is the ambient geometry localized along f, then cLjt + ωτ Λ ωτ = 0 and the monodromy representation Φωτ is trivial. (ii) Conversely, let Mn be a connected η-dimensional manifold and let (Ρ, ω) be a locally ambient geometry on Μ of codimension r satisfying du-\- ωΛω = 0 with trivial monodromy. Then there is an immersion /: Mn —» Rn+r? determined up to left multiplication by an isometry o/Rn+r? such that the Euclidean geometry of Rn+r localized along f is equivalent to the geometry (Ρ, ω). Proof. This is just an application of the fundamental theorem of calculus (cf. Theorem 3.7.14). ■ Fracturing the Local Geometry Now let (Ρ, ω) be a locally ambient geometry on Μ of codimension r. The next step in our investigation is to decompose (Ρ, ω) into the three pieces mentioned in the introduction to this section. First we show that the bundle Ρ itself decomposes as the fiber product of principal bundles determined by the tangent bundle of Μ and the associated normal vector bundle v. Proposition 5.9. Let Ptan = P/Or(R) and Pnor = P/On(R). Then (0 Ptan is a principal On(R) bundle and Pnor is a principal Or(R) bundle, (ii) Ρ = Ptan хм Pnor {fiber product), (iii) Ptan and Pnor are principal bundles associated to the tangent of Μ and the normal bundle ν over Μ arising from (Ρ,ω). Moreover, if (Ρ, ω) is the locally ambient geometry of an immersion f, then ν is the normal bundle of f.
§5. Subgeometry of a Riemannian Geometry 249 Proof, (i) Since the actions of On(R) and of Or(R) on Ρ commute, we see that On(H) acts smoothly on Ptan. Since the action of On(H) χ Or(R) is proper on Ρ and transitive on the fibers of Ρ —» Μ, it follows that the action of Οη(ΈΙ) is proper on Ptan and transitive on the fibers of Ptan —» M. Hence Ptan is a right principal On(H) bundle over M. The case of Pnor —» Μ is similar. (ii) The two smooth bundle maps (defined by projection) Or(R) —» Ρ —» Ptan and On(R) —» Ρ —» Pnor induce a smooth map into the fiber product Ρ —» Ptan xm ^Рпог· This map covers the identity on Μ and is obviously an On(R) χ Or(R) bundle map. (iii) Since the standard action of On(R) on Rn is faithful, the principal bundle Ptan —> Μ is determined by its associated vector bundle. The case of -Pnor ~~* Μ is similar. As for the identification of the associated vector bundles, we deal with the case of the normal bundle only and leave the similar case of the tangent bundle to the reader. Define Ρ χ u —» ν by sending (p,v) —» ψρ1(ν). Note that since φρ is an isometry and for ρ G Ρ, φρ(Τη^(Μ)) = t, it follows that ^р(^тг(р)) = u, and hence φρι(ν) — &V(p)· Since every /г G On(R) acts trivially on u we have φ~^(ν) = φ~ι(Αά(Η)ν) = ψρ1(ν)- Thus, the map induces a smooth bundle map Pnor x u = P/On(R) χ u —» г^. But we also have, for every fc G Or(R), φ~£(ν) = (p~l(Ad(k)v). It follows that (p, v) and (pk, Ad(fc_1)i;) have the same image in v. Thus, we get a further induced map Pnor xo(r) u —» г^. This map is vector bundle map covering the identity on M, and the fibers have the same dimensions. Thus, it is a bundle equivalence. ■ Exercise 5.10. Let (Ρ,ω) be a locally ambient geometry on M. Show that Ρ XAd t may be canonically identified with the tangent bundle of M. □ Writing /0 0 0 \ ω=[θ а -/?Ч, (5.11) \0 β 7 / the next step is to see that the parts α and 7 of the form ω determine, and are determined by, certain induced forms (to which we give the same names) on Ptan and Pnor. Proposition 5.12. Let (Ρ, ω) be α locally ambient geometry on Μ of codi- mension r. (i) The forms a and θ (respectively, 7) appearing in the block decomposition of ω are the pullback under the canonical projection Ρ —> Ptan (respectively, Pnor) of unique forms, which we denoted again by a and θ (respectively, 7) on Ptan (respectively, Pnor)-
250 6. Riemannian Geometry (ii) Let 7 be the form on Pnor guaranteed by (i). Then 7 is an Ehresmann connection on Pnor. (iii Let ωΜ=(0θ α) be the form on Ptan guaranteed by (i). Then {Ptan,им) is a Euclidean geometry on M. Moreover, if(P,u) arises from an immersion, then the Riemannian metric on Μ associated to (Ptan, им) is the metric induced on Μ from N. Finally, (Ptan, им) is torsion free if and only if (Ρ, u) is torsion free. Proof, (i) We deal with 7 only, the cases of a and θ being similar. The transformation law in Definition 5.2(b) implies that for h G On(R), Щ^у = 7; moreover, 5.2(c) says that u(X^) = X for X G i)r, so in particular a(X^) = 0 for X G a. Thus, 7 is basic for the principal bundle Ρ —> P/On(R) = Pnor, and since π*: Al(P/On(R)) -» Al(P) is injective, there is a unique form on Pnor pulling back to 7. (ii) The transformation law in 5.2(b) implies that for h G Or(R), jR^7 = Ad(/i-1)7; moreover, 5.2(c) says that ω(Χ^) = X for X e i)r, so in particular j(X^) — X for Χ Ε u. Since the projection Ρ —» P/On(R) is Or(R) equivariant and π*: Al(P/On(R),u) —» ^(Ρ,ιι) is injective, it follows that the formulas #£7 = Κά(Η~ι)η for /г G Or(R) and 7(Xf) = X for X G o(r) also hold for the version of 7 on Pnor. Thus, 7 satisfies the conditions for an Ehresmann connection (cf. Definition 2.4) on ΡηθΓ· (iii) To see that им is a trivialization of the tangent bundle of Ptan, since dim Ptan = dim a, it suffices to show that ujM'T(Ptan) —» α is surjective, or, equivalently, that the form п*(им)-Т(Р) —» α is surjective. But since tt*(^m) is just the canonical projection of ω on a, by condition 5.2(a) it is surjective. The proof that им satisfies the conditions (a) R*hUM = Ad(h~l)uM for h G On(R) and (b) и;м{Х*) = ХЬтХеЬ is similar to the proof of (ii). Thus, им is a Euclidean geometry on M. Suppose that (Ρ, ω) arises from an immersion /: Μ —» TV. Let (Рдг,о;дг) denote the Riemannian geometry on TV and let φ^Ν:Τυ(Ν) —» g/ί) be the usual isomorphism, where рдг G Рлг lies over у Е N. Since it is clear that φρ{ν) = φΦ о /*p(v), for any ν G TX(M) and for any ρ G Ptan covering x, the induced Riemannian metric qM'-T(M) —» i? arising from the immersion is given by
§5. Subgeometry of a Riemannian Geometry 251 9m(v) = 9n(/*(v)) where φρ refers to the geometry (Рг&п^м)· The last statement of (iii) is obvious. ■ Exercise 5.13. Suppose that G —> Ρ —» Μ is any principal bundle and # С G is a closed normal subgroup. Then G acts on Ρ and P/# so that the principal bundle map π: Ρ —» Ρ/Я is G equivariant. Let X G 0. Show that the two versions of X\ one on Ρ and one on P/H, correspond under π*. (This is a point we glossed over in the proof of Proposition 5.12(i) and (u).) □ At this stage we have two of the three pieces of the localized geometry (Ρ, ω), namely the Riemannian geometry on Μ (corresponding to a) and the Ehresmann connection on the normal bundle (corresponding to 7). The remaining piece, which corresponds to /?, is the second fundamental form. The Second Fundamental Form Proposition 5.14. Let (Ρ, ω) be a locally ambient geometry on Μ of codi- mension r and let ν be the normal bundle associated to it. The form β on Ρ given in Eq. (5.11) determines, and is determined by, the bilinear symmetric form В G Hom(T(M) θΤ(Μ),ν), which is given, independently of the choice of ρ lying over χ, by <pp(Bx(u,v)) = β(ω~ι(φρ(η)))φρ(ν), where u, ν G TX(M). Moreover, there are symmetric matrix-valued functions h(i): PT —» Mn(R) (= End{i)) defined by the equation βι = θιΗ{ι), where βι is the ith row of β. These functions are related to В by the formula φρ(ηΥϊι(1)φρ(ν)\ φΡ{μΥΗ{τ)φρ{ν)) Proof. First we show that the expression β(ω~ι(ιυι))νυ2 is symmetric in w\,W2 G t. The structural equation for ωτ in the (3,1) block is β Α θ = 0. This implies that the zth row of β may be expressed as βι — #έ/ι(ζ), where h(i) is an r χ r matrix-valued function on Pr. (We may also regard h(i) as taking values in End(t); cf. Lemma 5.21 ahead.) In fact, the condition βι Λ θ = 0 for all i implies that h(i) is a symmetric matrix. Thus, the expression φρ{Β{η,ν)) =
252 6. Riemannian Geometry /в^ш-^ымщ p(upl(wl))w2 = W2 y&iw^iwxmr)/ /w\h(l)w2\ \w\h(r)w2) (5.15) is symmetric in w\ and w2. Now consider the function b e C{Pr, (52(t*) ® u, p)) = 52(Ad*) ® Ad, where p, given by b(p)(wi,w2) = P(u~l(wi))w2, where Wi G t for г = 1,2. Let us show that the function b has the transformation properties indicated. For h = diag(^,D) e On(R) χ Or(R), we have b(ph)(wuw2) = p(u~f^(wl))w2 = p(Rh*u-l(Ad(h)wl))w2 = (RlP)(u-l(Ad(h)wl))w2 = D-lp(u-l(Ad(h)wl))Aw2 = Ad(h-l)b(p)(Ad(h)wuAd(h)w2) = (p(h-l)b(p))(wuw2). Clearly, β determines and is determined by b. Note that b(p)((pp(vi),(pp(v2)) e u. Since φρ~ι(\ή — vx — the normal space at χ G M, it follows that the formula defining Bx(vi,v2) makes sense; we must also show that it is independent of the choice of ρ lying over x. But this follows immediately from the transformation laws for b and for φ. Finally, we note that since φρ is an isomorphism for each ρ, Β and b determine each other. As for the last formula, we have <pp(B(u,v)) = b((pp(u)^p(v)) = β(ω-ι(φρ(ιι)))φρ(ν) 'φΡ(υ,γΗ(1)φρ(υΥ ^P{u)bh(r)ipp{v) t With Proposition 5.14 we have completed the decomposition of a locally ambient geometry into its three parts as mentioned in the introduction to this section. Moreover it should be clear that the original locally ambient geometry can be unambiguously reconstructed from these parts. The following exercise describes the mean curvature associated to a locally ambient geometry. Exercise 5.16. Let eb e2,..., en be an orthonormal basis for TX(M). Show that the normal vector Tr Β = Σ{ B(ei, ei) is independent of the choice of this basis. (The normal vector ^Tr В is called the mean curvature vector at χ G M. Its length is called the mean curvature at x.) □
§5. Subgeometry of a Riemannian Geometry 253 Now we pass to a study of the Weingarten maps associated to a locally ambient geometry. Definition 5.17. Let (Ρ,ω) be a locally ambient geometry on Μ of codimension r. The second fundamental form associated to (Ρ, ω) is the bilinear symmetric form В G Hom(T(M) φ T(M),v) given in Proposition 5.14. Given a normal vector X G vx, χ G M, the corresponding Weingarten map or shape operator Lx G End(Tx(M)) is defined by (Χ,Β(η,υ))„ = (u,Lx(v))M. * Lemma 5.18. Lx is a self-adjoint map depending linearly on X. Proof. The calculation (u,Lx(v)) = (X,B(u,v)) = (X,B(v,u)) = (v,Lx(u)) = (Lx(u),v) shows that Lx is self-adjoint. For the linearity in X, we have (u, LaX+bY (v)) = {aX + bY, B{u, v)) = a(X, В(щ ν)) + b(Y, В(щ ν)) = a{u, Lx{y)) + b(u, LY(v)) = (u, aLx(v) -f- bLY{v)), and so Lax+bY — aLx + bLy. Ш Since the endomorphisms Lx are self-adjoint, all their eigenvalues are real. This leads to the following definition. Definition 5.19. The eigenvalues of Lx G End(TX(M)) are called the principal curvatures associated to the normal vector X G vx{M). * Definition 5.20. The principal curvatures of Μ С TV are constant if, for any curve σ: I —» Μ and any parallel normal field X on Μ defined along σ, the set of eigenvalues of Lx is independent of the point on σ where they are calculated. ft Next we compare the Weingarten maps to the matrices h(i). Lemma 5.21. (i) Lx(v) = Σ^^ΨρίΧ^φ-^Η^φ^ν)). (ii) h(i) = φρο L^i{et) ο φ~ι e End(i). Proof. We have
254 6. Riemannian Geometry {u,Lx(v))M = {X,B{u,v))v = {φρ(Χ),φρ{Β(μ,ν)))η ι f φρ{ν)1Κΐ)φρ(ν)\, = UP(X)A : } · \ \<pP(u)*h(r)tpp(v)J'u Thus, (Vp(u),(pp(Lx(v)))t = ]Γ Ψρ(^ΨΡ(η)ιΗ{1ζ)φρ{ν) \<k<r = {ψρ{η), Σ 4>p{X)kh(k)4>p{y))i. \<k<r This yields (i). Taking X — φρι{βί) (where e*, 1 < г < η, is the standard basis of t) in (i), we get Lx(v) = φρ1(1ι(ι)φρ(υ)), which is (ii). ■ Corollary 5.22. The Weingarten maps all commute with each other if and only if all the h(i) commute with each other. Principal Curvatures for Relatively Flat Normal Bundles Definition 5.23. Let TV be a Riemannian manifold. An immersion /: Μ —» TV has a relatively flat normal bundle if β Α β1 = 0 or, equivalently, d^ + 7 Λ 7 = Ω, where Ω is the curvature entry on Pr of the (3,3) block of the locally ambient geometry. * Example 5.24. A codimension-1 submanifold of Euclidean space is automatically relatively flat since Ω = 0 in a flat geometry, and 7 = 0 since 7 is a 1 χ 1 skew matrix. More generally, a codimension-r submanifold of Euclidean space is relatively flat if and only if the Ehresmann connection on the normal bundle (cf. Definition 2.4) is flat (i.e., has curvature zero). ♦ Proposition 5.25. The following statements are equivalent: (i) Μ С N has a relatively flat normal bundle. (ii) h(i)h(j) = h(j)h(i) for all 1 < ij < r. (iii) The Weingarten maps Lx all commute with each other. Proof, (i) <3> (ii). By definition, (i) is equivalent to β Α β1 = 0. Thus, this equivalence is the same as (/? Λ /?%· = 0 & h(i) and h(j) commute.
§5. Subgeometry of a Riemannian Geometry 255 We have (/? Λ β% =βιΛ β] = 0έ/ι(ζ) Λ (θιΗ^)Υ (cf. Proposition 5.14) = ΘιΗ{ϊ) Λ h(j)0 = Σ °kh(i)kr Λ h{J)riei krl = Σ ( ^(h(i)krhU)rl ~ h(i)krh(j)H) \θίςΑθ1 k<l \ r J к<1 Since h(i) and h(j) are symmetric matrices, it follows that [h(i),h(j)] is skew symmetric. Hence, (/? Λ β% = 0& [ft(i), ft(j)]w =0 for all к < I & [ft(i), h(j)] = 0. The equivalence (ii) <=> (iii) is just Corollary 5.22. ■ Corollary 5.26. Suppose Μ С N has a relatively flat normal bundle. Define Лх = {λ e и* | Lxv = X(X)v for some ν e TX(M) - {0} and all X e vx), and set TX(M)X = {ve TX{M) | Lxv = X(X)v for all X e vx). Then TX(M) — φλΕΛ Τχ(Μ)χ is an orthogonal decomposition of the tangent space. Proof. Since the collection of all Weingarten maps is a commutative subset of End(Tx(M)), we know from linear algebra that they are simultaneously diagonalizable (cf., e.g., [N. Jacobson, 1953], pp. 132-133). This gives a decomposition TX(M) = ®Ei such that each Ei is an eigenspace for every Weingarten map. Let \i{X) denote the eigenvalue of Lx on Ei. Then the linearity of Lx in X implies that λ^ G v*. Thus, λ^ G Ax and Ei С Тх(М)\г. Since TX(M)X Π Τχ(Μ)μ = 0 for λ φ μ and TX(M) = ΘΕ{ С 0Τχ(Μ)λί, the direct-sum decomposition follows. To see that it is orthogonal, choose и е τχ(Μ)μ, ν e TX{M)X, λ φ μ. Then there is a vector Χ Ε vx with λ(Χ) φ μ(Χ) so that (A(X) - μ(Χ))(η, ν) = (X(X)u, ν) - (щ μ(Χ)ν) = (Lxu, ν) - {и, Lxv) = 0. Thus, (щ ν) = 0. ■ Definition 5.27. Suppose that f: Μ —> N has a relatively flat normal bundle. Referring to Corollary 5.26 we have the following:
256 6. Riemannian Geometry (i) The elements of Ax are called the dual principal curvatures4 at x. (ii) The eigenspaces TX{M)\ С TX(M), λ G Ax given in the proof of Corollary 5.26 are called the principal subspaces (or directions if they are one-dimensional) of TX(M) at x. (iii) The decomposition TX(M) = ®TX(M)\ is called the principal curvature decomposition. (iv) The normal vector Xx G T^M)1- given by (Χχ,υ) = λ(ν) for all ν G T^M)1- is called the mean curvature vector associated to λ Ε A (or to the principal subspace TX(M)X С TX(M)). Φ Exercise 5.28.* Let Μ С N have a relatively flat normal bundle. (i) Show that B(u,v) = {u,v)Xx for all u,v G TX(M)\. (ii) Assume that Ax contains a constant number of elements for χ G £/, a simply connected open set in M. Show that there is a canonical identification among the Λχ, χ G Ε/, and that under this identification the X\ are smooth normal vector fields along U. (iii) Deduce from (ii) that, for each λ G Λχ, the TX{M)\ yield a distribution along U. (iv) Show that the mean curvature vector is - ]T nxXx, where nx = dim T{M)\, λΕΛ and the mean curvature μ is given by μ2 = (1/n2) ]Γ^ λΕΛ ηκη\{Χκ, Χχ). □ This exercise allows us to identify the elements in different Ax for χ varying in a small, open set when #AX is constant. In fact, more is true in this case. Proposition 5.29. Suppose the immersion f:M—>N has a relatively flat normal bundle and the function #AX: Μ —> Ζ is constant. Then (i) rt\ = dim Τχ(Μ)χ is independent of x. Choose any orthogonal decomposition i = φλΕΛ ix, where dim ίχ = ηχ. (ii) There is a canonical reduction o/On(R) χ Or(R) —> Pr —> Μ to a principal bundle (хлелОпл(Я)) x Or(R) -> PA -> Μ 4In the case of a hypersurface, the numerical values of the elements of Λχ on the unique (up to sign) unit normal vector are what are usually known as the principal curvatures.
§5. Subgeometry of a Riemannian Geometry 257 given by Pa = {p G Pr | φρ(Τχ(Μ)χ) = ix for all λ G A, where χ = π(ρ)}. (iii) Regarding h(k) as taking values in End(i), we may also describe Рд as PA = {pe PT\ h(k) | ik = Χ(φ~ι(ek))idix for all X G Λ andl < к < г}. Set aSi — Xi((p~l(es)), 1 < г < η, η+1 < s <n + r, where Xi G Λ is determined by the condition e^ G i\t. (iv) The X\eAOnx(R) bundle Рд/(1 x Or(R)) —> Μ is a reduction of the bundle On(R) -> Ptan - M. (v) The (2,2) 6/ос/ь о/ the curvature function for Ptan (i-e., for the sub- manifold M) is given by #2,2+ Σ №Д;К*Ле*®е^ l<i,j<n where Xi is the normal vector given by {Xi, —) = X, where i is determined by the condition e^ G t\. #2,2 is the (2,2) block of the locally ambient curvature function restricted to Рд. It is Or(R) invariant and so is basic for the bundle Рд —» Рл/(1 х Or(R)) = Ptan- (In VaT~ ticular, the multiplicities of the Xs correspond to the multiplicities of the Xs.) Proof, (i) This follows immediately from Exercise 5.28(iii). (ii) We shall only verify that the fiber is as stated. Let ρ G Рд and h G On(R) x Or(R). Then phePA& ψρη(Τχ(Μ)χ) = ix for all λ G Λ <^> Αά(ϊι-ι)φρ(Τχ(Μ)χ) = ix for all λ G Λ <^> Ad(ft_1)tA = U for all λ G Λ ^ h e (xxeAOnx(R)) χ Or(R). (iii) Assume that ρ G Рд. If ν G U, then h(k)v = φρο L^-i(efc) ο φ-ι(ν) = φΡ(Χ(φ~1 ^))φ~ι(ν)) = Х((р~1(ек))у, and so Рд С {р G Pr I h(fc)|tA = A(^"1(eA:))idtA for all λ G A and 1 < к < г}.
258 6. Riemannian Geometry The reverse inclusion is similar. (iv) The inclusion Рд С Pr induces an inclusion Рд/Ог(К) С Pr/Or(R) = Ptan. (ν) We have (an 0 0 \ /αηθι ■■■ αι„#η О '·· 0 I so/J= I : О О Clin/ \0"п1&1 '*" Q"nn@n > and /?'л/? = ( Ι Σ α«α^Ι θ*Λθΐ \ \n+l<s<n+r J Now asi — \i((p~l(es)) = (Xi,ip~l(es)) is the 5th component of Xi in the orthonormal basis (^^(еД η + 1 < s < η + r, for ^(M)-1. Thus, 7 ^ aSidsj — (Xi,Xj). n-\-l<s<n-\-r From Eq. (5.1), the (2,2) block of the curvature on Рд is Ω22 = da + α Λ α — /?* Λ /?, from which the result easily follows. ■ Definition 5.30. The reduction of On(R) —» Ptan —* Μ described in Proposition 5.29(iv) is called the principal curvature reduction. ft Definition 5.31. Let Ρ —» Μ be a principal bundle and Q С Ρ be a reduction of it. Let X be a distribution on P. We say that the reduction is compatible with the distribution if, for each point g G Q, we have Lq С Tq(Q). * In general, there is no reason why the principal curvature reduction should be compatible with the Levi-Civita connection. Example 5.32. Consider the case of an umbilic-free surface Μ in 3-space with no flat points. The curvature reduction Q С Ptan is a principal bundle with group Oi(R) x Oi(R) = Z2 x Z2 and so is a covering space of the surface. Compatibility with the Levi-Civita connection would mean that the Levi-Civita connection form a would vanish on Q. Choosing a local section of Q would yield a gauge in which a vanishes, and hence so would the curvature. This would imply that Μ is flat. ♦ For a submanifold Μ С N with a relatively flat normal bundle, the condition of constant principal curvatures has a special significance in that the vector fields X\ are parallel. Lemma 5.33. Let Μ С N be a submanifold of a Riemannian manifold. Then the following conditions are equivalent:
§6. Isoparametric Submanifolds 259 (i) The principal curvature sections λ Ε A of the normal bundle are parallel along any curve. (ii) The principal curvatures of Μ are constant. (iii) The normal vector fields X\, λ G Λ, are parallel along any curve. Proof. Let σ be a curve on Μ and let Ζ be its tangent field. Let X be any parallel normal field along σ (i.e., normal to M) and let λ Ε Λ. (i) & {U) λ is parallel along σ ^ DzX — 0 along σ Γ λ(Χ) constant along 1 for all X parallel along ■Λ (ii) 45 (iii) ( \(X) constant along σ Λ ( {Χχ, X) constant along σ Λ \ for all Χ parallel along σ J (^ for all X parallel along σ J & X\ parallel along σ. ■ §6. Isoparametric Submanifolds Elie Cartan [E. Cartan, 1938] uses the term isoparametric ("same parameters") to refer to hypersurfaces Mn С ζ?£+1 having constant principal curvatures (these are the "parameters" of the submanifold) in a space Q™+1 of constant curvature c. Definition 6.1. A submanifold Μ С N of a Riemannian manifold is isoparametric* if it has constant principal curvatures and its normal bundle is relatively flat. * Isoparametric submanifolds of Riemannian space forms have been studied extensively in the past decade or so (see [C.-L. Terng, 1993] and the references given there). Cartan showed that isoparametric hypersurfaces of Euclidean space are spheres, hyperplanes, or a product of the two. He also 5 In the more usual definition, one assumes that N is a Riemannian space form and that the normal bundle of Μ С iV is flat. However, our definition generalizes this one since the flatness follows from the commuting of the Weingarten maps in this case; cf. [C.-L. Terng, 1993].
260 6. Riemannian Geometry showed that in hyperbolic space isoparametric hypersurfaces are totally umbilic, while in spheres there are many examples. Thorbergson [G. Thor- bergson, 1991] has shown that any isoparametric submanifold (Rn+r,Mn) of codimension r > 3 in Euclidean space is isometric to a pair (Te(G/H), Ad(H)v), where ν e Te(G/H), for some symmetric space G/H. For hypersurfaces in spheres ([H. Ozeki and M. Takeuchi, 1975 and 1976]) and for codimension 2 in Euclidean space [D. Ferus, H. Karcher, and H. Mun- zner, 1981], there are examples of isoparametric submanifolds that are not locally symmetric. We shall study only the case of hypersurfaces. Let Mn С Qc+1 be a hypersurface in a Riemannian space form of curvature c. Let us choose a section of the principal curvature reduction (Proposition 5.29(iv)) so that the 1-forms that make up the Cartan connection on Ρ pull back to Μ to yield a gauge of the form 0 θ 0 0 ω efh 0 -he 0 where h = h(l) = diag(Ai,..., λ„), the diagonal matrix of principal curvatures. The structural equation is then 0 0 0 άθ + ωΑθ άω + ω Λ ω - h0 Λ 6lh -d(h0) - ω Λ (ΗΘ) 0 * 0 Ό 0 0Ν 0 ев Α θ* 0 | , (6.2) ,0 0 0, whence, in particular, 0 = -d(h0) -ωΛϊιθ = -(dh) Λθ-Ηάθ-ωΛϊιθ (6.3) = -(dh) Λθ + ΗωΛθ-ω Λϊιθ. Umbilic Hypersurfaces Our first application of Eq. (6.3) is to the case when all the eigenvalues of h are equal. Definition 6.4. Let Mn С Qc +1 be a hypersurface in a Riemannian space form of curvature с. М is called an umbilic hypersurface if the principal curvatures of Μ are all equal at each point. ft Proposition 6.5. If η > 1, the principal curvature of a connected umbilic hypersurface Mn С Q™+1 is constant on M. Moreover, Μ is a Riemannian space form of curvature c + A2.
§6. Isoparametric Submanifolds 261 Proof. In this case h is a multiple of the identity, so hu Λ θ = ω Λ Ηθ. Thus, Eq. (6.3) becomes (dXI) Α θ = 0, or dX Λ 0< = 0 for 1 < i < n. Thus, there exist functions a^ such that d\ = a^ for each i. Since η > 2, it follows that a2^2 — ai#i — 0· Hence, a,2 = a\ = 0 and dA = 0. Thus, λ is constant. The induced geometry has Cartan connection I й J with curvature form 0 0 \ = /0 0 άθ + ωΛθ άω + ωΛω)~\0 (λ2 + с)в Л Θ1 where the last equality comes from the structural equation (6.2). Cartan's Formula for Isoparametric Hypersurfaces We return to the case where Μ is an isoparametric hypersurface in a space. Then the diagonal matrix h is constant over M. Equation (6.3) now reads hu Λ θ = ω Λ Ηθ or, in coordinate form, V"\Ai - λ3)ω^ Λ 6j = 0 for г = 1,... ,n. (6.6) Let us write Uij = Y^k9ijk0k (since the #s constitute a basis). Since ω^ — —ujji it follows that gijk = —gjik· Putting this in Eq. (6.6) yields У^(Аг - Xj)gijk0k Л 9j = 0 for г = 1,..., n, 3 к or (λ< - \j)gijk = (λ. - Afc)0tAy for all z,j, fc = 1,..., n. (6.7) Lemma 6.8. Let a^k — (K — ^j)9ijk- Then aijk is fully symmetric in all its indices and it vanishes if two indices are equal. Proof. Equation (6.7) shows that a^k is symmetric in the last two indices. Moreover, both g^ and (A^ — A^) are skew symmetric in the indices г and j, so aijk is also symmetric in the first two indices. It follows that а^к is symmetric in all the indices. Since а^к obviously vanishes if the first two indices are equal, by symmetry it must vanish if any two indices are equal. Corollary 6.9. If either (i) λ^ φ X3 — Xk or (ii) A^ = λ7 φ Xk, then 9ijk = 0. Proof, (i) Xj = Xk => ajki = 0 => aijk = 0 => gijk = 0 (using λ; φ X3). The other case is similar. ■
262 6. Riemannian Geometry Next, consider the exterior derivative of the equation ω^ = Y^kgijkQk- It is к к = Σ dViJb Л Ок - 5Z 9ijkWkl Λ θι к kl = /_^ dgijk Л Ok — 2_^gijkgkirQr Α 0z, к klr where we use the (2,1) block of the structural equation (6.2) for the second equality. From the (2, 2) block of the same equation, we get, for all 1 < hj <n, cOiAOj — 2^dgijkAek — 2_^gijkgkirurA0i+2_^9irkgrjiukA0i — XiXj0iA6j. к klr klr The only terms in Σ к dg^k A 0& that can contain 0* Л 0j are dg^ Л 0^ + dgijj AOj. If Xi Φ Xj, these terms vanish by Corollary 6.9, and equating the coefficients of 0* Λ 6j will yield с= / ^gjjkygkjj gkij) ~г у j^gjrxgrjj girjgrji) a%aj к г XiXj + c = ^{-gijkgkji + gijkgkij - gikjgkji) (since giri = 0). (6.10) Lemma 6.11. If Xi φ Xj, the only terms giving a nonzero contribution to the sum in Eq. (6.10) are those for which the index к satisfies Xk φ Χχ and Xk φ Aj. Proof. Consider a term -gijkgkji + gijkgkij + gikigkjj - gikjgkji for which *k = λ< Φ Xj. By Corollary 6.9, we have gijk = gkjj = g%kj = 0, so the term vanishes. The case οϊ Xk φ Xi = Xj is similar. ■ Thus, the only nonzero terms in the sum in Eq. (6.10) arise from indices fc for which all three of λ^, λ^, and λ^ are distinct. In particular, each of the gs appearing in these terms of the sum may be expressed in terms of the as (Lemma 6.8), and we get all к such that \ г j κ j 1 + or Xi ~ Xj Xk Xi Xi ~ Xk Xk ~ Xj
§6. Isoparametric Submanifolds 263 This formula quickly simplifies to yield Cartan's formula, a? A4Ai+c = -2 £ ijk all к such that (K - Afc)(Ajf - Afc) Proposition 6.12. 7/ an isoparametric hypersurface in a space form of curvature с has only two distinct principal curvatures, \\ and \2 say, then λιλ2 = -с. Proof. The right-hand side of Cartan's formula vanishes. ■ In [E. Cartan, 1938] other conclusions are drawn from this formula. In particular, it is shown that a connected isoparametric hypersurface in Mm С Q^+1 has at most two eigenvalues if с < 0.
7 Mobius Geometry Hermann Weyl defined a conformal structure on a manifold Μ as an equivalence class of Riemannian metrics, where two metrics are equivalent if they differ by a factor that is a smooth positive function on M. Suppose Μ is equipped with a conformal structure and u,v G TXM. Then the angle between и and ν makes sense, as does the ratio of the lengths of и and v, but the lengths themselves have no meaning. In Elie Cartan's view, a "conformal structure" on Μ is a Cartan geometry on Μ modeled on the Klein geometry whose space is the η-sphere and whose principal group is the full group of conformal (i.e., angle preserving) transformations. This group turns out to be the projective Lorentz group in dimension η+ 2. This is the geometry that we call the Mobius geometry. Weyl also considered a structure on Μ intermediate between the two described above which amounts to a Cartan geometry on Μ modeled on Rn with its full group of conformal transformations as the principal group, which is the group of Euclidean similarities. This is the geometry we call the Weyl geometry. If we regard Sn as being Rn U {00} (via stereographic projection), then the subgroup of the projective Lorentz group which fixes 00 is the principal group of the Weyl model. A recent survey and overview of many results about conformal geometry and its generalizations from a somewhat different point of view than ours may be found in [M.A. Akivis and V.V. Goldberg, 1996]. In §1 we study the Mobius and Weyl model geometries. In §2 we pass to the consideration of Cartan geometries with these models. In particular, we show that every Mobius (or Weyl) geometry on a manifold Μ deter-
266 7. Mobius Geometry mines a conformal structure on M. We also study some basic properties of Mobius and Weyl geometries, including special curvatures and normal geometries. In §3 we solve the equivalence problem for conformal structures. That is, for manifolds of dimension >2, we show that to a given conformal structure there corresponds a unique normal Mobius geometry giving rise to the original conformal structure as in §2 (Proposition 3.1). Since a Riemannian structure determines a conformal structure, which in turn determines a normal Mobius geometry, the curvature of this latter geometry is a conformal invariant of the original Riemannian geometry. This turns out to be the Weyl curvature (Table 6.2.5) of the Riemannian geometry, and so we have a conceptual proof for the conformal invariance of this part of the Riemannian curvature (Theorem 3.8). The section ends by solving the equivalence problem for Weyl structures (Definition 3.11) in terms of Weyl geometries (Proposition 3.14). The following figure summarizes the relationships among these geometries in dimensions >2. < conformal structure or normal Mobius geometry generalization > > ι Weyl structure ' or ι Weyl geometry ^ or special curvature ^Mobius geometry) generalization > general ] Mobius V I geometry Conversely (again for dimension >2), an arbitrary Mobius geometry canon- ically determines a normal Mobius geometry, but of course this correspondence is not one to one. In §4 we study immersions of submanifolds in normal Mobius geometries, showing how the ambient geometry localizes along the immersion to give data (a "locally ambient geometry"; cf. Definition 4.8 and Lemma 4.14) determining it. We then study how the localized geometry decomposes into tangential, normal, and gluing data1 (see Propositions 4.18, 4.20, and 4.21), and we characterize such data in simple cases. If the submanifold is not a surface, then these three data determine the locally ambient geometry from which they arise. For a surface immersed in Rn, we apply this to study the relation between Riemannian and Mobius geometries. Although the conformal structure inherited by this surface is necessarily flat, the (nonnormal) Mobius geometry it inherits is not flat. When η — 3, its curvature may be identified with the Willmore form (H2 — K)dA. In Riemannian geometry, the Gauss curvature К is regarded as intrinsic (since it can be calculated from the metric induced on the surface) while the mean curvature is regarded as extrinsic. Thus, the Willmore form, while appearing to be extrinsic data from the I.e., second fundamental form data.
§1. The Mobius and Weyl Models 267 Riemannian perspective, is seen to be intrinsic data from the perspective of the Mobius geometry2 induced on M. Section 4 ends with a brief comparison with Aaron Fialkow's work on the classification of submanifolds. In §5 we classify the immersions of vertex free curves in normal geometries of two and three dimensions. The method here is alternative to, and less general than, the one given in §4, but the form of the answer is simpler. In §6 we use the method of §5 to study the class of umbilic free surfaces in the conformal three sphere. This gives an approach that is an alternate to the one given in §4 and avoids the difficulty with the exceptional case of surfaces referred to above. Throughout this chapter, (G,H) = (Mobn(R),Mobn(R)0) denotes the Klein geometry of the Mobius group and its subgroup (see Definition 1.7); (g,i)) denotes the corresponding pair of Lie algebras (see Exercise 1.13); Mq will denote (a space canonically identified with) the homogeneous space G/H (see Definition 1.2); and Μ itself will denote the space of an arbitrary Cartan geometry modeled on (g, ()) with group H. §1. The Mobius and Weyl Models In this section we study the standard model of Mobius η space which, as we will see, may be identified with the η-sphere Sn together with its group G consisting of the group of all conformal transformations. To describe the group G and its action, we will regard Sn as the space of one-dimensional lightlike subspaces of (n -f 2)-dimensional Lorentz space (cf. Lemma 1.3). Thus, G will be related to the Lorentz group. Lorentz Space and Mobius Space Lorentz space is L = Rn+2 (with basis eo, eb ..., en+1) equipped with the indefinite quadratic form q^: L —» R given by 4l{x) = -2xoXn+i + Σ x^' \<k<n In the given basis, the matrix for this form is / 0 0-1 Ση+Μ = 0 In 0 \-l 0 0 2It was this fact that originally kindled the author's interest in Cartan's es- paces generalises.
268 7. Mobius Geometry Lemma 1.1. The affine map ^:Rn+1 —» L sending (Уо,У1,---,Уп) »-► ί--?=(1 + 2/ο)»2/ι»···»2/η,--^(1-2/ο)) zs an affine isometric embedding of Euclidean space into Lorentz space. Proof. The derivative of ψ at χ G Rn+1 is ψ*χ:Τχ(Κη+ι)-+Τφ{χ)(τ,), In particular, дь(Ф*х(у)) = qL l-j=v0,vi,...,vn,—y=u0 = 2Й"°)(71''°)+1<?</'=?°М· where qo is the standard metric on Rn+2. ■ Definition 1.2. A vector ν G L is called lightlike if qij(v) = 0. A subspace of L is called lightlike if it consists of lightlike vectors. The set of lightlike points in P(L) (the projective space of L; cf. Example 1.1.3) is denoted by M0 and is called the projective model of Mobius η space. Φ Lemma 1.3. The map ^:Rn+1 —» L induces a canonical diffeomorphism &Sn -» Af0. Proof. According to Example 1.1.3, the affine map ^:Rn+1 —» L given in Lemma 1.1 defines an affine coordinate system ^:Rn+1 —» P(L) for P(L). This coordinate system includes all points of P(L) except for those corresponding to the one-dimensional subspaces lying on the hyperplane ^o Η- ^η+1 — 0, and these latter are never lightlike. Thus, the space of lightlike lines in L lies in the image of ψ. In the coordinate system ψ, the condition of being lightlike is 1/S-1+ Σ Ук = ° or Σ vl = 1· ■ 1<Α:<η 0<Α;<π The Canonical Line Bundle The canonical (real) line bundle over P(L) (cf. Example 1.3.5) may be restricted to yield a line bundle Ε over the subspace of lightlike lines, M0 С P(L), which we again call the canonical line bundle. More precisely, we have the following definition. Definition 1.4. The canonical line bundle over M0 is Ε = {(ν, I) G LxM0 | vel}. *
§1. The Mobius and Weyl Models 269 Mobius Group Consider the Lorentz group £n+i,i(R) = {g £ G/n+2(R) | g Ση+ι5ι# = Ση+ι?ι}. It is clear that Ln+i?i (R) is a Lie group that acts smoothly on the projective space P(L) of lines in L and that this in turn induces a smooth action on the subspace of lightlike lines M0 С P(L). However, the induced action on M0 is not effective since ±1 acts trivially. Exercise 1.5. Show that the action of Ln+i>i(R) on P(L) is transitive on the Mobius space of lightlike lines M0 С P(L). □ Lemma 1.6. The action o/Ln+i>i(R) on Μβ has kernel ±1. Proof. As we mentioned above, the kernel К obviously contains ±1 so we need only show the reverse inclusion. Now Ln+i>i(R) acts on the canonical line bundle Ε (cf. Definition 1.4) according to the formula g(v, I) = (gv,gl). Ifg fixes I e M0, then for each ν 6 I we have gv = Xtv for some Χι e Spec(#), the set of eigenvalues of g. Thus, g £ К implies that gv = λιν (where I is the span of v) for every lightlike vector v. Since M0 (= Sn) is connected and Χι depends continuously on I 6 Sn and Spec(#) is discrete, it follows that Χι is independent of I, say λ/ = λ. Thus, gv = Xv for every lightlike vector weL. Since lightlike vectors given by e0,en+i,eJ- + -7=(e0 + en+i) for 1 < j < η 1 span L, it follows that g = XL Finally, we note that λ Ση+ι?ι = g Ση+ι5ι# = Ση+ι?ι => Χ — 1 => λ = ±1. ■ Definition 1.7. The group Mobn(R) = Ln+M(R)/{±/} is called the Mobius group in η dimensions. The pair (G, Я"), where Η — Mobn(R)0 = {h G Mobn(R) | h[eo] = [eo]}, is called the Mobius model in dimension n. Here [eo] denotes the class of eo in P(L). Φ It is the group Η = Mobn(R)o that will be crucial for an understanding of Mobius geometries, so we need to have a way of representing its elements. These may be expressed as matrices, as the following result shows. Lemma 1.8. Let Η be ζ 0 0 0 a 0 0 \ 0 z-4 Λ ° Vo я 1 0 r ?4 1 ζ e R+, a G On(R), q* € R", r = -qq*
270 7. Mobius Geometry (i) Η is a subgroup o/Ln+i?i(R). (ii) Every element of Η has a unique decomposition as a block diagonal times a block upper triangular matrix as above. (iii) Η acts effectively on Mq. (iv) The canonical projection Ln+i?i(R) —» Mo6n(R) induces an isomorphism Η —» Mobn(R)o «so ί/ιαί Я = Mo6n(R)o ma?/ 6e regarded as the matrix group H. Proof. Parts (i) and (ii) are easy. Part (iii) is a consequence of the fact that Η Π {i/} = {/}. Let us deal with part (iv). It is clear that Ln+i?i(R) —► Ln+i?i(R)/{±/} induces a homomorphism Η —► Mobn(R)o and that by (iii) this has trivial kernel. We must see that it is surjective. Any element of Mobn(R)0 may be covered by some element g G Ln+i>i(R), and we must have g[eo] = [eo]. Thus, g must have the form (• • *\ о • • J, 0 0*/ where the zero in the (3, 2) block arises as a consequence of the fact that <7*Еп+1д<7 = Ση+ι?ι. Multiplying g by —/ if necessary (which doesn't alter its image in Mobn(R)0), we may assume g has the form (z * * \ 0 5 • J , 0 0 z~lJ where ζ G R+. It is then an easy matter to use again the fact that g satisfies 9l Ση+ι,ι 9 = Ση+ι,ι t0 deduce that g G H. Ш Exercise 1.9. Prove the assertion of the last sentence in Lemma 1.8. □ Corollary 1.10. Each of G and Η has two components. Η is generated by its identity component He and the element diag(l, — 1,1,..., 1) G H. Proof. The proof of Lemma 1.8 provides a difFeomorphism Η —> R+ x On(R) x Rn, sending (in the notation of Lemma 1.8) h ι—► (z,a,qb). Now On(R) has two components while R+ and Rn are connected. Thus, Η has
§1. The Mobius and Weyl Models 271 two components. Since diag(l, —1,1,... ,1) lies in the nonidentity component of #, it, together with #e, generates H. On the other hand, for η > 1, G/H — Sn is connected and simply connected. Thus, the long exact sequence of homotopy groups 1 = n\(G/H) —» тг0(Н) —» kq(G) —► 7To(G/H) = 1 shows that Я and G have the same number of components (cf. [D. Husemoller, 1966], p. 92). We leave the case η = 1 for the reader. ■ By Lemma 1.12 and Corollary 3.7 ahead, we see that the group G may also be described as the group of all conformal transformations of the standard η-sphere (cf. also [M. Berger, 1987], Theorem 18.10.4). The precise definition of conformal transformations is as follows. Definition 1.11. A diffeomorphism 0:5n —» Sn is called a conformal transformation of the standard η-sphere if, for each χ G 5n, φ*χ: Tx(Sn) —> Тф(х)(Зп) preserves the canonical Riemannian metric qo up to scale. * Lemma 1.12. Mobn(R) acts on Sn as a group of conformal transformations. Proof. The action of Ln+i?i(R) on Sn as defined above may also be described by the formula дф{х) = μ(χ)ψ^χ), where xGS" and μ(χ) G R* is chosen so that μ{χ)ψ^χ) G Im ψ. Differentiating, we see that for any ν G Tx(Sn), we have 9Ψ*χ(ν) = μ*χ(ν)Ψ(9χ) +/i(x)iffx(^W)· Since g*x(v) G Tgx(Sn), it follows that gx and g*x(y) are perpendicular in Rn+1. Hence by Lemma 1.1, it follows that tp(gx) and ip*gx(g*x(v)) are perpendicular in L. Since we also know that ip(gx) is light like, it follows that qo(v) = Яь(Ф*х(у)) (by Lemma 1.1) = Ql(9iI>*x(v)) (by the definition of Ln+M(R)) = qL^*x{v)i>{gx) + μ(χ)Ψ*9χ(9*χ(ν))) (by the formula above) = (ϋ,(μ*χ(ν)Ψ(9χ)) + 2{μ^χ{ν)^χ))ιΈη+ιΛ{μ{χ)φ^9Χ^^χ{ν))^ + ςϊ,(μ(χ)Ψ*9χ(9*χ(ν))) = qL^{x)il>*9x(9*x(v))) (by the remarks above) = /i(x)2iL(^5x(5*xW)) = μ(x)2qo(g*χ(v)) (by Lemma 1.1). Thus, the action of Ln+i5i(R), and hence of Mobn(R), preserves go UP to scale. ■
272 7. Mobius Geometry Lie Algebras for the Mobius Model Exercise 1.13.* (i) The pair (G,H) has Lie algebras (g, f)), where i/z Q 0\] (fzqO \] 3= { Ρ s <?' > , f) = < 0 s 9* > , where sl = -s, l\° ρ1 -*)) IVo о -z)\ and g has blocks down the main diagonal of size Ιχΐ,ηχη, and lxl. (ii) g decomposes (as a vector space) as g = ρ θ 5 θ 3 θ q, so that ()=5030q, where P= < Ρ 0 0 >, s = fo 0 ρ 0 \0 pl ζ 0 0 0 0 0 0 0 0 0 0 —ζ 0 0 0 0 0 0 0 s 0 q 0 0 0N 0 0, 0 g4 0 (iii) Show that q,3 Θ q and $ 0 q are ideals of I). (iv) Show that for η -φ 4 the ideals in (iii) are the only ideals of \). (v) Show that for η = 4 we have s = 00(4) = so(3) Θ so(3) = WP+ Θ W_ as Lie algebras. Show also that the complete list of ideals of \) in this case is q,3®4»5 0q,W+0q,iy_eq,3elV+eq, and 3 Θ W_ Θ q. □ Exercise 1.14. Setting g_i = p, go — 5 0 3, and gi = q, show that [9a, 9b] С да+ь for all a and 6 (where ga = 0 if α φ -1, 0, 1). □ Adjoint Action It is a simple computation to show that the adjoint action of h G Η on g/f) is given by Ad(ft)v = sz-1?;, where /z 0 0 \ /1 q r\ г h= 0 5 0 0 1 ςΜ and r = -ад*. VOOz-WVOOl/ 2 Quadratic Form Invariant up to Scale on g/i) Lemma 1.15. There is a quadratic form qg/^'-g/t) —» R гу/ггсЛ zs Ad(H) invariant up to scale. Such a quadratic form is unique up to scale.
§1. The Mobius and Weyl Models 273 L-{0}- A Ψ a Ψ with maps 5Г-*->А4>< G/H diffeo. diffeo. Proof. Define χ h- (w, [w]), (ge0,[gej)< (v,0 I [v] + gH FIGURE 1.16. %/b Ι Ρ * 0 ρ* <fo(p) = Σ Pi KKn The formula above for the adjoint action implies that qQ/i)(Ad(h)v) = z~2%/\){v)· As f°r tne uniqueness, it is well known (and, in any case, easily shown) that qo is unique up to scale among the quadratic forms invariant under the subgroup On(R) С Η. Hence it is also unique up to scale under the larger group. ■ Identifying Mobius Space with G/H The Mobius model may be described in three ways: as the homogeneous space G/H; as the space Mq of lightlike lines in P(L); and as the n-sphere Sn. These descriptions are related in Figure 1.16, where Ε is the canonical line bundle over M0. Moreover, there is a left action of G on every space in this diagram. (The action on Sn is of course defined via the map ψ.) It is clear that all the maps are G-equivariant except for φ. As usual, we let N = e0 G Rn+1 be the "north pole" of Sn. Since ψ(Ν) = [e0], it follows that the composite map G/H —» Sn sends gH —» gN. Lie-Theoretic Interpretation of the Quadratic Form qQ Lemma 1.17. Let w G Te{G/H), and write it as w = ne*(v), where ν = 0 pl e 9 = TeG. Then under the identification G/H —» Sn given by Figure 1.16, ν corresponds to the tangent vector и — (Ο,ρι,... ,pn) G IV (5n) С T/v(Rn+1).
274 7. Mobius Geometry Proof. The map a: G —» L sending g ι—> ge$ has derivative a*e: % = TeG —> L sending ν ι-* г>е0. The derivative of the commutative diagram in Figure 1.16 yields the following commutative diagram. T\m(L) >7<>ь,ы)(£)<- Ψ*Ν TN(Sn)· Ψ*Ν •Te(G) в ν Я*£ ->7^(Μ0)< T[e](G/H) э vv For the proof of the lemma, it suffices to show that the elements ν G Te(G) = g and iP*n{u) £ Teo(L) = L correspond under a*e. But by the definition of ψ (Lemma 1.1) and the definition of a, Ψ*Ν(ν>) = (0,pi,...,pn,0) = | p /0 0 0 a*e(v) = ρ 0 0 \0 p* 0 The diffeomorphism С/Я" —» Sn sending gH ι—► #iV determines a quadratic form up to scale on TxSn for each χ G 5n as follows. Fix # G G such that χ = gN, and define the quadratic form as the composite Tx(Sn) £■ T[g](G/H) ^ g/t) 4-^X R. Of course the choice of g is not unique. It may be varied by right multiplication by any element (z * * 0 * * I G #, 0 0 z~\ which will replace qg/btpg by qg/^9h = q9/b(Ad(h-l)tpg) = z2qQ/b(tpg). Moreover, the following diagram is commutative, where the vertical arrows are induced by the left action of 7 G G. 1 i > Trx(S") < — T[n](G/H) ^
§1. The Mobius and Weyl Models 275 From this it follows that the quadratic form we have constructed on T(Sn) is invariant up to scale under the action of G. Lemma 1.18. Up to scale, the quadratic form onT(Sn) constructed above is the standard one. Proof. By Lemma 1.12 and the remarks above, both qo and the form constructed above are invariant up to scale under the action of G on Sn. Since this action is transitive, it suffices to prove the two forms agree at N. But from Lemma 1.17, we see that /• • 0N u = (0,p) = (0,pu...,pn)eTN(Sn)cTN(Rn+l)^[p * * \0 pl * '■■ a/b *-* Яв/ъ(р) = Σ P2i=^o(v)· Ki<n Ricci Hornornorphisrn For the next definitions, and for later purposes, we once again need the Ricci homomorphism, denoted by Ricci, which in the Mobius context is given as the composite of the following maps. Ricci:Hom(X2(Q/b),s) « \2{φ)* 0 5 id-^d A2(g/i))* 0End(g/i)) - A2(g/0)* 0 а/Ь ® (fl/W* contra-^°^ad (g/0)* 0 (в/Ы t* Aw* 0^01/;* »-► (u*(v)t* - i*(v)w*)0'u;* (1.19) Exercise 1.20. Show that q and q Θ s are Η submodules of i) under the adjoint action. It follows that we may regard 5 as an Я module via the canonical isomorphism s w (q0s)/q. Show that Ricci is an H module map. □ Definition 1.21. In Mobius geometry the normal Η submodule of Hom(A2(g/i)), fj) is the kernel of homomorphism Ricci : Hom(A2(g/i)),5) -> (φ)* 0 (fl/f,)V The following exercise is a partial analog of Proposition 6.1.4. Such an analog is possible since the adjoint action of Η on g/fj is the same as in the Riemannian case except for an additional scaling factor (see the subsection preceding Lemma 1.16).
276 7. Mobius Geometry Exercise 1.22.* Let η > 2, and fix an orthogonal equinormal basis e; e £j/f), 1 < г < η (i.e., orthogonal and of equal length with respect to the quadratic form of Lemma 1.15). Let epq С s, 1 < ρ < q < η be the unique (and standard) basis such that ad(epq) corresponds to ep 0 e* — eq 0 e* under the canonical isomorphism End(g/i)) ^ (β/ί)) 0 (β/β)*· (i) Show that Ricci is an Я module homomorphism satisfying Ricci(e* Л e* 0 epq) = 6jpe* 0 e* - ^-pe* 0 e* - <5ipe* 0 e* + 6iqe* 0 e* so that, in particular when гфкф j, Ricd(e* Л e*, 0 ekj) = e* 0 e* + <%e£ 0 e*k. (ii) Show that (a) for η = 2, Ягссг is injective so that the normal submodule is trivial, and (b) for η > 2, Ricci is surjective. (iii) Let η > 2 and set φ = J2ijpq ^ijVqe\ л ej ® eP9 G Hom(A2(g/i)), ()) where α^·Ρ9 is skew symmetric in the first two indices and the last two indices. Assume in addition that akijk = a>hjik for all indices. This latter condition would be a consequence of the Bianchi identities were they to hold (cf. the proof of I in Proposition 3.1 ahead). Show Ricci(ip) = 22akiJkei ® e** ijk (iv) Let η > 2 and set bij = Y^i<k<n(ei л et ® ekj + е!· Л e*k 0 efc<), 1 < hj <n. Show that Ricci(bij) = (n - 2)(e* 0 e* + e* 0 e*) + 26ia ^ e£ 0 e£. Show the elements b^ are linearly independent, and span a submodule i?n С Hom(A2(g/i)),()) (the analogue in Mobius geometry of the Ricci submodule, cf. Definition 6.1.5(ii)). Show that Rn has trivial intersection with Wn — ker Ricci (the analogue of the Weyl submodule). □ The Weyl Model Definition 1.23. The Weyl model is the Euclidean geometry of similarity. Its groups (Сдг,Ядг) are given by zGR+,pGRn,seOn(R) , H^HIn Ϊ ) UeR+,seOn(R)l A
§2. Mobius and Weyl Geometries 277 Exercise 1.24. Show that the Lie algebras of the Weyl model are BN = \(l °0)\zeR,peRn,3eon{R)\, m ζ g R, 5 e on(R) The Euclidean group of similarities Gjy acts on Rn according to the formula (z o^ -i/ , a • χ = ζ (sx H-p). ζ 0 ρ s 0 ρ* 0 0 —ζ If we identify Rn with Sn — N via stereographic projection (where N is the north pole of 5n), then Gn and Η ν are identified with the subgroups fixing TV of the corresponding Mobius groups G and H. Thus, the Weyl model may be regarded as a subgeometry of the Mobius model. In particular, the inclusion ддг С д is given by ζ Ο ρ s §2. Mobius and Weyl Geometries In this section we study some elementary aspects of Mobius geometry. We show that a Mobius geometry on Μ determines a conformal metric on M. We also investigate the various possibilities of special curvatures including normal geometries and Weyl geometries. Definition 2.1. Let Μ be a smooth manifold. A Mobius geometry on Μ is a Cartan geometry on Μ modeled on the Mobius model. ft Let us begin by studying the conformal metric canonically associated to a Mobius geometry. The Conformal Metric Determined by a Mobius Geometry Definition 2.2. Let Μ be a smooth η-manifold. Two metrics qi,q2- T(M) —» R are called conformally equivalent if there is a positive smooth function λ: Μ —» R such that q2 — \q\. A conformal equivalence class of metrics Q on Μ is called a conformal metric. If χ G M, then Qx denotes the restrictions of the elements of Q to TX(M). ft Lemma 2.3. Let Μ be a smooth manifold. A Mobius geometry (Ρ, ω) on Μ determines a conformal metric on M.
278 7. Mobius Geometry Proof. As usual, we have the isomorphism φρ:Τχ(Μ) —» g/ί) for each ρ Ε Ρ situated above χ G M. Using a fixed choice, say qQ/^, of one of the quadratic forms given in Lemma 1.15, we can construct, on each open set U С Μ over which there is a section σ: U —» P, a metric defined by 0*(v) = 3Wf)(<Ar(*)(^)) for each v e T*(M)· Now two such sections differ by a change of gauge of the form (z * * \ 0 • • I : £7 -» Я, 0 0 z"1/ and as in the remarks following Lemma 1.17, the corresponding metrics differ by a factor of z2, where ζ = z(h) is as in the expression for h above. From this it is also clear that, by an appropriate change of gauge in this way, we can obtain any metric on U conformally equivalent to qx. Thus, we have a well-defined conformal metric on U. The only remaining point is to show that there is a globally defined metric on Μ restricting to the local conformal metrics we have described above. For this we use the notion of a partition of unity (cf. [W.M. Boothby, 1986], pp. 193-198). Choose sections aa:Ua —» P, where the Ua cover M, and let qa:T(Ua) —» R be the corresponding metrics. Since Μ is paracompact, we may find a locally finite refinement of the cover {Ua} and restrict the aas to it. Thus, without loss of generality, we may assume that {Ua} itself is locally finite. Then we may choose a partition of unity pa subordinate to {Ua} and set q — J2apaQa- Then q is a smooth metric on M, and on each Ua it is a finite sum of conformally equivalent local metrics of the type described above and is therefore conformally equivalent to any one of them. ■ Example 2.4. Consider the model geometry G/H, which by Lemma 1.3 we have identified with the unit sphere Sn С Rn+1. According to Lemma 1.18, the conformal metric constructed by the procedure of Lemma 2.3 is the class of the standard metric on Sn. ♦ Special Curvatures In Mobius geometry, the Cartan connection and its curvature take values in g, and so they may in general be written in block form as f1 ω=\θ Vo η ν a θ1 1 О
§2. Mobius and Weyl Geometries 279 (with blocks of size 1 χ Ι,ηχη and lxl down the main diagonal) and Έ Υ Ο Ω = Ι Θ Α Υ1 0 θ* -Ε/ de + υ Α θ άυ + ε/\ν + υ/\α 0 \ άθ + θΛε + αΛθ аа + вЛу + аЛа + уьЛвь * . (2.5) О • •/ We may say that the torsion free Mobius geometries that are "farthest" from the Riemannian geometries are those for which the curvature has no Riemannian part, that is, no rotational part. These are the geometries with θ = 0 and A — 0. However, the following result shows that nontrivial examples of such geometries can exist only in dimensions <3. Proposition 2.6. Suppose the curvature takes values in the ideal3 0q (c/. Exercise 1.13) so that it has the form (Ε Υ 0 Ω = Ι 0 0 Υ1 \0 0 -Εt Then (i) dim M > 3 => Ε = 0, (ii) dim M > 4 =ϊ Υ = 0. Proof. Since E is semibasic, we may write it as Ε = 2_^ fijOi Λ 0j. i<j Recall the Bianchi identity dft = [ft,ω] (Lemma 5.3.30). The (2,1) block of this identity is 0 = ΘΑΕ or O = 0kAE for к = 1,... ,n = dim M. Thus, for all fc, i<j For к fixed and к φ г < j φ к, the forms θ к Л θ ι Λ 0j are independent. Thus, fij = 0 whenever there is a & φ i,j. But for η > 3, such a & exists for every pair (г,^). Thus Ε = 0.
280 7. Mobius Geometry Now consider the (2, 2) block of the Bianchi identity. It reads 0 = Yl Αθι -ΘΑΥ or 0 = Y{ л 0j - θι A Yj for г J = 1,..., n. Thus, for all г, j, we have Yi A 0j Α θι — 0* A Yj Α 0* — 0. This means that the 3-form Yi Авг has a factor 6j for all j. For η > 4, this means that У*Λ0; = О for all i. Thus, we may write Yi — φ{ Α θ ι for some 1-forms φι. Putting this in 0 = Yi A 0j -θιΑ Yj yields фгАвгА 0j -OiA φά A 0j = 0 or {фг+Фэ)Г\в{Г\ва=й. From this we see that for any three distinct indices г, j, &, we have φι + φ^ is a linear combination of θι and 0j, 0j -f 0^ is a linear combination of 0j and ^^, 0i -f фк is a linear combination of 0* and 0fc. We may solve these equations for ф{. The solution shows that, for any three distinct indices г, j, k, the form φι is a linear combination of 0;, 0j, and 0^. For η > 4 and / any index different from г, this process will yield a linear expression for φι that is free of θι. This shows that φι can only be a multiple of 0;. But then У^ = φι Α θι = 0. ■ Normal Mobius Geometries Somewhat more subtle in their definition than the subclasses of geometries considered above are the normal geometries. Definition 2.7. A Mobius geometry is called normal if (i) it is of type s Θ q (i.e., Θ = 0 and Ε — 0), and (ii) the 5 component Ks: Ρ —» Hom(A2(g/i)),s) of curvature function takes values in the normal submodule described in Definition 1.21. * One of the consequences of the proof of Proposition 3.1 ahead is that, in the presence of condition (i), condition (ii) may be regarded as saying that the block ν of the form ω (see Eq. (2.5)) is determined in a certain way by the other blocks of ω. We shall need the explicit formula for Ks in terms of the block A in the curvature form Ω of Eq. (2.5). Let e* G g/f) and epq G в be as in Exercise 1.22. Since A is semibasic, we may express it as A. = у A-ijpqUi A Uj Qs) epq. i<j
§2. Mobius and Weyl Geometries 281 Lemma 2.8. Ks = £r<e>p<g Arsvqe% Λ e* 0 epq. Proof. Since K{u,v) — Ω(ω_1(ΐ/),α;_1(ΐ;)), for r < s we have Ks(er,es) = A(u~l(er),u~l(es)) = ^2 AiJPQ^i Λ 9j)(uj~1 (er),u~l(es)) ® epq i<j v<q = У ^ s\.ijpq\0ir0js 0jr0is)6pq = У^ s*-rSpq€pq. i<j p<q p<q Thus, Ks = ^К5(ег,е3)е1 Л e* = ^ Агзрде* Л е* ® evq. r<s r<s p<q Three-Dimensional Normal Mobius Geometry Proposition 2.9. Let (Ρ, ω) be a Mobius geometry on the three- dimensional manifold M. Then (Ρ,ω) is normals the curvature of (Ρ,ω) takes its values in q. Proof. =$> By Definition 2.7(i), the connection and curvature take the form and Ω To show that the curvature takes its values in q is to show that A = 0, which we do now. Since the space of the geometry has dimension 3, it follows that ^Л^з, ^Λ^ι, and θ\ Λ#2 constitute a basis for the semibasic 2-forms on P. Since curvature is semibasic, we may express A in the form A = (αχ * *)02 Л 0з + (*а2*)#з Λ θι -f (* * аз)#1 Λ 02, where a\ is the first column of (αϊ * *), and so forth. The Bianchi identity in the (2,1) block has the form Α /\θ — 0. Mutiplying this out yields (Μι Λ 02 Л 03,0,0) + (0, Μι Λ 02 Λ 03,Ο) + (0,0, Μι Λ 02 Λ 03) = 0, and hence αχ = α2 — α3 = 0. After a brief reflection, it follows that A must have the form 0 Μι Λ 02 Μι Λ 03 ' -Μι Λ 02 0 М2Л03 -Mi Л 03 -М2 Л 03 0
282 7. Mobius Geometry Thus, Ks = Мз Л е* ® e23 + 6263 Л е* 0 e3i + 63ei Ле^ ei2. Applying the Ricci homomorphism (cf. Exercise 1.22(i)), we get -Ricci(Ks) = 6i(e2 ® e\ + e3 0 e3) + 62(β2 0 e^ + e\ 0 e3) + 63(et 0ei-he2 0e;) = (b2 + &зК 0 ei + (63 + 6i)e2 0 e\ + (6X + 62)e3 0 e£. Since the geometry is normal, Ricci(Ks) = 0 and hence b\ = b2 = 63 = 0. Thus A = 0. ΦΦ· Since the curvature of (Ρ,ω) takes its values in q, condition 2.7(i) holds and A = 0. Thus, we need verify only 2.7(ii). But by Lemma 2.8, A = 0=^Ks=0=> Ricci(Ks) = 0. ■ Weyl Geometry In 1918, H. Weyl introduced a generalization of Riemannian geometry in an attempt to formulate a unified field theory (cf. [H. Weyl, 1952]). The fundamental idea was that, since there is no standard of length in the universe, the final length of a ruler moved along a path might well depend on the path. From this ingredient he constructed a theory unifying gravitation and electromagnetism. Although Einstein rejected Weyl's unreliable ruler as physically wrong, and the theory fell into disrepute, its mathematical structure has nevertheless resurfaced as a fundamental part of the "standard theory" of particle physics, where the unreliable length has been reinterpreted as an unreliable phase [S.-S. Chern, 1977]. We shall introduce Weyl's geometry, not as Weyl originally did,3 but as a particular case of a Cartan geometry. Definition 2.10. Let Μ be a smooth manifold. A Weyl geometry on Μ is a Cartan geometry modeled on the Weyl model. ft In a Weyl geometry, the Cartan connection and its curvature take values in gw? and so they may in general be written in block form as n ] and Ω = ( ~ . θα) \θ Α Theorem 2.11. A torsion free Mobius geometry (Ρ, ω) on the surface Μ with connection and curvature given by ε v\ v2 0 #i 0 —a v\ θ2 a 0 v2 0 0i θ2 -ε j ω=\θ: ° "α Ul | and Ω = θ2 a 0 v2 Ε 0 0 0 Υι 0 Λ21 0 Yi -A2i 0 0 0 Vi Yi Ε 3Cf. Definition 3.11 for a definition equivalent to Weyl's.
§2. Mobius and Weyl Geometries 283 for which Ε and A never vanish simultaneously, canonically determines a Weyl geometry (Pweyb ^Weyi) on Μ of curvature (de 0 0 " ^Weyi = I 0 0 —da 0 da 0 Proof. Every entry in the curvature form is a semibase 2-form. Since θχ Αθ2 is a basis for the semibasic 2-forms, it follows that each entry is a multiple of it, so we may write Ω = Ί о Vo 0 a 0 "2 —a 0 0 Ui "2 -el θχΑθ2 Now the element k = 1 Qi 42 r\ 0 1 0 qx 0 0 1 ,0 0 0 ΐ/ еЯ, where г =-(ς?+^), transforms the curvature according to the formula (2.12) r \ -Qi -42 1 / /e 0 0 Vo «1 0 α 0 Ui —a 0 0 °\ щ 1 ^2 -e/ <9ιΛ6>2 'e Ui—q2d + qie u2 + qia + q2e 0 0 -a 0 a 0 6>ιΛ6>2. ,0 0 0 The equations ^l - q2a + q\e u2 + q\a -\-q2e 0, 0 may be solved uniquely for qi and q2 provided that det -a e e a -(a2 + e2)^0,
284 7. Mobius Geometry that is, provided that Ε and A never vanish simultaneously. Since we are assuming this, it follows not only that in any fiber there are points for which У ι = Уч = 0, but also that changing such a point by right multiplication by any element к ф I of the form given in Eq. (2.12) will never preserve this property. On the other hand, elements of the form k = 0 0 0 \ cos b — sin b 0 sin b cos b 0 0 0 z~lJ always preserve this property. Set Pweyi = {p e Ρ | Yx = y2 = 0}. Then by Proposition 4.2.14, Pweyi is a reduction of Ρ to the Weyl group HN. The form (ε vi υ2 Ο \ 0i 0 -a vi \ θ2 ol 0 v2 I 0 6>x <92 -ε/ restricts to give a form on Pweyb but of course it need not happen that vi = v2 — 0. Nevertheless, the part of this form given by / ε 0 0 ^Weyi = I 0i 0 -a \θ2 a 0 on Pweyi determines a Weyl geometry on Μ with curvature Ω Weyl — G^Weyl + ^Weyl Λ CJ\Veyl = <te 0 0 \ (άε 0 0 d0i -h 0i Λ ε - α Α θ2 0 -da I = I 0 0 -da Kd02 + 02 Λ ε + α Λ 0Χ da 0 / \0 da 0 §3. Equivalence Problems for a Conformal Metric In this section we study two equivalence problems associated with conformal metrics. The first is the equivalence problem for a conformal metric itself. The solution here is a certain normal Mobius geometry. The second is the equivalence problem for a "Weyl structure," which is a conformal metric together with a certain compatible family of 1-forms. Here the solution is a Weyl geometry.
§3. Equivalence Problems for a Conformal Metric 285 The Equivalence Problem for a Conformal Metric As we saw in Lemma 2.3, a Mobius geometry on Mn uniquely determines a conformal metric on M. Here we study the converse and show that to a conformal metric Q on Mn, η > 3, there corresponds a unique normal Mobius geometry on Μ giving rise to Q by the process described in Lemma 2.3. (The case η = 2 is exceptional in this regard; cf. the discussion given in the course of the proof below of Proposition 3.1.) This result is an immediate consequence of the following proposition. Proposition 3.1. To each conformal metric Q on the open set U С Rn, η > 2, there corresponds a unique normal Mobius geometry on U whose associated conformal metric is the original one. Proof. We first find a Cartan gauge on U. Let q be a representative metric in the given conformal class of metrics on U. We may choose, with respect to g, an orthonormal basis of tangent vector fields еДх) along U. Let Qi. T(U) —» R, i = 1,..., n, be the dual 1-forms on U. Then <?(«) = 5>(V)2. Conversely, every decomposition of q as a sum of squares arises in this way. A gauge on U modeled on Mobius geometry may be written in block form as (ε ν Ο ω = Ι θ α ν1 \0 θ1 -ε We claim it is no restriction to consider only gauges satisfying ε = 0, which we shall assume in the sequel. To justify this, consider a change of gauge of the form h:U —> Я, where 1 r \rr1^ 0 1 r* ^0 0 1 This will effect the replacement ω —> Ad(h~l)u + Ь?ин, or ε θ 0 υ a θ' 0 υ1 —ε Since the 0$, г — 1,... ,n, is a basis for the 1-forms on [/, the form гв = Σϊ Ггвг is an arbitrary 1-form on U. Thus, we may take it to be equal to ε so that the (1,1) entry of the new gauge vanishes.
286 7. Mobius Geometry The curvature of our gauge (with ε = 0) has the form υ Αθ dv + v Λα 0' άω + ωΑω=\άθ + αΑθ da +α Αα + Θ Αν+ νι Αθ1 * 0 • * (3.2) where the components denoted by stars are determined by symmetry from other components already displayed. We aim to choose ω so that Τ and Ε vanish and so that "A has no Ricci curvature." According to Cartan's lemma 6.3.3, given forms 0*: T(U) —» R, г = 1,... ,n, there is a unique skew symmetric matrix-valued form a = (α^·): T(U) -» on(R) = Skewn(R) such that the torsion Τ = άθ-\-αΑθ = 0. Thus, the condition that Τ — 0 forces the choice of a. We may write ν = 0*/, where / = (Ζ^·) is some η χ η matrix-valued function on U. We claim that the symmetry of / is necessary and sufficient for the vanishing of E. In fact, Ε = νΑθ = θΗΑθ= Σ hjOiAOj l<i,j<n = у (lij — lji)0i Л 0j. l<i<j<n Thus, Ε = 0^1* = I. Finally, we ask if we can choose ν (i.e., /) so that Ks lies in the kernel of the Ricci homomorphism Ricci: Hom(A2(g/i)),s) —» (β/f))* ® (β/f))*. There are two parts to this. I. We show that no matter what υ is, Ricci(Ks) lies in the symmetric submodule 52(д/())* С (β/ί))*®(β/ί))*· (Since the 5 here corresponds to i) in Proposition 6.1.4, the symmetry of the values of the Ricci homomorphism follows from that discussion. However, for the convenience of the reader, we reprove this result in the present context.) II. For η > 3, there is a unique choice of / such that Ks G ker Ricci. Proof of I. Ricci(KB) lies in the symmetric submodule of (g/i))* 0 (&/Ъ)*· Since Τ = 0, the Bianchi identity corresponding to this block reads Α Α θ = 0, or J2S Aks Α θ3 = 0 for 1 < k,s < n. Evaluating this on ei A ej A e^ yields 0 = Σ Λι™ л °Лег Л ej А ек)
§3. Equivalence Problems for a Conformal Metric 287 = 2 22{Aks(ei,ej)6sk + Aks(ej,ek)8si + Aks(ek, ei)6sj} s = 2{0 + Аы(е0,ек) + ^fcj(efc, e»)}. Thus, i4fci(ej, ek) = Akj(ei, ek), which shows that Ks=^2 Apq(ei, ej)e\ Λ e* 0 epq i<j p<q satisfies the conditions of Exercise 1.22(iii), and so Ricci(Ks) = Σ Akj(ei,en)el 0 e* г J, к is symmetric. ■ Proof of II. Let us see first how the curvature block A depends on /. The structural Eq. (3.2) implies (where Epq is the matrix with 1 in position (p, q) and zeroes elsewhere) A - (da + a A a) = θ Α υ + υ1 Α θ1 = θα (β*/) + (ϊ*0) л α1 — / J Op A Ojljg 0 Epq -f- ^ ^ /jp02 Λ 0ς 0 -Epg = / ^ bipljqQi Α θj 0 i£p9 -f- ^ ^ bjqhpOi A 9j 0 -Ep9. ij,p,q i,j,p,q In the second sum we switch г <-> j, ρ <-> g and reorder the wedge product to get A — (da -\- aAa) = Y^ 6ipljq6i A 9j 0 2£pg i,j,p,q — У ^ 6ipljq9i A 6j 0 Eqp ij,p,q = /2 ^ipljqQi Л 9j 0 epq, i,j,p,q where epq = Epq — Eqp. Thus, we see that the terms in Ks involving / are /] bipljqe[ A e*j 0 epq = - ^ hqej л ei ® e«9 t,J,P,q i,j,q - Σ hq(e*j л e* ® ei9 -f e* Л е\ 0 e^·) 2 3,q
288 7. Mobius Geometry where we have used the symmetry of / and the definition of bij given in Exercise 1.22(iv). Both of these ingredients are used again to show that the terms in Ricci(Ks) involving / are "2 ^hqIHcci(bjq) = -- ^1оя{{п - 2)(e* 0 e* + e* 0 e*) j,q j,q + 26j^kel 0 e*k) = --{n - 2)Ej)9/J9(e* 0 e* + e* 0 e*) = -(n - 2) ^ /i9e* 0 e* - Trace(/)Efce^ 0 e*k It is easily seen that for η > 2 this is an arbitrary element of the symmetric submodule of (g/f))* 0 (β/f))* and that / is determined by this expression. Thus, in that case there is a unique value of / for which Ricci(Ks) = 0. On the other hand, when η — 2, the expression reduces to -TYace(/)^e*0e*, г in which case we do not get an arbitrary element of the symmetric sub- module of (g/t))* 0 (β/f))*, nor is / determined by this element. Let us review what we have done so far. We have described a procedure that may be represented schematically as in the following diagram. choice of representative Q ι ► q £$ (a conformal metric on U) choice of decomposition η ι > q{v) = Σ 0f(v)2 ι = 1 normal conformal gauge satisfying , , /£ = 0 1 > \θ = (θ, ft)' normal conformal ι > Cartan geometry on U It is clear from the construction that the conformal class of metric obtained from the resulting geometry is Q again. Moreover, the procedure we have
§3. Equivalence Problems for a Conformal Metric 289 described is completely determined except for the choices in the first two stages. Let us investigate the effect of varying these two choices. If θ{ is another coframe decomposing another choice q' G Q, then we may write θί — Y^hijej, where at each point (bij) = za~l, a positive scalar multiple of ζ of a rotation a~l. Let us set /z 0 0 \ h= 0 a 0 . \0 0 z~l I Then h: U —► Η is a gauge transformation and ώ = Ad(h~l)u + /г*о;я is also a normal gauge (since normality is a gauge-invariant notion). Moreover, θ = Αά(Ιι~ι)θ mod I), so ώ is the unique normal gauge associated to the decomposition of q' by θ as in the construction above. In particular, varying the two choices merely replaces the gauge ω by the equivalent gauge ώ and has no effect on the corresponding normal Mobius geometry. Exercise 3.3. In the derivation of step I of Proposition 3.1 it is assumed that e{ A ej Λ e^ φ 0, which requires η > 2. Verify that Aki(ej,ek) = Akj{ei, ek) even when η = 2. □ Corollary 3.4. Let (Ρ,ω) be a Mobius geometry on Mn, η > 2, of type 5 0q. Then there is a unique replacement for the block ν of ω (с/. Eq. (2.5)) so that the resulting geometry on Μ is normal. Proof. Let Q be the conformal structure on Μ determined by the given Mobius geometry. By Proposition 3.1 there is a unique geometry on Μ of type 5 0q with Ricci Ks in the normal submodule. Moreover, the construction of this geometry given in Proposition 3.1 shows that, except for the block υ, all the blocks of its Cartan connection are determined by Q and the condition that the geometry have type 5 0 q. It follows that, except for the block υ, the geometry (Ρ,ω) is the one constructed in Proposition 3.1 and that there is a unique choice of the block ν yielding a normal geometry. Corollary 3.5. Let U С Sn (n > 2) be any connected and simply connected open subset and let f:U —> Sn be any smooth immersion pulling back the canonical metric q^ on Sn to a metric f*qo conformally equivalent to the induced metric qo \U. Then there is an element g G G such that f(x) = gx for all χ G U. In particular, f is infective and extends to a conformal automorphism of the whole sphere.
290 7. Mobius Geometry Proof. Since η > 2, there are canonical normal Mobius geometries associated to the conformal metrics on U and Sn. By the theorem, / must be a local geometric isomorphism of these geometries. So, by Theorem 5.5.2, / is the restriction of the action by left translation of an element g G G. Ш Exercise 3.6. Show that this corollary is false if η = 2. [Hint: the Riemann mapping theorem says that any connected and simply connected proper open subset U С С is diffeomorphic to the open unit disc by a holomorphic and therefore conformal mapping.] Corollary 3.7. The group of conformal diffeomorphisms of Sn (n > 2) is the Mobius group G. Proof. By Lemma 1.12 we know that G is a subgroup of the conformal diffeomorphisms of Sn. By Corollary 3.5 we know that, for η > 2, any conformal diffeomorphism is an element of G. Ш This corollary also holds for η = 2, although the proof given is no longer valid. The Mobius Geometry Associated to a Riernannian Geometry In this subsection we shall identify the Weyl curvature, introduced in Table 6.2.5 for a Riernannian manifold Mn with metric q, with part of the curvature for the normal Mobius geometry associated to the conformal class of metric determined by q. This will show not only that the Weyl curvature of a Riernannian manifold is an invariant of the conformal class of the metric but that for η > 4 its vanishing is a necessary and sufficient condition for the metric to be locally conformally flat. More precisely, we have the following result. Theorem 3.8. Let Mn be a smooth manifold. If q is a Riernannian metric on M, let W(q) denote the Weyl curvature of the Riernannian manifold (M,q). (i) Ifqi and q2 are two Riernannian metrics on Μ differing by a smooth positive factor, then W(q\) = Wfa). (ii) If W(q) — 0 and η > 4, then on each sufficiently small open set U about any point in Μ, there is a flat Riernannian metric on U that differs from q by a smooth positive factor. Proof. This theorem is local, so we work with a gauge on an open set U С М. We may write the Riernannian gauge and (torsion free) curvature
§3. Equivalence Problems for a Conformal Metric 291 and 0 0 \ fO 0 ^Rie — dO + α Αθ da + a A a J \0 da + α Λα The corresponding normal Mobius gauge and curvature for the conformal metric determined by q according to the procedure in Proposition 3.1 are ^Mob υ Αθ dv + v A a 0' J^Mob = ( άθ + αΑθ άα + αΑα + θΑυ + υιΑθ* * О • • 0 dv + υ Α a 0' О άα + αΑα + θΑυ + ^Αθ1 * о о оу Note that we are taking Riemannian and Mobius gauges with the same choice of Θ. Since in each case the condition that the geometry be torsion free is the same (i.e., άθ + θΑα = 0), it follows from Cartan's Lemma 6.3.4, that the same a appears in both gauges. Moreover, since η > 3, Corollary 3.4 shows that the block ν is determined by the condition that the Mobius geometry is normal. We claim that the identity da + a A a = (da + a A a + θ Α ν + ν1 Α Θ1) - (θ Α υ + ν1 Α Θ1) is the decomposition of the Riemannian curvature into its Weyl curvature and its Ricci curvature (cf. Proposition 6.1.3(iii) and Table 6.2.5). To see this, note first that by the Definition 2.7 of a normal connection, the part Ks of the Mobius curvature function corresponding to da -\-aAa+θ Αν -\-νι Α Θ1 must lie in the Weyl submodule (the kernel of the Ricci homomorphism). Now consider the other term, θ Α υ + υ1 Α θ1, and write ν = θ11, where I — (kj) is symmetric. The part of the curvature function corresponding to this term is, as in the proof of II in Proposition 3.1, / j ^ipijq^i Λ ej 0 epq = — - 2_^ IjqVjq i,j,P,q j,q which, by Exercise 1.22(iv), lies in the Ricci submodule. In particular, W(q) is the curvature of the associated normal Mobius geometry and hence depends only on the conformal class of q, proving (i). On the other hand, when η > 4, Proposition 2.6 assures us that when W(q) = 0, then Ωμ(Λ = 0, so the Mobius geometry is flat. ■ The following exercise shows that even if W(q) — 0 and η > 4, q need not be globally conformally equivalent to a flat metric.
292 7. Mobius Geometry Exercise 3.9. Let Μ = Rn — {0} be equipped with the standard metric induced from Euclidean η space Rn. Define an action Ζ χ Μ —► Μ by the formula η · χ — 2nx. (i) Show that this is an action by conformal covering transformations so that Μ = Μ/Ζ acquires a conformal structure Q. (ii) Show that Μ is diffeomorphic to the product Sl χ Sn. (iii) Show that Μ is a locally flat conformal structure with holonomy group Z. (iv) Deduce that Μ is not complete. (v) Deduce that there is no global flat Riemannian metric on Μ in the conformal class Q. □ On the other hand, we have the following result. Proposition 3.10. Let Mn be a simply connected, complete, locally flat Mobius geometry. Then it is the model geometry. Proof. This is a special case of Theorem 5.5.3. ■ The Equivalence Problem for a Weyl Structure A Weyl structure is a refinement of the notion of a conformal metric. This is the structure actually used by Weyl (see [H. Weyl, 1952]) in his attempt to unify the forces of gravity and electromagnetism. We are going to show that a Weyl structure is the same as a torsion free Weyl geometry. Definition 3.11. A Weyl structure on Μ is a conformal metric Q on Μ together with a function F: Q —► Al(M) satisfying F(exq) = F(q) - d\ for every smooth function λ: Μ —► R. ft Lemma 3.12. A Weyl geometry on Μ determines a canonical Weyl structure on M. Proof. Let U С Μ be a small open set and let ω be an arbitrary gauge on U for the Weyl geometry on M. Write (ε 0 ω={θ α, П = ^ + уЛу=(й + Ш + аЛв da + ала Define q — J2i<j<n@] t0 ^e tne Riemannian metric on Μ associated to the gauge ω. Then we define F(q) = —2ε. To see that this is a Weyl
§3. Equivalence Problems for a Conformal Metric 293 structure on Μ, we need to see that every expression of the form exq (for any λ: Μ —► R) appears as the Riemannian metric on Μ associated to some gauge and that F(exq) — F(q) — d\. If ώ is another gauge on U for the given Weyl geometry, then it differs from a; by a unique change of gauge of the form i) h=iZ0 "):U^HN, and the gauges themselves are related by the formula ώ = Ad(h~l)u + h*UHN ζ 0\_1/ε 0\fz 0\ (z OV1 fdz 0 0 α) \θ α)\0 a) + \0 a) \0 da ε + ζ~ιάζ Ο \ ζα~ιθ a~laa + a~lda J ' (3.14) Moreover, any such choice of h will yield a valid change of gauge. It is clear from this that the Riemannian metric associated to ώ is z2q and that F(z2q) = -2ε - 2z~xdz = F(q) - 2z~ldz. Writing ex = z2 so that d\ = 2z~ldz then yields the result. ■ Now we solve the equivalence problem for Weyl structures. Proposition 3.14. Let F by a Weyl structure on the smooth manifold M. Then there is a unique torsion free Weyl geometry on Μ giving rise, as in Lemma 3.12, to this Weyl structure. Proof. Fix q G Q. On a sufficiently small open set U about any point of Μ, we may express q as q = ΣΘ2 for some coframe θ = {θ\, θ2, · ·., θη). We take for the gauge associated to this coframe the expression — i'-if»> ° в а where (as usual, by Cartan's Lemma 6.3.4) a is the unique form such that the torsion, άθ — \θΛF(g) + αΛθ, vanishes. Suppose we change our choice of q to q = z2q and the choice of coframe θ to θ = ζα~ιθ, where z:U —> R+ and a: U -+ On(R). Setting Λ = ( q °a]:U-+HN, we calculate Ad(h-1)w + h*wHN ζ oyV-2F(9) 0\fz 0\ (z Oy1 fdz 0 0 α V θ * \0 a \0 a
294 7. Mobius Geometry = (-\F{q)+z~ldz 0\ ^ ζα~ιθ •J ^ za"^ •у V 0 */" Since a gauge transformation will not alter the torsion, the vanishing of the torsion in the new gauge is automatic. Thus, the (2, 2) entry of this gauge is the correct one. In summary, we see that altering the choice of metric and coframe has no effect on the Weyl geometry. It merely replaces the given gauge by one equivalent to it. Thus, the torsion free Weyl geometry is uniquely determined by the Weyl structure. Moreover, it is clear from the discussion that the Weyl structure determined by this geometry according to Lemma 3.12 is just the original one. ■ §4. Submanifolds of Mobius Geometry- Suppose we are given a torsion free Mobius geometry ATn+r and an immersion /: Mn —► ATm+r. After establishing some algebraic notation, we explain how to obtain the "locally ambient geometry of N along /" (this notion is formalized in Definition 4.8). The main technical point is to find a certain reduction of the restriction to Μ of the ambient principal bundle to a principal bundle Η χ —► Ρχ —► Μ with group Η χ « Mobm(R)0 x Or(R). This is achieved by taking a subbundle of the tangent reduction along which a certain collection of traces vanishes. The locally ambient geometry of N along / is the bundle Ρχ together with the restriction to it of the Cartan connection for N. This is analogous to the Riemannian case in that this geometric entity contains, in principle, all of the geometry of the map /. In Proposition 4.17 we show that, when N is the Mobius sphere, this datum does indeed determine the immersion /: Μ —► N up to a Mobius transformation of N. Thus, in this case, the locally ambient geometry of N along Μ is a complete invariant for the map /. The next step in studying the locally ambient Mobius geometry on Μ is to see how it fractures into the tangential part (the "intrinsic" Mobius geometry on M, which is not necessarily normal), the normal4 part (the Ehresmann connection on the normal bundle u), and the "gluing" data (the second fundamental form). This decomposition depends on factoring the bundle Ρχ as a fiber product Ρχ « Ptan Хм Рпот- The tangential and normal parts of the locally ambient geometry on Μ are essentially given by the restrictions ω \ Ptan and ω \ Pn0r of the Cartan connection ω. These 4Note the two different uses of the word normal here. The meaning should always be clear from the context.
§4. Submanifolds of Mobius Geometry 295 pieces exist for any torsion free, locally ambient Mobius geometry on M. In Corollary 4.30, we show that when dim Μ φ 2 and the locally ambient geometry is normal, then it can be reconstructed from these pieces. A different approach is needed when Μ is a surface; this is discussed in §6 when codimension is one. Since the "intrinsic" Mobius geometry (Pt&n,u \ Ptan) on Μ is not necessarily normal, it is not the geometry on Μ corresponding to the conformal metric induced on it by the conformal metric associated (by Lemma 2.3) to the Mobius geometry on N. Nevertheless, the former determines the latter (cf. Proposition 4.20(iii)). This means that more data about the geometry of the immersion Μ —► N are packed into the "intrinsic" geometry of Μ than is possible with a normal geometry. The discussion of submanifolds given here parallels the modern treatment of the Riemannian case given in the last chapter. We remark that a more classical treatment, analogous to the classical treatment of the Riemannian case, may be found in [A. Fialkow, 1944]. See page 315ff for a brief look at Fialkow's method. Whereas the present treatment is quite general, Fialkow's treatment requires the absence of umbilic points. Model Algebra The models for the Mobius geometries on ATm+r and Mn involve the Klein pairs of Lie algebras (g, ()) and (a, b), respectively, where Q= \ * • * 0 • • 0 0 • • • • • • 0 0 • • • • 0 0 0 0 °\ * 1 * */ °\ • 1 0 ι */ b={ (* о 0 Vo f* 0 . 0 Vo * * * 0 * * 0 0 * * • 0 0 0 0 0 °\ • * */ °\ • 1 0 ι */ and g has blocks down the main diagonal of size lxl, η χ η, r χ r, and 1 χ 1. Of course, we must bear in mind that not all possible entries are allowed; a matrix in any of these Lie algebras must have the form (z и V Vo ρ a b Я -bl 0 Pl where a and с are skew-symmetric matrices. Let us set
296 7. Mobius Geometry 0(n) q = = < • о о о 0*00 0 0 0 0 0 0 0 0 /0 0 0 0\ * о о о 1 о о о о I \o * о o/ ( /0 * * 0\ 0 0 0*1 о о о * l\o о о о/ o(r) = /0 0 0 0\ ^ f о о о о I 0 0*0 Vo о о o/ /0 0 0 0\ 0 0 0 0 1 * о о о I \0 0 * 0/ Then we have the canonical з®о(гг)фо(г) module decomposition g/f) = t®u. The First Reduction Recall that we are studying an immersion /: Mn —> Nn+r, where N is equipped with a Mobius geometry (Ρ, ω). The description of the geometry of N localized along / is more complicated than in the case of Riemannian geometry, although it begins in the same way with the tangent reduction PT along /. This is the reduction of the pullback bundle f*(P) given by Pt = {(x,p) G /*(P) | ψρ{ί,{Τχ{Μ))) = φ С φ). The reduction PT sits in the commutative diagram Pr Μ ПР) id Μ / / ■* Ρ -> Ν The tangent reduction PT has group HT = {h e Η \ Ad(h~l)a/b = a/b}. An easy calculation (starting from the description of Η given in Lemma 1.8) shows that HT consists of all elements of Η having the form 1 ρ q s 0 1 0 pl 0 0 1 q] ,0001 ζ 0 0 0 \ 0 α 0 0 0 0 с 0 0 0 0 z~lJ where рь G Rn, <f <E Rn, 5 = \{ppl + qql) G R, ζ e R+, a e On(R), с G Or(R). The Lie algebra \)r of HT consists of elements of the form
§4. Submanifolds of Mobius Geometry 297 If we set ωτ = f*(u) | Pr, then we have the following. Lemma 4.1. f*(u)(T(PT)) = a mod i). Proof. а/Ъ = φρ{/*(Τπ(ρ)Μ)) = φρ(/*π+(ΤρΡτ)) = ωρ(/*(ΤρΡτ)) mod t) = f*(uj)(T(PT))moat). ■ Corollary 4.2. ωτ mod I) takes values in a/b С g/f) and Zience a;r /ms £fte /orra /ε ν ν 0 _ [ θ α -β1 υ1 Ur~ \0 β δ ν1 \0 6>έ 0 -ε, Т/ге Second Fundamental Form on the First Reduction To continue we need to find a reduction of Pr involving some of the "normal" information encoded in the second fundamental form. We note that this procedure has no analog in Riemannian geometry, where the tangential information alone is sufficient to describe all of the reduction needed. The appearance of an aspect of the second fundamental form for the very definition of the second bundle reduction requires that we introduce it right away, not in the final form in which it will eventually appear, but in its raw form as a collection of matrices. The form ωτ transforms under к G Hr according to R\uT Ad(k-l)uT = k' 0 Vo a β θ1 ν -β1 7 О °\ к, where Ί ρ q О 1 О О 0 1 ,0 0 0 with \{ppb + qqb). (4.3) For any I G l)r, we have ωτ(Χ^) = X and hence β(Χ^) — 0. Since the value of X^ at some point ρ G Pr runs through all the fiber directions as X varies in i)r, it follows that the form β vanishes on each fiber and is therefore semibasic. This implies that the zth row of β may be expressed
298 7. Mobius Geometry as β{ = θ*Ίτ(ϊ), where h(i), 1 < г < r, is an η χ η matrix-valued function on Pr. These h(i) constitute the raw version of the second fundamental form5 on M. Lemma 4.4. The vanishing of the (3,1) curvature block is equivalent to the symmetry of h(i) for all i. Proof. The vanishing of the (3,1) curvature block is equivalent to βΑθ = 0, or, for all г, 0 = β Λ θ = θιΗ{%) Αθ = Σ eph(i)pq Λ 0q p<q which is equivalent to the symmetry of h(i) for all г. Ш Since we are dealing here only with the case when the ambient geometry is torsion free, then in particular the (3,1) block of the curvature does vanish. Thus, the following expression, Eq. (4.5), is symmetric in v\ and V2'· b(p)(vi,v2) = β(ωρ1(ν1))ν2 = ··· \v2 fv\h(l)v2\ = \ ··· I, where v\,v2 G t. (4.5) \v\h(r)v2J To proceed further, we need to know how the h(i) transform under the action of к G HT. Lemma 4.6. For к as in Eq. (4.3), and for each j, Rlh(j) = z-'z-'a-1 | ^Су(Л(0 + qiI) U. Proof. Since /?г· = 0th(i), we first need to calculate how Θ and β on PT change under the action of к G Hr. Using the formula RkUJ = Ad(k~l)u with к as in Eq. (4.3), we easily calculate that Π*,β = ο-1(β + (ΐθί)α and Щ.0 = ζα'χθ. 5Of course, it is only in the Riemannian case that there is, properly speaking, a first fundamental form. It is the Riemannian metrics itself.
§4. Submanifolds of Mobius Geometry 299 On the one hand, RlP = c-1№ + qtet)a = c-1 On the other hand, R-Ιβ = R*k Ά+ίι*' 'вг(Ц1)+Я11)а' a = с Pr + qrOt j \ei(h{r)A-qrl)a) θ%(1)\ (Rl^RlhilY fh{r)J \ICketRrkh{r)i '{za-4YRlh{\)\ fPzaRlh{iy (za-^YRthir)) \etzaRkh(r)i Combining these yields, for each j, e^aRlhiJ) = jth row of c"1 \θί(Η{τ) + ςΓ1)α/ = Y^cijet(h(i) + qiI)a, г which, since θ has linearly independent entries, yields the result. ■ The Second Reduction P\ From Lemma 4.6 we see that, for any point pGPT, there is а к G Hr of the form /1 0 q s\ k = 0 10 0 0 0 1 ql Vo ο ο ι/ where s = -qql, such that, at pk G PT, all of the symmetric matrices h(j) have trace zero. (Just take qj = —trace h(j) for all j.) Moreover, given any point ρ G PT at which trace h(j)(p) = 0 for all j, then in the notation of Eq. (4.3), for к e #r, trace h(j)(pk) = 0 for all j & qj = 0 for all j. This, together with the smoothness of the function trace h(j), is sufficient (cf. Proposition 4.2.14) for the equations trace h(j) = 0, 1 < j < r, to give a further reduction of the principal bundle HT —» PT —» Μ to a principal bundle H\ —» P\ —» Μ, where #л = < /■ ^ /z 0 0 0 α 0 0 0c \0 0 0 ° \ 0 0 z-4 /1 ρ 0 s \ 0 1 0 p1' 0 0 1 0 , \0 0 0 1/ where 5 = -ppl > .
300 7. Mobius Geometry Note that Ηχ is isomorphic to a product of the two subgroups #л ^ Mobm(R)0 xOr(R) corresponding to the decomposition of group elements given by 9 = (z 0 0 0 0 α 0 0 0 0 с 0 V0 0 0 z~l ζ 0 0 0 0 α 0 0 0 0 10 \\θ 0 0 ζ- 1 0 0 0\ 0 10 0 0 0 с 0 о о о ι/ (where s = -рр1)- (4.7) It follows that the Lie algebra of Ηχ is given by f)\ = b0o(r) = /* 0 0 Vo • 0 0\ • 0 • 0*0 0 0* We set ωχ = /*(u;) | Ρ\. The pair (Ρχ,ωχ) is our candidate for the invariant describing the immersion / (cf. Lemma 4.14). Geometry of N Localized Along f The preceding discussion leads us to the following definition, which is justified in Lemma 4.14. Definition 4.8. Let Μ be a smooth manifold of dimension n. A locally ambient Mobius geometry on Μ of codimension r is a pair (Ρ,α;), where (i) Ρ is a principal Η χ = Mobn(R)0 x Or(R) bundle over Μ, (ii) ω is an (I) + a)-valued form on Ρ satisfying the conditions (a) ω:Τ(Ρ) —* ί) + α is injective at each point; ω-1(ί)) is tangent to the fiber at each point; and the composite of u; and the canonical Ηχ module projection ί) + α —» α Θ о (г) are surjective at each point. (b) R%u = Aa(k~l)u for all к е Hx. (c) cj(Xt) = χ for each X e ί)λ (= Ь Θ o(r)).
§4. Submanifolds of Mobius Geometry 301 Just as for the case of a submanifold, we may write ω = ν 0 ?' ν1 7 vb 0 -ε) (d) The zth row β(ϊ) of β has the form β{%) = θιΗ(ϊ), where h{i) is an nxn symmetric matrix of trace zero. (This is the condition that the ambient geometry be torsion free along Μ in the direction normal to M.) The curvature of (Ρ, ω) is άω -f ω Α ω = (de + ν Α θ dv + εΛν + νΛα+νΛβ dv + ε /\ ν — ν Λ β* + ν Λ j 0 άθ + θ/\ε + α/\θ άα-\-θ/\ν + α/\α-βί/\β + νί/\θί * * 0 άβ + βΛα+ΊΛβ + ^Λθ* d-y - β Λ β1 + 7 Λ 7* * 0 * 0 * (Ρ, ω) is called torsion free if άθ + ε Λ θ + α Λ θ = 0. (Ρ, ω) is called normal if it is torsion free, of type s Θ q, and Ks lies in the kernel of the Ricci homomorphism restricted to Hom(A2(a/b),s) С Hom(A2(g/i)),s). (Ks is the part of the curvature function corresponding to the (2,2), (2,3), (3,2), and (3,3) curvature blocks.) A gauge on U С Μ for a locally ambient geometry (P, cj) is a form σ*(ω) where σ: U —» Ρ is any section. Two local ambient geometries (Pi,cji) and (Рг,^) on Μ are said to be equivalent if there is an Mobn(R) χ Or(R) bundle map b: P\ —» P2 such that 6*(o;2) = cl>i. * We note that since cj(X^) = X for each X G ()a, it follows that ^(-X"*) = 0 for X G ()λ· Thus ν vanishes on all vectors tangent to the fiber of P, namely, ν is semibasic. We also note without proof that an effective locally ambient geometry may be described by its gauges just as for Cartan geometries (cf. Chapter 5, §2). Proposition 4.9. Let Mn be a manifold of dimension η > 0 and let ω — be a gauge for a locally ambient Mobius geometry of codimension r on M. Regard α, β, 7, and ε as fixed. Let us write the forms ν and ν in terms of the basis θ arising from the gauge as {v v) = θ11 where I is an η χ (n + r) matrix. (a) (i) If ηφ2 there is a unique choice of и such that Ricci(Ks) takes values in the span of e* 0 e^ for 1 < г, j < n. ε θ 0 0 ν a β θ1 ν -β1 Ί 0 0 υ1 ν1 —ε
302 7. Mobius Geometry (ii) Ifn = 2 then ν has no influence on Ricci(Ks). (b) (i) Ifn>2 there is a unique choice of υ such that Ricci(Ks) takes values in the span of e* 0 e!· for l<i<n,n+l<j<n + r. (ii) If η = 2 only the trace of ν {= l\\ + I22) has an influence on Ricci(Ks). It contributes the terms — (l\\ +l22)(e\ ®z\ -fe^e^). (iii) If η — 1 then υ has no influence on Ricci(Ks). Proof. We begin by calculating the contributions made by ν and ν to Ricci(Ks). From Definition 4.8(d) we see that the terms in the s block of the curvature matrix arising from υ and ν are θΑυ + υ'Αθ1 θΛν\ ίθ\Α , ., .t л (θ\ $)a« + «a(JP' 1<ς<η+Γ 1<ς<η+Γ = 2L, ^9^Ρ Λ @k{Epq — Eqp) l<p,k<n l<g<n+r = 2L/ 'fcg^pл QkCpq- l<p,k<n l<g<n+r It follows that the terms in Ks arising from υ and ν are - 51 'fcg4Ae£®epg- l<g<n+r Hence, by Exercise 1.22(i), in terms of Ricci(Ks) arising from υ and ν are - ^ lkqRicci(el Λ e* 0 eP9) l<p,fc<n l<g<n+r Σ Me* ® ej + ife,e; ® φ l<p,fc<n,l<i/<n+r l<p,fc<n>l<9<™+^ l<p,fc<n,l<i/<n+r ^фрфч.кфщ k^p^q,k=q Σ tkqSkqep ® e* l<p,k<n,l<q<n+r k^p^q )еЯ
§4. Submanifolds of Mobius Geometry 303 \<k<n,l<q<n+r l<k<n,l<q<n+r k^q k=q - Σ 1кквр0 ер l<p,k<n k-φν = -(n-2) ς Wfc®e;- Σ ^4®4 e9 l<fc<n l<fc<n l<g<n+r - Σ 'fcfcep <8) e* Η- Σ 1кке*к®е1 \<p,k<n l<k<n = -(n-2) ς Мь®е;- Σ /fcfc Σ ep0ep (*) l<fc<n l<fc<n l<p<n \<q<n+r The terms in the formula (*) involving e\ 0 e* for 1 < A; < η, η + 1 < g < η -h r, which is the same as the terms depending on v, are -fa-2) 5Z Mfc®^ (**) l<fc<n n+l<g<n+r Since this vanishes for η — 2, we verify (a)(ii). When η φ 2 it is clear from the expression (**) that any variation of ν will be reflected in a variation of the coefficients of e£®e* for 1 < к < η, η-hi < q < n + r and, moreover, any variation in these coefficients may be achieved by a suitable choice of v. This proves (a)(i). The terms in the formula (*) involving e\ 0 e* for 1 < k, q < n, which is the same as the terms depending on v, are ~(n -2) Σ hqel®e*q- ^T lkk ]T e^ej. (* * *) l<fc<n l<fc<n l<p<n \<q<n Note that the sum of the coefficients of the terms e£ 0 e£, 1 < A; < η in this expression is -2(n-l) ^ lkk. l<k<n Thus for η φ 1, the expression (* * *) determines the expression ~{n -2) Σ 'fc^fc®^ l<fc<n 1<9<η and for η φ 2 the converse holds. These remarks show that, for η > 2, any variation of υ will be reflected in a variation of the coefficients of e£ ® e* for 1 < A; < n, 1<<7<п and, moreover, any variation in these coefficients
304 7. Mobius Geometry may be achieved by a suitable choice of v. This proves (b)(i). If η = 2, the expression (* * *) reduces to -(hi + k2)(el <8> e* + e£ 0 ej) which verifies (b)(ii). Finally, if η = 1, the expression (* * *) vanishes and (b)(iii) is verified. ■ Exercise 4.10.* Prove the following analog of Proposition 2.6. Let (Ρ,ω) be a locally ambient Mobius geometry on Mn of type 3 Θ q. (i) If dim Μ > 3, then the 3 block of the curvature vanishes. (ii) If dim Μ > 4, then the curvature vanishes. □ The following result is an analog of Theorem 5.3.15. Proposition 4.11. Let (Ρ,ω) be a locally ambient Mobius geometry on Μ of codimension r and let χ G M. Then for each ρ Ε Ρ over χ, there is a unique linear isomorphism φρ:Τπ^Μ —» α/Ь such that the following diagram commutes. W) ^ И+» I , I TX(P) - ~ ν" ■> Φ c (Ь + α)/ϊ) Moreover, if h G Mo6m(R) χ Or(R), Йеп (/^ = Ad(/i_1)(£p. Proof. The existence, linearity, uniqueness, and injectivity of φρ follow from the fact that ωρ(ν) = 0 mod ϊ) <=ϊ υ is tangent to the fiber of Ρ —» Μ. Since dim Μ = dim α/Ь, it follows that </?p is an isomorphism. The final fact comes from applying Ad(/i)_1 to the equation φρ ο π*ρ = ωρ mod f) to get Αά(Η)~1φρ ο π*ρ = Ad(/i)_1u;p mod f) = Upfc о Дл# mod fj = ^ρΛ ° 7Γ*ρ/ι ° -R/i* = φΡΗ°π*ρ (since πο Rh = π). Since π*ρ is surjective, it follows that ψφ = Αά(Ιι~1)φρ. Ш
§4. Submanifolds of Mobius Geometry 305 Corollary 4.12. In the case of a locally ambient geometry arising from an immersion f:M —> N into a torsion free Mobius geometry, the map φρ:Τχ(Μ) —» a/b given by the theorem is the pullback, via /*, of the corresponding map φ?(ρ):^(^χ)(Ν) —» g/f) for the Mobius geometry on TV. Proof. This is simply the fact that, in terms of the form cj, in both cases the map φ is given by the block Θ. Ш Definition 4.13. Let (Ρ, ω) be a locally ambient Mobius geometry on Μ of codimension r. The normal bundle associated to (Ρ,ω) is the r-dimensional vector bundle unOT = Ρ χ#λ Rr, where the action of H\ = Mobm(R)o x Or(R) on Rr is the standard second factor action. qnOT:vnoT —» R is the canonical (up to constant scale) metric on ипот. Ф The point of Definitions 4.8 and 4.13 is that they describe the data associated to the immersion /: Μ —» TV we have been discussing. Lemma 4.14. The pair (Ρχ,ωχ) described on page 300 is a normal, locally ambient Mobius geometry on Μ of codimension r. Proof. Condition (i) of Definition 4.8 is obvious, as is the fact that ωχ is an (f) -f a)-valued form on Ρχ. Condition (ii)(b) is an automatic consequence of the formula on Ρ that R*hu = Ad(h~l)u for all h G H. Condition (ii)(c) comes from the definition of a Cartan geometry and the fact that the inclusion Ρχ С /* (Ρ) is equivariant with respect to the right actions of #λ; hence for X e (ц, the vector field X^ on /*(P) restricts to the vector field X^ on Ρχ. The injectivity of (ii)(a) comes from the fact that ωχ = /*(cj) | Ρχ, f is an immersion, and ω is injective. The rest of (ii)(a) comes from combining (ii)(c) with the fact (Lemma 4.1) that ωτ(Τρ(Ρτ)) = a mod i) for all ρ Ε Ρχ. Condition (ii)(d) follows from the nature of the first and second reductions. ■ We emphasize this example of a locally ambient geometry by giving the following definition. Definition 4.15. Let TV be a Mobius manifold and let /: Μ —» TV be an immersion. The locally ambient Mobius geometry of η along f (or along Μ if / is an inclusion) is the pair (Ρχ,ωχ). * Exercise 4.16. Let φ: Ν —» TV be a geometrical automorphism of a Mobius geometry on TV. If /: Μ —» TV is an immersion, show that the locally ambient Mobius geometry of TV along / is equivalent to the locally ambient conformal geometry of TV along 0/. □ Now we can justify the terminology "ambient Mobius geometry localized along /" by showing that, at least in the case when the ambient geometry
306 7. Mobius Geometry is the model Mobius sphere, it does indeed encapsulate all the geometry of /· Proposition 4.17. (i) Let f:Mn -> 5n+r (standard Mobius sphere) be an immersion. If (Ρχ,ωχ) is the ambient geometry localized along f, then dwT -f- ωτ Λ ωτ = 0 and the monodromy representation Φωτ is trivial. (ii) Conversely, let Mn be a connected η-dimensional manifold and let (Ρ,ω) be a locally ambient Mobius geometry on Μ of codimension r satisfying άω -f ω Α ω — 0 with trivial monodromy. Then there is an immersion f:Mn —» 5n+r? determined up to left multiplication with a Mobius motion o/Rn+r? such that the Mobius geometry of gn+r localized along f is equivalent to the geometry (Ρ,ω). Proof. This is just an application of the fundamental theorem of calculus, Theorem 3.7.14. ■ Decomposing a Locally Ambient Geometry Let (Ρ,ω) be a torsion free locally ambient Mobius geometry on Μ of codimension r. As in the Riemannian case studied in the last chapter, we wish to decompose (Ρ, ω) into three pieces. The first two of these pieces are the induced Mobius geometry on Μ and the Ehresmann connection on the normal bundle. These are described ahead in Propositions 4.18 and 4.20. The final piece is the second fundamental form described in Proposition 4.21 and Definition 4.22. As we mentioned before, although these three pieces exist in general, the locally ambient geometry may not in general be reconstructed from these pieces alone without the presence of some extra hypothesis such as that of normality (cf. Theorem 4.29). First we show that the bundle Ρ itself decomposes as the fiber product of principal bundles determined by the tangent bundle of Μ and the associated normal vector bundle vnOT. Proposition 4.18. Let (Ρ,ω) be a locally ambient Mobius geometry on Μ of codimension r and let Ptan = P/0(r) and Pnor = Р/МоЬь(R)o· (i) Ptan is a principal Mobn(H)o bundle and Pnor is a principal Or(R) bundle. (ii) Ρ = Ptan хм Pnor {fiber product). (iii) There is a canonical bundle equivalent Ptan xMobn(R)0 &/Ь, where the action of Mobn(R)o on a/b is induced from the adjoint action. (iv) Let vnor be the normal bundle over Μ associated to (Ρ, ω). Then Pnor is the principal bundle associated to vnor. Moreover, if (Ρ, ω) is the locally ambient geometry of an immersion f, then vnor is the normal bundle of f'.
§4. Submanifolds of Mobius Geometry 307 Proof, (i) Since the right actions of Mobn(R)o and of Or(R) on Ρ commute, we see that Mobn(R)o acts smoothly on Ptan· Since the action of Mobn(R)o x Or{H) is proper on Ρ and transitive and effective on the fibers of Ρ —» Μ, it follows that the action of Mobn(R)0 is proper ОП Ptan (cf. the proof of Lemma 4.3.12) and transitive and effective on the fibers of -Ptan —* M. Hence Ptan is a right principal Mobn(R)o bundle over M. The case of Pnor —* Μ is similar. (ii) The two smooth bundle maps Or(R) —» Ρ —» Ptan and Mobn(R)0 —» Ρ —* Ρηοτ (defined by projection) induce a smooth map into the fiber product Ρ —» Ptan Хм -Pnor· This map covers the identity on Μ and is obviously an Mobn(R)0 x Or(R) bundle isomorphism. (iii) Define Ρ χ α/b -» T(M) by (ρ, ν) -» (ρ"1 (ν) (cf. Proposition 4.11). Since every h G Or(R) acts trivially on α/b, we have φ~^{ν) = φ~ι{Αά{Η)ν) = φ~ι{ν). Thus, the map Ρ χ α/b -> T(M) induces a smooth bundle map Ptan x α/b — (P/On(R)) χ ci/b —> T'(Af). But we also have, for every к G Mobn(R)0, ψρ^(ν) — Ψρ1(^(^)ν)· ^ follows that the points (p,v), (pk,Ad(k~l)v) e Ptan x α/b have the same image in T(M). Thus, we get a further induced map Ptan Хмоьп(Я)0 α/β ~* T(M). This map is a vector bundle map covering the identity on M, and the fibers have the same dimensions. Thus, it is a bundle equivalence. (iv) This is similar to the proof in the Riemannian case (cf. 6.5.9(iii)) and we leave it to the reader. ■ Now we are in a position to study how the form a; of a locally ambient Mobius geometry (Ρ,ω) decomposes. Let us write ω — (ε 1 θ . 0 Vo υ a β et ν -β1 Ί 0 0 υ* ι/ -ε ^tan — (ε 'θ 0 \0 υ α 0 0* υ 0 0 0 υ υ1 0 —ε (4.19) The next step is to see that the parts u;tan and 7 of the form ω determine, and are determined by, certain induced forms (to which we give the same names) on Ptan and Pnor· Proposition 4.20. Let (Ρ, ω) be a locally ambient Mobius geometry on Μ of codimension r. (i) The form ωίαη {respectively, 7) is basic for the canonical projection Ρ —» Ptan {respectively, Pn0r)· We use the same name utan {respectively, unor) to denote the corresponding form on Ptan {respectively, ±nor)· (ii) Let 7 be the form on Pnor guaranteed by (i). Then 7 is an Ehresmann connection on the principal bundle Or(R) —» Pnor —» M. (iii) Let Utan be the form on Ptan guaranteed by (i). Then {Ptan^tan) is a Mobius geometry on Μ. Moreover, г/(Р, ω) arises from an immersion
308 7. Mobius Geometry f:M—>N, then the conformal metric on Μ associated to (Ptan^tan) is the conformal metric induced on Μ from N. (iv) (Ρ,ω) is torsion free <=> (Ptan^tan) is torsion free. Proof, (i) We deal with 7 only, the cases of utan being similar. The transformation law in Definition 4.8(ii) (b) implies that, for h G Mobn(R)0, ЩП — 7' and moreover, 4.8(H) (c) says that ω(Χ^) = X for X G f)A, so in particular j(X^) = 0 for X G a. Thus, by Lemma 1.5.25, 7 is basic for the principal bundle Ρ —» P/Mobn(R)0 = Pnor· (ii) The transformation law 4.8(ii) (b) implies that, for h G Or(R), #£7 = Ad(/i_1)7, and moreover 4.8(ii) (c) says that ω(Χ^) = X for X G i)A, so in particular, ί(Χ^) = X for X G o(r). Since the projection Ρ -» P/Mobn(R)0 is Or(R) equivalent and π*: ^1(P/Mobn(R)0, o(r)) -» Л*(Р, o(r)) is injective, it follows that the formulas #£7 = Ad(/i_1)7 for /i G Or(R) and 7(X^) = X for X G o(r) also hold for the version of 7 on Pnor· Thus, 7 satisfies the conditions for an Ehresmann connection (cf. Definition 6.2.4 of Appendix A) on Pnor· (iii) To see that utan is a trivialization of the tangent bundle of Ptan, since dim Ptan = dim 0, it suffices to show that uJtan-T(Ptan) —* 0 is surjective, or, equivalently, that the corresponding form utan:T(P) —» 0 is surjective. But since utan is just the canonical projection of u; on 0, by condition (ii)(a) of Definition 4.8, it is surjective. The proof that utan satisfies the conditions (a) R*hLutan = Ad(/i~1)o;tan for h G On(R), (b) utan(Xi)=XfovXeb is similar to the proof of (ii). Thus, utan is a Mobius geometry on M. If (Ρ,ω) arises from an immersion /, then the induced conformal metric qM-T(M) —» R on Μ arising from the Mobius geometry (Ptam^tan) is, using Corollary 4.12, given as follows: Qm(v) = \\φρ(ν)\\, where φρ arises from the geometry of Μ = ||ρ/(ρ)(/*(ν))||, where φ^ρ) arises from the locally ambient geometry of / = Qn(Mv))- Part (iv) is obvious. ■ The Second Fundamental Form The second fundamental form is the last of the three pieces of a locally ambient geometry. From one point of view, the second fundamental form
§4. Submanifolds of Mobius Geometry 309 on Μ is just the restriction to P\ of the trace zero matrices h(j) defined on PT. This may be rephrased by saying that the second fundamental form is a differential form on Μ with values in a certain vector bundle P\ xp (52(t) ® u). Equivalently, we may also say that it is a bilinear symmetric form В on T(M) of trace zero and with values in the normal bundle unor. All of this is described in the following result. Proposition 4.21. Let (Ρ,ω) be a locally ambient geometry on M, and let v-nor be the associated normal bundle. The block β of the form ω on Ρ given in Definition 4.8 determines, and is determined by, either of the following: (i) the function b G A°(P, p), where ρ = S2(Ad*) <g> Ad: #λ -» Gl(S2(C) <g> u)), given by b(p)(v\,v2) = β(ω~ι(νι))υ2, where v< G i for i = 1,2; (ii) The symmetric bilinear form В G Hom(S2(T(M)),vnor) given by φΡ(Βχ(υι,υ2)) = b{p){ipp(yx),ipp{v2)), where v< G TX(M) for г = 1,2. Proof, (i) β is a 1-form on Ρ taking values in Hom(t,u) (= t* 0 u). From the discussion on page 298, b(p)(v\,v2) — β(ω~ι(ν\))ν2 is symmetric in v\ and v2, so that b may be regarded as a function on Ρ with values in 52(t*) ®u. Now let us see how b transforms. Write g e Mobn(R)0 x Or(R) as in Eq. (4.7) so that, in particular, Ad(g)v\ — av\. Now β transforms according to Д*/? = c~lβα, and so 6 transforms according to b(pg)(vuv2) = β{ω~91(υι))υ2 = P(Rg*u>-1(Ad(g)vi))v2 = (Ιΐ;β)(ω-1(αν1))ν2 = ^~ιβα)(ω~λ(ανχ))ν2 = ο~ιβ(ω~1(αυι))αυ2 = c~xb(p)(avi,av2) = ((p(9-1)b)(p))(vuV2)· Thus, b G A°(P,p). Clearly, β determines and is determined by b. (ii) Note that b(p)((pp(v\),φρ(υ2)) G u. Since φ~χ()Χ) — vx = the normal space at χ G M, it follows that the formula defining Bx(v\,v2) makes sense (i.e., it takes values in vnor); neverthless, we must show that it is
310 7. Mobius Geometry independent of the choice of ρ lying over x. But this follows immediately from the transformation laws for b and for φ. Finally, we note that since φ ρ is an isomorphism for each p, it follows that В and b determine each other. ■ Definition 4.22. Let (Ρ,ω) be a locally ambient geometry on Μ of codi- mension r. The second fundamental form associated to (Ρ,ω) is the bilinear symmetric form В G Hom(S2 (T(M)), vnOT) given in Proposition 4.21(ii). Given a normal vector X G vx (x G M), the corresponding Weingarten map Lx G End(Tx(M)) is defined by (X,B(u,v))v = {u,Lx(y))M· * Definition 4.23. A bilnear symmetric form В G Hom(52(T(ra)), νηοτ) has trace zero if, for any orthogonal equinormal basis ei,e2,...,en of T(M), Σί B(e<, e<) = 0. An element b G A°(P, p), where ρ = 52(Ad*) 0 Ad: Hx -> G/(52(t*) 0u)), has irace zero if, for any othogonal equinormal basis ei, e2, ..., e„ of t, X)i b(p)(et, e<) = 0 (cf. Exercise 1.2.29). * Exercise 4.24. (a) Show that the two notions of "trace zero" given in Definition 4.23 agree if В and b are related as in Proposition 4.21(H). (b) Show that an element В е Hom(S2(T(M)),zynor) has trace zero if and only if the corresponding Weingarten maps Lx have trace zero for all X e vx. (c) Show that the second fundamental form given in Definition 4.22 has trace zero. □ Lemma 4.25. Lx is a self-adjoint map depending linearly on X. Proof. (u,Lx(v)) = (X,B(u,v)) = (X,B(v,u)) = (v,Lx(u)) = (Lx(u),v). Also, (u, Lax+bY (v)) = {aX + bY, B(u, v)) = a{X, В(щ ν)) + Ь(У, В(щ ν)) = a{u,Lx(v)) +b{u,LY(v)) = (u,aLx(v)+ bLY(v)), and so Lax+bY — aLx H- bLy. Ш Definition 4.26. Let TV be a torsion free Mobius geometry and let /: Μ —» TV be an immersion. A point χ e Μ is called umbilic for this immersion if the second fundamental form vanishes at χ. Φ Note that when Μ is a curve, every point is umbilic, so this notion is not important for curves. (See Definition 5.1 for the appropriate notion for curves.) Exercise 4.27. Let Μ С TV be a Riemannian inclusion. Then, by Proposition 3.1, we may regard TV as a Mobius geometry. If Вще and £?моь are the second fundamental forms in the corresponding geometries, show that #Mob — #Rie — ^Mob #Rie· Deduce that χ G Μ is umbilic in the Riemannian sense if and only if it is umbilic in the Mobius sense. □
§4. Submanifolds of Mobius Geometry 311 Reconstructing a Locally Ambient Geometry from Its Parts Consider the following collection of data: (a) a Mobius geometry on Μ of type s Θ q (the "intrinsic geometry" (Ptan,^tan)); (4.28) < (b) a vector bundle on Μ equipped with an Ehresmann connection (vnor and 7); (c) a trace zero element of Hom(52(T(M)), unOT) (the second fundamental form B). We have seen (Propositions 4.20 and 4.21(H)) how a locally ambient geometry (Ρ,ω) decomposes to determine data of this form. Now we study the converse. Theorem 4.29. Let Μ be a manifold of dimension η φ 2. If we are given the data of the form in (4.28), then there is a unique locally ambient geometry (P,u>) on Μ giving rise to it such that Ricci(Ks) takes values in the span o/e*®e*, l<i,j<n. Proof. Note first that by Proposition 4.18(ii) we must have Ρ = Ptan Хм Pnor, and since Pnor is the principal bundle associated to unor (Proposition 4.18(iv)), the data (a) and (b) determine P. We must now see that every block in the form 0 ω = is also determined. Pulling up the form Wta /ε ν 0 0 \ θ α 0 υ1 0 0 0 0 \θ θ1 0 -ε) of (a) to Ρ determines every block of ω except for 7, /?, and v. But 7 and β are determined by (b) and (c), respectively. Proposition 4.9(a) shows that ν is determined by the condition on Ricci(Ks). Ш Corollary 4.30. Let Μ be α manifold of dimension η φ 2. Then a normal locally ambient geometry (Р,ш) on Μ is determined by the data of the form (4.28) it gives use to. Proof. The final condition of Theorem 4.29 is automatic in the presence of normality. ■
312 7. Mobius Geometry Subrnanifolds of Mobius Spheres Submanifolds Mm of the flat space gn+r constitute an important special example of the general theory discussed above. We have already noted in Proposition 4.17 that the ambient geometry localized along / is a complete invariant for an immersion /: Mn —» gn+r up to a Mobius transformation of the ambient space 5n+r. Moreover, by Proposition 4.17, an arbitrary local ambient geometry arises from an immersion into the sphere if and only if its curvature and monodromy vanish. It is interesting to try to restate this result in terms of data of the form of (4.28). The immediate difficulty is that the monodromy condition doesn't seem to break up in this way. We avoid this difficulty by assuming that Μ is simply connected, so that the monodromy condition is automatic. Theorem 4.31. Let Μ be a simply connected manifold of dimension n>4, and assume that (a) (Ptan,utan) is an arbitrary Mobius geometry on Μ of type s 0 q (c/. Exercise 1.13(H)), (b) vn0r is an arbitrary r-dimensional vector bundle on Μ equipped with an Ehresmann connection 7, (c) В G Hom(S2(T(M)),vnor) is an arbitrary element of trace zero. Let β denote the Ηom(t,u)-valued form on Ρ corresponding to В via the formula of Proposition 4.21. Let ν be the form guaranteed by Proposition 4.9(a). Then (a), (b), and (c) arise from an immersion Μ —» gn+r if and only if the following equations hold: da + а A a + θ Λ υ + ν1 Λ θ1 = β1 Λ /?, d/? + /?Aa + 7A/? + i/A0*=O, (4.32) б?7 + 7 Λ 7* = β Λ β1 Proof. By Theorem 4.29, parts (a), (b), and (c) will always fit together to give a unique locally ambient geometry of codimension r of type s Θ q such that Ricci(Ks) takes values in the span of e* 0 e^, 1 < i,j < n. Referring to the curvature given in Definition 4.8(d), we see that the vanishing of the curvature implies Eq. (4.32). Conversely, in the presence of Eq. (4.32), the locally ambient geometry is normal and indeed of type q. Since η > 4, the curvature vanishes by Exercise 4.10(H). The monodromy also vanishes since Μ is simply connected. Thus (a), (b) and (c) arise from an immersion Μ -» 5n+r. ■
§4. Submanifolds of Mobius Geometry 313 Comparison of Riernannian and Conformal Surfaces in Space Suppose we are given a two-dimensional submanifold M2 С Rn+2 in Euclidean space. We may equip Rn+2 with the canonical normal Mobius geometry associated with its Riernannian structure and then take the induced Mobius geometry on M. We may also take the induced Riernannian structure on M. This yields two geometries on Μ with different models and different curvatures. Here we study the question of the relation between them. It is convenient to work with the gauge version of the geometries. Because of the existence of the tangent reduction in the Riernannian case, we may choose a gauge on Rn+2 such that the restriction to Μ of it and its curvature have the form ^Ri( ^Rie = 0 θ 0 0 a β 0 άθ + α 0 0 -β1 Ί Αθ 0 άα + αΛα-βιΛβ άβ + βΛα+^Λβ 0 * άΊ + Ί л 7 - β Λ /?* The Riernannian geometry induced on the surface has gauge and curvature given by ^Ric 0 0 θ a ^Rie \άθ + αΛθ da + αΛα) \0 β1 Α β /0 0 0 = 0 0 ΚθλΛθ2 \ , V 0 -Κθχ Αθ2 0 respectively, where K is the Gauss curvature. Now we pass to the Mobius case. Once again the existence of the tangent reduction (which is, of course, the same as in the Riernannian case) allows us to choose a gauge on Rn+2 such that the restriction to Μ of the gauge and its curvature are /° ''Mob 'Mob = -/3' °\ 0 β 7 v \0 Θ1 0 0 / 0 dv+νΛα+νΛβ άν-ν/Κβ*+ν/\η 0> άθ+α/\θ άα+θΛν+αΛα-β* Λβ+ν* Λθ* * 0 άβ+βΛα+^Λβ <*7-/ЗЛ/3*+7Л7 0 * 0 0, = ο,
314 7. Mobius Geometry where the 0, a, /3, 7 are the same as in the Riemannian case. But this is not yet the final reduction for a submanifold in conformal geometry (see the second reduction on page 302). The final reduction has the effect of making β trace free, replacing it with β0 = β — \ Trace /3, and hence replacing h(i) by ho(i) = h(i) — \ Trace h(i)I, while leaving the entries 0, a, and 7 unaltered. The entries υ and ν do change, to ν and ν say. Thus, the Mobius geometry induced on the surface itself has gauge and curvature given by /0 ν 0 ^7Mob = \ θ a vl \0 Θ1 0 0 άν -f ν A a SMob = \ άθ + α/\θ άα + θΑϋ + αΛα + ^Λθ1 0 • /0 Li0iA02 £20ιΛ02 0\ 0 0 —JRTo^i Л 02 • 0 #ο0ι Л 02 0 • \0 0 0 О/ -ϋΑβ0 о Definition 4.33. i^o is called the Willmore curvature6 of the conformal surface Μ. * Exercise 4.34. (a) Show that Kq: Μ —» R is independent of the choice of conformal gauge. (b) Show that at points where K0 doesn't vanish, there is no information contained in the functions L\,L<i. [Hint: show that there is a choice of conformal gauge for which L\ = L2 = 0.] (c) Show that at a point χ G Μ where Kq does vanish, the functions 1/1,1/2 determine a canonical covector ηχ e TX{M)*. □ Writing ft(z) = ί , 1, it follows that ft(i)0 Л 0* ft(i) = a b\ f 0 0i Λ 02 λ (a b b ο)[-θ1Αθ2 0 J \b с 0 1 -1 0 det(ft(i))0i Λ 02. We also have *»<»)- l ь J "I 0 Ha + c) -{ b -Ш-с) ' 6The analogy with the Gauss curvature suggests the opposite sign for Ко- However, the present convention ensures that Ко > О.
§4. Submanifolds of Mobius Geometry 315 so that det(MO) = -τ(α " c)2 " ^ (< 0) = --(a + c)2 + ac-b2 4 = - ( -Trace h(i) J + det h(i). Since /б*Л(1)\ β* Λβ = (Η(1)θ-··Η(η)θ)Λ it follows that ]Γ ft(t)0A0fft(t), 1<г<п \(?'Л(п)/ ^л^=(_°1 о) '9ι Λ ^2 Σ dethO), ^ / 1<Г*'<Гп 1<г<п and similarly, βί0Αβο=(^1 J)^iA^2 Σ det Μ*)· ^ ' 1<г<п Thus, the Gauss and Willmore curvatures of the surface may be expressed as κ=ς det(M*)), 1<г'<п Κο = - Σ ^(^W) (>0) 1<г<п = Σ J Qtrace Μ*')) " det(ft(i)) > = Я2 1<г<п I ^ ' J К This is conceptually interesting as it shows that the conformal curvature, which is "intrinsic" to the conformal structure, when looked at from the Riemannian point of view involves the mean curvature, which is "extrinsic" in that view. Willmore [T.J. Willmore, 1966] conjectures that for a torus T2 embedded in R3, the inequality fT2(H2 — K)dA > 2π2 holds. This conjecture is still open. (See [T.J. Willmore, 1993] for a report of recent work.) Fialkow's Relative Conformal Invariant Λ Aaron Fialkow gave a classification of submanifolds of a conformal manifold provided the submanifold is free of umbilic points ([A. Fialkow, 1944]).
316 7. Mobius Geometry Since a conformal manifold supports a unique normal Mobius geometry (Proposition 3.1), the present classification is certainly more general than his. The assumption that the submanifold is free of umbilics allows Fi- alkow to associate to each representative metric q on Μ a canonical positive smooth function Λ on Μ that scales the same way the representative metric scales. Then q/A is a confermally invariant Riemannian metric on M. The same remarks apply to the metric on the normal bundle, and one may consider a conformally invariant second fundamental form b/A2. In this way Fialkow converts conformal notions to Riemannian ones, making available the whole machinery of Riemannian geometry for the study of the conformal geometry of submanifolds. He gives a local classification in terms of symmetric tensors consisting of (i) the metric q/A, (ii) r additional tensors (some of which may vanish) of orders 4,6,..., 2r + 2, (iii) the "deviation tensor" of order 2 (necessary only for surfaces). The case of a surface in space is somewhat special in that (ii) consists of a single tensor, which is in fact the square of a tensor of order 2 so that three symmetric tensors of order 2 suffice for the local description of a hypersurface. For a comparison of Fialkow's classification and the present one in this case, see Exercises 6.6 and 6.7. Here we limit our discussion to Fialkow's function Λ. Let ei, e2, · · ·, er be an equinormal7 basis for the normal space ux and let Li, L2, · · ·, Lr be the corresponding Weingarten maps. Lemma 4.35. Let Xik, г = 1,... ,n, be the eigenvalues of Lk, 1 < к < r. Then A = Y^ijk(^ik — Xjk)2 is invariant under an orthonormal change of the basis e\, e2, · · ·, er of the normal space. Proof. Set V = TX(M) so that Lk: V -» V for each k. Then L\: V -» V and Lk 0 Lk'· V 0 V —» V 0 V. Since each of the Lk is diagonalizable, it is easily verified that Л = 2n Σ Trace(LJb) - 2 ]T Trace(Lfc 0 Lfc). l<fc<n l<fc<n Replacing the basis {e^} by the basis {ei = 5^7au'ej'}» wnere (aij) 1S an orthogonal matrix, will replace the corresponding Weingarten maps {Lk} ЬУ {^ = Y^j^jLj}. Then I.e., a basis whose elements are orthogonal and of equal lengths.
§4. Submanifolds of Mobius Geometry 317 Y^ Trace(Z^) = Y^ aijaijTr&ce(LjLk) l<k<n l<i,j,k<n = Σ SjkTmceiLjLk) = ^ Trace(L^), l<i,j,k<n l<k<n and similarly, У^ Trace(Zfc ®Zfc) = ^ akpakqTr8Lce(Lp ® Lq) 1<к<п 1<&>Р>9<™ = 5Z ^P9TraCe(^P ® Lq) l<p,q<n = Σ Trace(Lfc0Lfc). l<fc<n ■ Corollary 4.36. χ Ε Μ is umbilic <^> A(x) = 0. Proof. First note that if Lx is a multiple of the identity for each X in a basis of the normal space vx, then by Lemma 4.25, Lx is a multiple of the identity for every X G vx. Thus, we have A(x) = 0&Xik = Xjk for 1 < г J, k<r <=> Lk is a multiple of the identity for 1 < к < г <& χ is umbilic. щ Now let us reinterpret Л as a function on P\. For this we fix, once and for all, an equinormal basis {es} for u. Then for each ρ G P\, {^^(es)} is an equinormal basis for vx, and we may write A(p) for Λ with respect to this basis. The effect of replacing ρ by ph, where /μ 00 0 \ /1 ρ 0 s\ 0 α 0 0 0 0 d 0 \0 0 0 μ-1/ 0 1 0 pl 0 0 10 ,0 0 0 1 / G#x, where s = ψ ' p\ is to replace φρ l(es) by <р^(ёа) = (Ad(h-l)<pp)-l(Ba) = φ;1(Αά(Η)β3) The orthonormal change d of the basis {<£p(es)} has no effect on Λ, while the scaling μ-1 multiplies Λ by μ-1. Thus, up to a universal choice (the choice of the equinormal basis {es} for u), Λ determines a well-defined function on Ρ transforming according to A(ph) = μ~χΑ(ρ).
318 7. Mobius Geometry In the case where Μ has no umbilics, so that Λ is never zero, we obtain a further reduction of the bundle P\ by setting Ρμ = {ρ e Ρλ Ι Λ(ρ) = 1}. Exercise 4.37. Suppose that Μ С N С Q are three conformal manifolds with Μ and TV receiving their geometries from Q. Assume that Μ is totally umbilic in TV and TV is totally umbilic in Q. Show that Μ is totally umbilic inQ. □ §5. Immersed Curves Let us assume we are given an abstract locally ambient geometry on a curve. We study this geometry in the case of a curve of codimension one or two that is without vertices. These hypotheses allow a reduction of the locally ambient geometry to a principal bundle with discrete fiber. In this case the locally ambient geometry on Μ may be fully described by certain canonical functions and forms on Μ itself.8 Definition 5.1. Let Μ be a connected manifold of dimension one, and let (Ρ, ω) be a locally ambient Mobius geometry of codimension r on M. A point χ G Μ is a vertex for this geometry if the block ν of ω vanishes (see Definition 4.8). * Recall that by Lemma 4.14 an immersion of a submanifold Μ into a Mobius geometry determines a locally ambient geometry on M. It follows that Definition 5.1 defines the notion of a vertex for any immersion of a one-dimensional manifold into a Mobius geometry. Exercise 5.2. Show that for curves in the plane (regarded as a Mobius geometry) a vertex is just an inflection point. □ Proposition 5.3. Let Μ be a connected manifold of dimension one, and let (P,u) be a vertex-free locally ambient Mobius geometry of codimension one on M. Then (Ρ,ω) is determined (up to equivalence) by a canonical function κ\: Μ —» R and a canonical (up to sign) 1-form ±θ G Al(M). Proof. We refer to Definition 4.8, where the locally ambient geometry has the connection form Actually, it would be more proper to say that these forms and functions take their values in certain flat line bundles over Μ according to the principles described in footnote 12 of Chapter 1. However, the situation is so elementary here that we confine ourselves to the brief Remark 5.4.
§5. Immersed Curves 319 ω — 'ε υ ν Ο θ Ο Ο υ Ο Ο Ο ν ,0 θ Ο -ε, in which the blocks a and 7 vanish because they are lxl skew-symmetric matrices and the block β vanishes because its ith row β(ϊ) is 6h(i) and h(i) is a 1 x 1 matrix of trace zero (cf. 4.8(d)). Since ν is semibasic, we have у — f$ for some function / on P. The vertex-free condition means that ν never vanishes and so neither does /. We are going to find two successive reductions of the bundle P. These reductions will be defined by ν = θ and {ν = θ, ε = 0}, respectively. The initial group (i.e., of P) is Ял=МоЬ!(К)0х{±1}. The First Reduction, ω transforms according to R*lj = Ad(g )ω. Let us write g G H\ as 9 = Then ν transforms according to Щи — α2μ lv and θ transforms according to R*0 = αλμθ. Thus, ν = /Θ =» RTgu = R*gfR*ge =ϊ α2μ~1ν = (Ι1*9/)αιμθ =>α2μ-1/θ=(Ρ*9/)αιμθ =» R*gf = αια2μ~2ί. Hence / assumes the value 1 on every fiber. It follows that we may reduce the bundle to a subbundle P\ C. Ρ defined by / = 1 (or ν = θ) which has group 1 0 0 0 0 a 0 0 0 0 a 0 °\ 0 0 1/ /! 0 0 Vo Ρ 1 0 0 0 0 1 0 8\ Ρ 0 , 1/ :Ρ2,α ±1 On this reduction the form ε is semibasic, so we may write ε — /θ for some (new) function on the reduced bundle. The Second Reduction. The final reduction depends how ε transforms under the element
320 7. Mobius Geometry 0 1 0 pM 0 0 1 0 1 0 0 0 1/ 1 2 A simple calculation shows that Ще = ε — ρθ. By choosing ρ = /, we see that there is a point on each fiber where ε = 0. Thus, there is a final reduction P2 С Pi С Ρ via the equation ε = 0 to the discrete group Z2 generated by w = diag(l, —1,-1,1). On this reduction, the form restricts to 0 θ 0 0 υ 0 0 θ θ 0 0 0 0 υ θ 0 ω and this time ν is semibasic. Thus, we may write ν = κ\θ for some function K\ on P2. Now an easy calculation shows that ν and θ change sign under the action of w. Thus, κ\ is constant on the fibers of P2 and hence is basic. On the other hand, θ is defined up to sign on Μ, so we may express the induced form on Μ as ±θ. Clearly, this function κ\ and form ±θ constitute a complete set of invariants for the immersion. ■ Remark 5.4. Note that prescribing an orientation for Μ would allow a canonical choice among the forms ±θ (if e\ G TX(M) is positively oriented, choose the sign so that θ{β\) > 0). We remark that θ G AX(P2, R) may also be regarded as an element θ Ε Α1 (Μ, С), where ζ is the (trivial) line bundle associated to P2 by the canonical representation of Z2 —» Οχ (R) sending w ι-* — 1, where it; = diag(l, —1,-1,1,1) as above. Similar remarks apply to the θ and κ2 appearing in Theorem 5.6. Definition 5.5. ±θ is called the conformal arclength of Μ, and κ\ is called the conformal curvature of Μ. Й It may be shown (cf., e.g., [G. Cairns and R. Sharpe, 1990]) that for a curve in the (x, г/)-р1апе with Taylor expansion at the origin of the form У ~ \χ2 + ^Ακ^χΑ + 0(x5), then, at the origin, θ = dx and K4 is the conformal curvature. Now we turn to the case of a vertex-free curve in an ambient geometry of codimension two. Theorem 5.6. Let Μ be a connected manifold of dimension one, and let (Ρ,α;) be a vertex-free, locally ambient Mobius geometry of codimension two on M. Then (Ρ,ω) is determined (up to equivalence) by a canonical function κ\\ Μ —» H, a canonical (up to sign) function ±κ,2'· Μ —» R? and a canonical (up to sign) 1-form ±θ G Al(M). Proof. Referring to Definition 4.8, the locally ambient geometry has the form
§5. Immersed Curves 321 ε θ 0 0 ν 0 0 θ ν 0 7 0 0 ν ν —ί where the block a vanishes because it is a 1 χ 1 skew-symmetric matrix, and the block β vanishes because its zth row β(ι) is 0h(i) and h(i) is a 1 x 1 matrix of trace zero. Since ν is semibasic, we may write ν — /θ for some function f:P —> R2. Since ν never vanishes (cf. Definition 5.1), neither does the function /. We are going to find two successive reductions of the bundle Ρ defined by [v — θβχ] and {v = #ei, ε = 0}, respectively. The initial group of Ρ is #A = M5bi(R)0xO2(R). The First Reduction, ω transforms according to Щи — Ad(^_1)a;. Let us write 9 = where ζ > 0, a\ G {=Ь1}, с G 02(R), s — \p2. Then ν and θ transform according to R*v = z~xvc and R*9 = α\ζθ. Thus, v = fe=*irgV = Щ!Щв ^z-lvc = ax{R*gf)z6 ^z-1fce = al(R*gf)ze ^R*gf = aiz-2fc. Since / never vanishes, it follows that it assumes the value e* G R2 on every fiber. Thus, we may reduce the bundle to a subbundle P\ С Р defined by f = t\ (or ν = θβι), which has group ζ 0 0 0 0 αϊ 0 0 0 0 с 0 ° > 0 0 z-Ч /! ° 0 Vo Ρ 1 0 0 0 0 1 0 8\ ρ ι 0 , 1/ ί1 о о о Vo о αϊ 0 0 0 0 0 ах 0 0 0 0 0 αϊ 0 °\ 0 о о 1/ Ζ1 о о о Vo о о 1 о о Ρ о о 1/ On this reduction, the form ω restricts to ω (ε υ θ (9 0 0 0 0 0 0 0 7 Vo θ о 0 0 -7 0 0 оьа2 6 {±1}, s= -ρ2 г; θ 0 -ε J in which, except for v, all the forms are semibasic. In particular, we may write ε — /θ for some (new) function / on the reduced bundle.
322 7. Mobius Geometry The Second Reduction. The final reduction depends on how ε transforms under the element /! 0 0 Vo Ρ 1 0 0 0 0 h 0 s Ρ 0 1 Β=ψ A simple calculation shows that R*e — ε — ρθ. By choosing ρ — /, we see that there is a point on each fiber where ε = 0. Thus, there is a final reduction P2 С P\ С Ρ via the equation ε — 0 to the group {±1} χ {±1} = ^ 0 = /1 0 0 0 0\ 0 αϊ 0 0 0 0 0 αϊ 0 0 0 0 0 α2 0 \0 0 0 0 l/ аьа2 G {±1} (5.7) On this reduction, the form restricts to /0 г; θ 0 <9 0 0 0 0 0 0 -7 θ 0 0 7 0 0 \0 0 0 0 0 ν I and this time every entry is semibasic. Thus, we may write ν = κ\θ and 7 = /ς2# for some functions tti, к,2: Рч —» R. An easy calculation shows that for g as in Eq. (5.7) we have #*<9 = αι<9, R*gv = aiv, Я*7 = aia27, so that #*tti = «i, Я*/^2 = α2κ;2. Thus, we may regard κ\ as a function on Μ, κ:2 as a function on Μ defined up to sign, and θ as a 1-form on Μ defined up to sign. The functions κ\, κ2 and the form θ obviously constitute a complete set of invariants for (Ρ,ω). Ш It may be shown (cf., e.g., [G. Cairns, R. Sharpe, and L. Webb, 1994]) that, up to confermal transformation, every vertex-free curve in 3 space has a Taylor expansion at the origin of the form 1 3 1 ,rt Ζ=^Κ2Χ4+0(Χ5) κ1)χ5+0(χ6), and that, in this coordinate system, θ = dx at the origin and κ;ι, κ2 are the values at the origin of the corresponding functions described in Theorem 5.6. Exercise 5.8. Generalize Theorems 5.3 and 5.6 to the case of arbitrary codimension. □
§6. Immersed Surfaces 323 §6. Immersed Surfaces For a surface Μ, the classification of immersions given in §4 fails in that a normal locally ambient geometry on Μ cannot be reconstructed from the intrinsic geometry on Μ, the Ehresmann connection on the normal bundle, and the second fundamental form (cf. Theorem 4.29). In this section we provide, for surfaces in R3, a partial alternative of the theory of immersions given in §4, which is analogous to our study for curves in §5. We study in some detail the case of a flat locally ambient geometry on a surface under the restriction that the geometry is without umbilic points and both the ambient and induced geometries are oriented. The absence of umbilic points allows us to reduce the principal bundle of the local ambient geometry to a bundle with discrete fiber. This principal bundle may be described as the bundle of "double orientations" of Μ, where a double orientation is an orientation for each of the two principal subspaces which are compatible with the orientation of the surface at that point.9 The local ambient geometry on Μ may then be fully described by certain canonical functions and forms on Μ itself. All but one of these forms and functions have values in a certain real flat line bundle L over the surface. L is the line bundle described in footnote 12 of Chapter 1. Proposition 6.1. Let Μ be a connected, topologically oriented manifold of dimension two, and let (Ρ,ω) be an umbilic-free, topologically oriented, flat, locally ambient Mobius geometry of codimension one on Μ (e.g., one arising from an immersion Μ —» 53). Let us write the form as Iе 01 #2 0 Vo Vi 0 a βχ θχ V2 —a 0 fo 02 V βχ βΐ 0 0 °\ Vi V2 V . -ε) Then, canonically associated to (Ρ,ω) are (i) a two-fold cover Μ —> Μ with associated flat line bundle L over Μ, (ii) a function Φ: Μ —» R, (iii) sectionspi,p2 G A°(M,L) (= A£2(M,(R,-)), where (R, -) denotes the nontrivial representation of z2, (iv) L-valued 1-forms θχ,θ2 e Al(M,L)(= Л^2(М, (R,-)). Let ξ\,^2 be the vector fields dual to θ\,θ<ι. Then these data satisfy the relation 9Since the principal directions have distinct eigenvalues, they may be canonically ordered, so "compatible" makes sense here.
324 7. Mobius Geometry ω (ν) άώ + ώ Λ ώ = 0, w/zere [\ρ\Θ\-\ρίΘί • • Ι 0ι 0 0 — θ2 0 0 0 б! -02 V ο • • • • * 0 0 0 \ -^((1 + Φ)0ι + (ξιΡ2+ΡιΡ2)02) * 5(«2Ρΐ-ΡΐΡ2)βΐ-(1-Φ)02) -^(Ρ101+Ρ202) ги/геге Ζ2 ac£s 6г/ ί/ге covering transformation on Μ, and by Ad(r), τ = diag(l, —1,-1,1,1) on q. (As usual, the entries denoted by stars are determined by symmetry.) Moreover, given data as in (i), (ii), (iii) and (iv) satisfying relation (v), there is, up to equivalence, a unique locally ambient geometry giving rise to it Proof. As usual, we let Η χ be the group of the locally ambient Mobius geometry. We claim the absence of umbilic points means that it is possible to obtain a reduction of the principle bundle P. This reduction is given by ■{ P e Px I h[p) = 0 -1 Q = ui, where β = 0*h. Let us see that this reduction exists. According to Definition 4.8(b), we have jR£cu = ka(k~x)u for all к G Яд, so calculating as in Lemma 4.6, we get R*kh = (z~1cl)a~lha, where k = where r = \uul, ζ > 0, a G О2(Г0> с= ±1. The absence of umbilic points means that h is a nonsingular matrix. Of course, h is also symmetric of trace zero. Thus, we may choose a rotation a and a scaling z~lc such that l Д), proving the existence of the reduction. Upon restriction to Pp, the form ω becomes 1 0 0 0 и h 0 0 0 0 1 0 r\ ul 0 1/ (z fo 0 Vo 0 a 0 0 0 0 с 0 0 0 0 ζ'1 (z lcl)a lha (ε θι θι 0 Vo V\ 0 a θχ θι V2 —a 0 -02 ^2 ι/ -01 02 0 0 v2 V -ε) A simple calculation shows that the stabilizer of subgroup of Η ρ generated by elements of the form 1 0 in Яд is the
§6. Immersed Surfaces 325 9\ (\ 0 0 0 0\ 0-1000 0 0 10 0 0 0 0 10 Vo ο ο ο ι/ и — 0 0 0 Vo ux 1 0 0 0 u2 0 1 0 0 92 (l 0 0 0 Vo 0 0 0 0 0 10 0 -10 0 0 0 0-10 0 0 0 l/ 3(«1+«2)\ u2 0 1 (6.2) / so that Pp is a principal Hp bundle. The restrictions of the forms ε and a to Pp vanish on the fibers, and thus they are semibasic. Another calculation shows that λά{υ)ω /• • • • *\ * • • • • ' * a — 112Θ1 + ιΐιθ2 * * • * • • • • It follows that on each fiber of Pp there is a point where a = 0, so we get a further reduction to the finite group Z2 Θ Z4, generated by g\ and g2l by setting 1 0 Μ = {pePP\h(p)=(l _°1),α = θ}. It follows that Μ —> Μ is an eight-sheeted principal covering space. Upon V\ 0 0 01 01 f2 0 0 -02 02 1/ -01 02 0 0 0 \ V\ ' f2 ^ 1 -ε/ restriction to Pp, the form ω becomes Iе 01 ω — 02 Is Because Μ —> Μ has discrete fiber, all of the entries in this matrix restrict to zero on the fiber and hence are semibasic. Moreover, the structural equation implies certain relations among these forms, which we now determine. The (2,3) component of the structural equation reads 0i Λ υ2 + 0ι Λ 02 + υι Λ 02 = 0, so that if we express v\ and v2 in terms of the basis θ\ and 02, they must have the form Ы =--(1 + Ф)0! --П02, 1 v2 -<Ζ201--(1-Φ)02
326 7. Mobius Geometry for some functions n, q2, and Φ on M. On the other hand, the structural equation may be used to compute dOi in two different ways (from the first column and from the fourth row). We get dOx = -0i Λ ε = θχ Α ι/, dQ2 = -θ2Αε = -02 Λ ν, so that if we write ε = \ρ\θ\ — \p2Q2 for some functions p\ ,p2 on M, then we must have 111 1 ν = 2Ριθι + 2P2°21 d°l = 2Ρ2θΐ Λ 6>2' and d°2 = 2Ρχθχ Α °2' Putting these expressions in the (1,1) block of the structural equation, we calculate de — —v\ Α θ\ — υ2 Α θ2 or d ί^ρχθχ - ^Ρ2θλ = 1((1 + Φ)0ι + Πθ2) Λ 01 - ^2ΘΧ - (1 - Φ)62) Λ 02 = -^Γι0ιΛ02-^2^ιΛ02. (6.3) Substitution in the (1,4) entry of the structural equation, which is dv — — ε Α ν + v\ Α θχ — υ2 Λ 02, yields d ί 2^ι6>1 + 2^2<92) =~\2Ριθι~ 2Ρ2θ2) Λ ( 2Pl6>1 + Ψ2®2 -^((1 + Ψ)0ι+η02)Λ0ι -^(926ι-(1-Φ)62)Λ02 1 11 = -ψ\Ρ2θ\ Α 02 + -7Ί01 Λ 02 - -9201 Λ 02. (6.4) Adding and subtracting Eqs. (6.3) and (6.4) yields d{p\ A 0i) = - ί q2 + -P1P2 1 01 Λ 02 and d(P2 Λ 02) = Ι η - -ριρ2 J 0i Λ 02,
§6. Immersed Surfaces 327 which in turn imply that dpi Λ 0i +ρχάθι = -(q2 + -ρχρ2 J θχ Α θ2 and dp2 Λ θ2 +ρ2άθ2 = ί η - -Ρ\ρ2 J θχ Λ 02, or dpi Λ <7ι<9ι + (q2 -\-ρ\ρ2)θ2 and dp2 Λ θ2 — [τχ - Ρχρ2)θχ Λ θ2. Thus, we may write dpx =ςχθχ+ (q2 -\-ρχρ2)θ2 and dp2 = (n - ρχρ2)θχ + r2<92 for some functions qx and r2. Note that from this equation it follows that, if ξχ and ξ2 are the vector fields dual to θχ and #2, then ξιΡι = Qu bPi = 42 +P1P2, ξιΡ2 = rx- ρχρ2 and ξ2ρ2 = r2. From the equation R*u = Аа(д~1)и, we obtain the following table showing how the various items transform. V R9lV R92V θχ -θχ -θ2 θ2 θ2 θχ Ρι -Ρι Ρ2 Ρ2 Ρ2 -Ρι φ φ -φ It follows that the 2-form θχ Αθ2 is invariant under g2 but changes sign under gx. If there were a path on Μ joining ρ Ε Μ to pgx, this path would cover an orientation-reversing loop on M. By hypothesis, no such loop exists. Thus, Μ = Mf U Μ1 gx for some submanifold (a union of components) of M' С Μ, and Μ' —» Μ is a four-fold cover (a principal Z4 bundle with group generated by g2). If there were a path on M' joining ρ e M' to pg2, this path would cover a loop on Μ that would be orientation reversing for the locally ambient geometry. Again by hypothesis, no such loop exists, so M' = MUMg2 for some submanifold (a union of components) Μ С М' and Μ —» Μ is a two-fold cover (a principal Z2 bundle with group generated by τ — (#2)2). From the table above we have the following: V Rrv θχ -θχ θ2 -θ2 Ρι -Ρι Ρ2 -Vi Φ Φ Let L be the real line bundle over Μ associated to Μ via the representation Z2 - {±1} С Gii(R). It follows that puP2 e A°(M,L), θχ,θ2Ε A\M,L), Φ G A°(M, R) and that these items fit together to yield the form ώ as described in (v). Conversely, given the data (i), (ii), (iii), and (iv) satisfying (v), we can construct the principal bundle Z2 —» Μ —» Μ associated to L
328 7. Mobius Geometry and interpret Φ, pi, p2 as functions on it and θχ,θ2 as forms on it. These data then determine an element ώ G ^22(M,g) by the formula in (v), which in turn determines an umbilic-free, oriented, flat, locally ambient Mobius geometry of codimension one on M. We leave the details of this to the reader. ■ The three functions p\,p2, and Φ given in this result are related to the Monge form of the equation for the surface. In [G. Cairns, R. Sharpe, and L. Webb, 1994] it is shown that for a surface in 53 with invariants Pi, p2, Φ at some given point, there is an orientation-preserving Mobius transformation g G МоЬз(К) throwing the point to the south pole such that standard stereographic projection of the 3-sphere into space sends the image surface to a surface in space given locally by an equation of the form z=-(x2-y2) + -(plX3+p2y3) + 24 {(3 + Pi + f iPi )χΑ + 4(&Pi - PiP2)x3y + 6Φχ V + 4(ξιΡ2 +PiP2)xy3 + (-3 -pl+ ЬР2)У4} + 0(5). (6.5) Moreover, the forms θ\ and θ2 agree with dx and cfa/ at the origin. There are exactly two such elements g G МоЬз(К) that will do this job, and they differ by left multiplication with r = diag(l, —1, —1,1,1). The effect of r on the equation is to change the signs of χ and y, and this accounts for the indeterminacy of the signs of pi, p2, #ι, θ2. We may interpret the existence of g as a map 7: Μ —» {e, r} \ G, in which case the canonical map G-+S3 is a left inverse for 7. Exercise 6.6. For a torus of revolution in space with radii r and R, show that the invariants p\ and p2 vanish identically while Φ = 4 . □ Exercise 6.7. For a closed oriented surface Μ with generic umbilic points, show that the monodromy of L along any closed path counts the number, mod 2, of umbilics enclosed in it. □ Exercise 6.8. Read enough of Fialkow's paper [A. Fialkow, 1944] to show that, for the case of surfaces in space, the tensors described in Fialkow's Theorem 29.2 may be described in terms of our invariants by (i) Gijdxidxj =θ\+θ\, (ii) Bijdxidxj — Θ2 — 0f,
§6. Immersed Surfaces 329 (iii) Eijdxtdxj = (2 + ±(3p? -vl) ~ Шт) - *))*? + ((ζιρ2) - (&Ρι) + VxVl)Wl + (2 - |(Ρ? - 3ρ1) - |(Φ - (ξ2Ρ2)))θΙ In particular, show that in the case of a torus of revolution, the deviation form is Eijdxidxj = h + ^Φ J 0? + f 2 - ^Φ J 0|. □ Exercise 6.9. Verify that Fialkow's invariants do indeed determine our invariants in the case of a surface in space. □
8 Projective Geometry In the elementary sense, to solve a differential equation means to find a change of coordinates in which one recognizes the equation as one of known type. To truly be able to carry this out means that one is somehow able to deal with the differential equation in all possible coordinate systems. One of Elie Cartan's ideas was to develop a theory of differential equations that was independent of the coordinate system used to express it. This idea of Cartan is very ambitious and includes both ordinary and partial differential equations within its scope. It can be especially useful in the study of "overdetermined" partial differential equations, which often arise in differential geometric settings. In this context, the following kind of question arises. Given a system of differential equations on a manifold, can one "geometrize" the system in the sense of associating to it, independently of any coordinate system, a canonical Cartan geometry (i.e., a connection) from which alone one can recover the original system of differential equations? Generally speaking, the answer is no. There do exist, however, some remarkable cases of both ordinary and partial differential equations in which the answer is yes. The first example of this is Cartan's five-variables paper referred to in the preface. There he introduced the method of equivalence for solving this question. Many authors have continued to study Cartan's method of equivalence. For example, Tanaka (see [N. Tanaka, 1978] and the references therein) took up this line of work and addressed the general problem of attaching a Cartan geometry to a certain kind of A;-dimensional distribution on an η-dimensional manifold. In favorable cases, this yields a Cartan geometry.
332 8. Projective Geometry We will not study Cartan's general program here. For that we refer the reader to [R. Bryant, S.-S. Chern, R. Gardner, and H. Goldschmidt, 1991]. In this chapter, the theme we have been discussing is taken up in the particular context of the "geometrization" of a certain class of systems of second-order ordinary differential equations on a manifold. These systems include the equations for geodesies in a Riemannian manifold as examples. The Cartan geometry that solves the equivalence problem here is modeled on projective space, and the solutions of the equations are recovered as the geodesies of the projective geometry. In §1 we establish the basic notation for the η-dimensional real projective model space Pn with fundamental group G, the group of (almost) all projective transformations (Definition 1.3). In §2 we introduce the basic terminology of projective Cartan geometries and define the special geometries, including the normal geometries. We also show the existence of a canonical projective gauge associated to a fixed coordinate system (Proposition 2.2). In §3 we define geodesies in a projective Cartan geometry (Definition 3.4). We show that the set С of all curves on a manifold Μ that satisfy an ordinary second-order differential equation of a certain type determines a unique normal projective connection on Μ with С as its set of geodesies (Theorem 3.8(iii)). This is Cartan's version of the geometry of paths of O. Veblen and T.Y. Thomas. In Exercise 3.3 we point out that this theory applies to the second-order ODE y" = A(x, y) + B(x, y)y' + C(x, y){y')2 + D[x, y)(y'f, an example of which Cartan was particularly proud. In §4 we show that the set of geodesies of a Riemannian geometry on Μ is also of this type, and we determine the corresponding normal projective geometry (Theorem 4.1). We then apply this to prove a result of Beltrami stating that a Riemannian manifold Μ whose geodesies are all straight lines in some coordinate system must be of constant curvature (Theorem 4.2). The material in the first four sections is loosely modeled on [E. Cartan, 1924]. Finally, in §5 we take a brief tour of some of the literature associated with projective connections. We mention here the recent publication of [M.A. Akivis and V.V. Goldberg, 1993] in which the reader will find further details about projective Cartan geometries. §1. The Projective Model Projective space Pn = Pn(R) was defined in Definition 1.1.3. Three points of Pn are said to be collinear if they lie on a single line (cf. Definition 1.11). One version of the fundamental theorem of projective geometry (cf. [P. Samuel, 1988], p. 18) says that, for η > 2, the group of diffeomorphisms of Pn that preserve collinearity is PG/n+i(R) with its standard action on
§1. The Projective Model 333 Pn (induced from the standard action of G/n+i(R) on Rn+1). Prom that point, of view it would make sense to take PG/n+i(R) as the fundamental group of the geometry. However, it is more convenient to work with the subgroup G = PS7n+i(R), which is not much different from PGln+i(R), as the following exercise shows. Exercise 1.1.* (i) Show that there is a canonical inclusion PS7n+i(R) С PG/n+i(R). (ii) Show that PS7n+i(R) is the identity component of PG/n+i(R). (iii) Show that PGZn+i(R) is J connected if η -h 1 is odd [ has two components if η -f 1 is even. □ Lemma 1.2. Let Ζ = {A7n+i G S7n+i(R) | A G R*} so that PS7n+i(R) = S7n+i(R)/Z. Then Ζ is the center o/S7n+i(R) and -{i if η -f1 is odd, ±1 if η Η-1 is even. Proof. To calculate Z, we have A7„+1 6 Sln+l (R)^A^ = 1^A = {1 * n + ) is odd' -rv/ ^ _j_^ if η-h 1 is even. It is clear that Ζ lies in the center. Let us show that if g is central, it must lie in Z. If g G S7n+i(R) is central, it must commute with diag(±l, ±1,..., ±1) e S7n+i(R) for all possible choices of the sign; when η > 2, this forces g to be diagonal. When η = 1, the fact that g must commute with ( and ( J forces it to be diagonal. Since the diagonal matrix g must also commute with all matrices of the form //. ' 0 0 Vo 0 0 -1 0 0 1 0 0 0 0 0 J-n—s— 1 it follows that g = XIn-\-i for some A G R. ■ In working with elements of G = PS7n+i(R), it will often be convenient in this chapter to represent these elements as matrices. Such an expression is only defined mod Z, but we shall not always include this proviso even though it must be understood to be present.
334 8. Projective Geometry Let eo, ei,..., en be the standard basis of Rn+1. It is easily seen that the stabilizer in G = PS/n+i(R) of the point [e0] G Pn is given by H=ihePSln+l(R) \h= (Z0 |Λ mod zj, where the diagonal blocks of the matrix have sizes lxl and η χ η, so it follows that the map G —» Pn sending # ι—► ^[e0] induces a diffeomorphism G/Я и Pn. As usual, we will generally work with G/H instead of Pn. Definition 1.3. The group PSin+i(R) = G is called the projective group in η dimensions. The pair (G, Я), where Я is described above, is called the projective model in η dimensions. φ Exercise 1.4. Show that the Lie algebras of G and Я are given by /1 η g = {<?GMn+1(R) | TYace(<?) = 0} , f|=^Gfl|/i=« ρ V0 α Exercise 1.5.* Show that the following three subgroups of Η are connected, are normal in Я, and have the following Lie algebras: b = Ъг = H2 = f>2 = hePSln+xCR) \he (z0 J4modz,*>o|, heg\h = ζ ρ 0 α hePSln+i{B) \he (J Η mod zj, /iGMn+1(R)|/i=[[ J J,Tracea = o|; hePSln+l(K)\he fj ^modzj, /i G Mn+i(R) I /i = 0 ρ 0 0 □ To describe all the normal subgroups of Я, we shall use the homomor- phism , я f R-* for η + 1 odd, [ R+ for η -f1 even, 0: sending :)~tt for η Η-1 odd, z\ for η Η-1 even.
§1. The Projective Model 335 Lemma 1.6. (i) He, Hi, #2, ο,ηά {e} are the only connected normal subgroups of Η. (ii) The other normal subgroups of Η are given by φ~ι(Α), where A is a nontrivial subgroup of Г R* for η + 1 odd, I R+ for η + 1 even. Proof. Let N С Η be a normal subgroup, Α/" φ {e}. The commutator relation ζ ρ 0 α 1 Q О 1 1 zga О 1 shows that N contains an element of the form ρ φ 0. The commutator relation ζ ρ 0 α (mod Z) with ζ ρ 0 α 2/ 0 0 / 1 (I-y)pa l 0 1 shows that N contains an element of the form ρ φ 0. The commutator relation 1 ρ 0 1 (mod Z) with 1 ρ 0 1 1 0 0 b 1 p(I-b-1) 0 1 contains an element #2 we are done, so let us assume that TV (mod Z) with ζ φ 1 or α φ 1. Since AT С shows that N D H2. If N ' ζ ρ 0 a j Sln+i(R)/Z, it follows that if a = I then 2=1. Thus, we may assume that α φ I. Since AT D #2, it follows that (z J (mod Z) G TV. The commutator relation ζ 0 0 a 1 0 0 6 0 [a A shows that (mod Z)e N for all b G Sln(R). Since P5i„(R) is 1 0 ,0 [o,b]. simple, it follows that N D H\. Since #i is the kernel of 0, it follows that Ν = φ~ι(Α), where vv y I R+ for η + 1 even. Now apply Exercise 1.5.
336 8. Projective Geometry Now we work toward the introduction of the normal submodule of Hom(A2(fl/l)), i>). Lemma 1.7. There is a canonical isomorphism ad\\)/\)2 ~ End(%/\)). Proof. Consider the homomorphism ad: f) —» End(g/f)). This is given explicitly by , , ζ ρ \ ( * * \ _ ( z Ρ \ ( * * \ ( * * \ ( z Ρ 0 a I \v • / \ 0 a ) \v • 7 \v * 7 V 0 α (α — zl)v * ) ' so it induces a map ad: f)/f)2 —* End(g/f)), which is easily seen to be an isomorphism. ■ For the next definitions and for later purposes, we will need the following composite Η module map, which is the projective analog of the Ricci homomorphism (cf. Definitions 6.1.3 and 7.1.19) and which we again denote by Ricci. 0 projection 0 Нот(Х2(в/ЬШ ► Hom(X2(g/i)),tytk) о * id®ad 9 л - X2(0/F)f<8>[)/F)2 » X2(fl/f)f ®End(0/f)) о ^ ^ contraction ^ ^ = λ2(0/ί))*<8>0/^<8)(0/())* > (g/i)f <8)(fl/f)f ί*Λ M* ® V ® W* . > (ll*(v) ί*-ί*(ν) И*) <g> W* (1.8) Definition 1.9. In projective geometry, the normal submodule of Hom(A2(g/fj), i)) is the kernel of the Я module map Ricci: Hom(A2(g, fj), fj) —» (β/l))* ® (β/1))* given by (1.8). Similarly, the symmetric submodule of Hom(A2(g/fj), fj) is the kernel of the composite Я module map Hom(A2(£,/l>), f,) ^? (fl/ί,)* <g> (fl/f,)* - A2(fl/^)*. ■ Exercise 1.10.* For 0 < p, q < n, define f £p9 if Ρ Φ q, Eo<Kn^r if Ρ = q, where the Epq G Mn+1 (R) are the standard elementary matrices. Let ег· G β/ϊ), 1 < г < n, be the standard basis (i.e., a is the class of Ei0).
§1. The Projective Model 337 (i) Verify that the epq (for 1 < p, q < n, and 0 = ρ < q < n) constitute a basis for \). (ii) Show that the element evq G f), 1 < p, q < n, corresponds to ep <g) e* under the composite homomorphism t) —► f)/f)2 —* End(g/f)) « Ы/Ь) ® (fl/f))*· (iii) Show that Accitf л e; β eP9) = {J*** " W ® «5 jff J > J; Q Lines Definition 1.11. A line in Pn is the set of all one-dimensional subspaces of a two-dimensional subspace of Rn+1. Three points of Pn are said to be collinear if they lie on a single line. Φ Note that the projective group acts transitively on the set of lines of Pn. A line on Pn meets any affine coordinate chart of Pn in a line with respect to the affine coordinate system. In such a coordinate system a line may be parametrized in the usual way; such a parameter is called an affine parameter. These lines, unparametrized, will be the models for geodesies in the next section. Let us study the generalized circles (cf. Definition 5.4.16) on Pn with a view toward determining which generalized circles are straight lines. Lemma 1.12. Let X G g have block decomposition of the form X = '. The generalized circle corresponding to the vector X is a straight q a line if and only if q ^ 0 and is an eigenvector for a. Proof. Recall that a generalized circle is the projection to Pn of the integral curve of an uq constant vector field V on G. We shall exclude the generalized circles that degenerate to points. Let X = lug(V) Ε g. Then the integral curves of V are the left translates of the one-parameter subgroup generated by X. This subgroup φ: R —» G is given by 71 = 0 Since left translation by g on Pn moves generalized circles to generalized circles and straight lines to straight lines, the character of any generalized circle corresponding to X will be determined by that of the projection of φ{ί) to Pn, which is the curve <£>(£) [eo]. If this curve on Pn lies on a line, then φ(ί)βο lies on a plane, which implies that the derivatives <^m)(0)eo = Xmeo, m = 0,1,..., span a space of dimension at most two. Consider
338 8. Projective Geometry *-(ί). Μ;)· **-(<£:?>.)· If the span of the first two of these vectors is one-dimensional, it follows that q — 0 and hence that Xkeo — zkeo, so that <^(£)[ео] = [^(0eo] = [ei2e0] = [eo]. This is the degenerate case that we are excluding. On the other hand, if the span of the first three of these vectors is two-dimensional, then g/0 and the vectors (zl + a)q and q are linearly dependent. This means that q is an eigenvector for a and it implies, by an easy calculation, that the span of all the Xmeo, m = 0,1,..., is also two-dimensional and hence that ^(i)[eo] lies on a Hne· " §2. Projective Cart an Geometries In this section we study some elementary aspects of projective geometry. We briefly study the canonical gauge associated to a coordinate system and the special geometries including the normal geometries. Definition 2.1. Let Μ be a smooth manifold. A projective geometry on Μ is a Cartan geometry on Μ modeled on the projective model. φ Let us write the connection and curvature forms as '=0 a)' _ /Ε Y\ _ , _ / άε + νΛθ άυ + εΛυ + υΛαλ \θ Α)~+πΑπ~\άθ + θΛε + αΛθ da + θΛυ + αΛα) As usual, we will also denote by π and Π the corresponding connection and curvature forms in their various gauge incarnations. Proposition 2.2. Suppose that (U,x) is a coordinate system on a manifold Μ equipped with a projective Cartan geometry. Then this geometry has a unique gauge (C/, π = ( . J) such that ε = 0 and θ{ = dxi for i = Ι,.,.,η.1 Proof. It is enough to show that this is true on a sufficiently small neighborhood V of each point χ G U since, by the uniqueness, all these gauges will fit together to give a gauge on U. Since, by Lemma 1.7, the adjoint representation ad: f) —» End(g/f)) is a surjection in projective geometry, then by Exercise 5.1.5, there is a neighborhood V of each point χ 6 U and a gauge (V, π) such that 0* = dxi for 1This result explains the frequent occurrence of the condition ε = 0 for projective gauges in the literature, e.g. [E. Cartan, 1937] and [S.-S. Chern, 1937].
§2. Projective Cartan Geometries 339 г — 1,..., η. Now we make a change of gauge via h: V —► Η of the form h — I 1 J so that ^(θ ~1&)(β α)(θ J)+/l*W/r=(e~eW * Since the components of θ span the space of 1-forms on V, we may choose the function b: V —► Rn in a unique manner so as to kill the first component of the new gauge. This yields the existence of the required gauge on V. Now we study the uniqueness. Note than an arbitrary change of gauge h:V —> Η will alter a gauge with first column I ) as follows. Writing h = ' 0 A '' We haVe Ad(h-l)n + h*u>H . a wH • + l 0 • dx a~l -a~lbA-l\ ( О Л fa b О Л /Ιώ •До Л -α Х6Л ldx + а*ин * aA~ldx • For the new gauge to have the same form as the old one, we must have A — al and — a~lbA~ldx + α*ω# = 0. The first equation implies that 1 = det h — adet A — an+1. Thus a = ±1, and so the second equation implies that 6 = 0. Hence h — ±1, and so h is the identity in PSZn+i(R). Torsion and Torsion Free Geometries Recall that a torsion free geometry is one for which the curvature takes values in the Η submodule f) С д. Here is a criterion for a projective torsion free geometry. Lemma 2.3. Let Μ be equipped with a projective geometry. Fix a coordinate system on U С Μ, and let (С/, π — ί ~ 1) be the gauge provided by Proposition 2.2. If we write ctik = Σκ/<η Гш#ь then the geometry is torsion free if and only if Г^ is symmetric in the last two indices. Proof. Torsion free θθ = 0θάθ + θΛε + αΛθ = 0. But in the gauge of Proposition 2.2, θ — dx so άθ = 0. Also ε — 0, so torsion free <=ϊ α Λθ = 0
340 8. Projective Geometry ^ Σ, aik Λ 0k = 0 for 1 < i < η 1<к<п ^ Σ Vikl°l A°k=Q for 1 < i < η l<fc<n &ГШ =Тцк for 1 < i,k,l < η. ■ Exercise 2.4.* Show that in a torsion free projective Cartan geometry the curvature satisfies the equation Α Αθ — θ Α ε = 0. [Hint: use the Bianchi identity.] □ The Curvature Function For later work it is useful to calculate the curvature function with respect to the gauge given in Proposition 2.2. Proposition 2.5. Let π = ( ~ J be a gauge of the form guaranteed by Proposition 2.2 and let di = -£^-. Assume the geometry is torsion free so that we may write the corresponding curvature as Π = ( . 1. Then the curvature function satisfies W K = 2 Σ Apq(d" dj)ei л ej* ® epq mod Hom(A2(g/f)), fj2), (ii) Ricci(K)= Σ Ι Σ Aqj{dhdq)\e*®e*. l<i,j<n \l<g<n / Proof, (i) Since we can write the curvature as Π = (Hpq) = 2_^ RpqEpg = - ^^ Hpq(di, Oj)0i A 0jEpq 0<P,9<7i 0<p,q<n l<i,j<n (where as usual Epq G Mn+i(R) is the standard elementary matrix), it follows that we can write the curvature function (using the notation of Exercise 1.10) as Κ = \ ]T Ilpq(di,dj)e*Ae*®E1 0<p,q<n l<ij<n = o Σ Прд(9ь^)е*Ле^®е 0<ρ,ς<η 1<2,J<71
§2. Projective Cartan Geometries 341 + 2ί^ΤΤ> Σ ПРрй-^КЛе*® Σ Етг 0<ρ<η 0<r<n l<2,J<n 2 Σ иря(д" дз)е* л ej* ® еРя 0<p,q<n I since 2_] npp(9i,9j) = 0 for all г, j'J ^ 0<p<n ' - Σ Apq(di,dj)e* Λ e* 0 eP9 mod Hom(A2(g/f)), f)2). (ii) Prom (i) and Exercise l.lO(iii), we have Ricci(K) — - V^ Apq(di,dj) Ricci(e1 Λ e* 0 epq) 2 l<i,j,p,q<n = 2 Σ A^(&, fy) (e*<5ip - e*5<p) 0 e* l<i,j,q<n l<M>i/<n = Σ ( Σ ^(ад))е:®е;. 1<г,д<п \l<j<n / Special Geometries As we saw in Lemma 1.6, t) contains two ideals, t) D t)i D t)2- We study the two subclasses of torsion free geometries for which the curvature takes values in f^> г = 1,2, respectively. Curvature type t)\. This means that the curvature has the form Exercise 2.6. Show that for a geometry of curvature type fji, the Bianchi identity implies, in addition to the relations of Exercise 2.4, the relation ΥΑθ = 0. □ Curvature type f)2· This means that the curvature has the form loo
342 8. Projective Geometry Proposition 2.7. A projective Cartan geometry of dimension η > 2 with curvature in f)2 is flat- Proof. The Bianchi identity is Ε Υ θ Α Ε Υ\ (ε ν θ ΑίΛθ α In the present case, where Ε = θ = A — 0, the (2, 2) component of this identity is 0 = θ A Y. It follows that Υι — θιΑ фг for some 1-forms фг. Thus, for η > 2, Yi has all three of #ι, θ2, #з as factors, which is impossible for a nonzero 2-form. Hence Yi = 0 for ι > 1, and so Υ — 0. ■ Normal and Symmetric Projective Cartan Geometries Somewhat more subtle in their definition than the subclasses of geometries considered above are the normal geometries and the symmetric geometries. Definition 2.8. A projective Cartan geometry is called normal (respectively, symmetric) if its curvature function Κ: Ρ —> Hom(A2(g/f)), f)) takes values in the normal (respectively, symmetric) submodule described in Definition 1.9. # Note that a normal geometry is symmetric and that both geometries are of type f)i. Proposition 2.9. A type f)i projective geometry is symmetric. Proof. It suffices to show that this is true for a single gauge, so we shall use the gauge of Proposition 2.2. Let us consider first the case η — 2. In this case we have Ricci(K)= Σ Ι Σ Agj(dudg)\e;®e;, 1<M<2 \1<9<2 J so the symmetry follows from the calculation Σ V(diA)- Σ АяЛд2,дд) = А22(д1,д2)-Ап(д2,д1) l<q<2 1<9<2 = А22(д1,д2) + А11(дид2) together with the fact that Ε — 0 and Ε + Trace A — 0. For the case η > 3, we use the part of the Bianchi identity iffl = [Π,π] coming from the (2,1) block, which reads (from Exercise 2.4 and the fact that ε = 0) Α Α θ = 0. Evaluating this equation on the triple dq A di A <9j, we get
§3. The Geometry of Geodesies 343 ° = Σ Aqk^Ok(dqAdiAdj) l<k<n = Σ {Agk(9qAdi)6kj + Aqk(diAdj)6kq + Aqk(dj Adq)8ki) l<k<n = Aqj(dq A di) + Aqq(di A dj) + Aqi{da A dq). Now Trace A — Trace Π — Ε — 0, so summing on q we get 0 = Σ ^«(0дЛй)+ Σ Aq№jAdq), l<q<n 1<9<^ or J] Л„-(йлЗ,)= 53 A9i(djAd9). l<q<n 1<9<™ It follows that Ricci К takes its values in the symmetric part of (g/i))* §3. The Geometry of Geodesies In this section we are going to study collections of regular curves c: I —> U С Rn satisfying a system of ODEs of the form E(P): '^ + P^ = ^ + p2(c) = _ = cn + Pn(c)^ where Ρ = (РЬ...,Р„) with Р*(г;) = Σ having coefficients α^·: £7->R. Lemma 3.2. (i) For each point χ G U and for each vector ν G TX(U), a system of the form E(P) has a unique solution c(t), defined on a neighborhood ofO, satisfying c(0) = χ and c(0) = v. (ii) Let E(P) and E(Q) be two systems of ODEs of the above type. They have the same solutions if and only if there is a 1-form φ G Al(U) such that Pi{v) — Qi(v) + У{ф{у) for all v. Proof, (i) This is a consequence of the usual existence result for solutions of ODEs. Of course, it does not depend on the Ps being quadratic functions, (ii) 4= Since с\+Рг(с) Ci+Qi(c) + Ci<f)(c) Ci + Qi(c) r = : = : + 0(c), Cj Cj C{ it follows that E(P) is satisfied 44> E(Q) is satisfied. => Suppose that a curve с satisfies E(P) 44> с satisfies E(Q). Taking the difference of the equations E(P) and E(Q) yields
344 8. Projective Geometry P1(c)-Q1(c)= =Pn(c)-Qn(c)=m (gay)_ Cl Cn By (i) there are solutions с of E(P) satisfying c(0) = χ and c(0) = ν for χ and г> arbitrary so that φ is defined for every χ G £/ and for every г> G TX(U). Since Ρ and Q are quadratic, it follows that ф{у) is linear in v. Thus φ is a 1-form and Pi(v) = Qi(v) -f νιφ{ν) for all v. ■ Exercise 3.3. Show that for η = 2, the system of ordinary differential equations E(P) reduces to a single equation that may be put in the form y" = A + By' + C(y')2 + D(y'f (where A, B, C, D are functions of (x,y) G R2 and ' denotes derivative with respect to x). □ Geodesies Definition 3.4. Let Μ be a manifold equipped with a projective geometry. A geodesic on Μ is a curve that develops to a straight line in the model geometry. ft Note that by Proposition 5.4.13, development is the same no matter what gauge is used. Thus, the notion of geodesic is well defined. Note also that we are concerned with unparametrized geodesies.2 We are going to show that these geodesies of Μ satisfy a second-order ordinary differential equation of type E(P) for some P. Then we shall show a converse, that a second-order differential equation of this type determines a unique normal projective geometry on Μ whose geodesies are the solutions of the differential equations. (ε ν \ * I) be a projective Cartan gauge and let c: (a,/?) —> U be a smooth curve. Then с is a geodesic if and only if for all i = 1,2,..., n, the following expressions are equal: gj(c)4-Pi(c) Oi(c) ' where "·" — d/dt and Pi (1 < г < η) are the homogeneous quadratic polynomials 2 We could of course also discuss parametrized geodesies, that is, those geodesies for which the parameter of the development on Pn is a projective parameter, that is, a parameter that, in any affine coordinate system, is a linear fractional transformation of an affine parameter.
§3. The Geometry of Geodesies 345 Pi{v)= Σ <*ijWj(v)· 1<3<п In particular, the geodesies are completely independent of the choice of v. Proof. We must show that the given differential equation is a necessary and sufficient condition for the curve c(t) to develop (cf. Definition 5.4.15) to a straight line in the model space Pn. We can develop the associated function 7r(c): (α,/?) —► g, to obtain a curve c: ((a,/?),0) —► (G,e) satisfying (jg(c(t)) = n(c(t)). The projection of this curve to the model space Pn is c(£)[eo], where [eo] G Pn, and this curve in Pn is covered by the curve c(t)eo in Rn+1. Now c(£)[eo] lies in a straight-line segment in Pn if and only if c(t)eo lies in a plane of Rn+1; the latter condition is equivalent to the condition that c(t)eo, c(t)eo, and c(t)e$ are linearly dependent. Thus, c(t) is a geodesic if and only if there are functions μ(ί) and X(t) such that ceo = /лсео Η- Aceo or, equivalently, c_1ceo = /ieo H- Ac-1 ceo- But we have c_1c = cjg(c) = 7r(c). Also, since (c_1) = —c~lcc~l, we have 7r(c) = -C_1CC_1C H- C~lC = -7r(c)2 -f C_1C. Using these identities, the equation that с be a geodesic becomes 7r(c)"e0 + тг(с)2е0 = /ie0 -l· λπ(έ)β0. The eo component of this equation is the only component involving μ, and it may be satisfied by taking it as a definition of μ. The other components of this equation may be expressed in terms of the constituent forms of π to yield 0*(c) +Eo<j<naij(c)0j(c) λ г · 1 fp- =λ, for г = 1,...,η. ■ Corollary 3.6. If (C/,π = ( й I) is a gauge for which dxi = θι, then the equation for the geodesies is the system E(P), where Ρ = (Pi,..., Pn) and Pi(v) = T,o<j<nai3(v)°3(v)· Proof. In this case we simply have щ{с) = di, so ^тгг(с) = J^c* = c\. Ш Exercise 3.7. Verify that the condition of the equality for all i = 1, 2,..., η of the expressions Oi(c) is a condition that is independent of the way in which the curve с is parametrized. □
346 8. Projective Geometry Theorem 3.8. Suppose we are given, on an open set U С Rn, a collection of curves С constituting the solutions of a system of ordinary differential equations of the form E(P) (cf. Eq. (3.1)). (i) There are many projective geometries on U having the curves С as their geodesies. We can represent the geometrical equivalence class of (ε ν \ й J with 6i = dxi, 1 < i < η, and ε — 0. Among these geometries, (a) the block a — (ctij) is determined by С up to addition by a matrix S = (Sij) = (/ ^Ajjfcflfc) where Л^ is skew symmetric in the last two indices but is otherwise arbitrary, (b) the block υ may be chosen arbitrarily, and (c) the variation by the matrix S together with the form ν parametrizes in a one-to-one fashion the geometric equivalence classes of geometries whose geodesies are C. (ii) Among the geometries described in (i), there is exactly one choice of the a yielding a torsion free geometry. This choice is independent of the block v, so this block parametrizes the torsion free geometries in a one-to-one fashion. (Hi) Among the geometries described in (i), there is a unique normal projective Cartan geometry. Proof, (i) Let us find all the projective Cartan geometries with С as geodesies. First we note that although the curves С satisfy the system E(P), this doesn't mean that Ρ is determined by C. In fact, by Lemma 3.2(ii) we know that С determines Ρ only up to the addition by a term of the form νφ(υ), where φ is an arbitrary 1-form on U. By Proposition 2.2 we know that any projective geometry on U С Rn, with standard coordinates χ = (χχ,... ,xn), has a unique gauge π = (ε ν\ a J on Rn satisfying ε — 0 and θ = dx. Thus, we merely need to determine all the ways of choosing the remaining blocks a and ν so that the curves С are geodesies. The only a priori condition on the block a is that Trace a) — 0 (recall that ε = 0). Now by Proposition 3.5 the geodesies of a geometry are the smooth curves ο:(α,β) —► U that are solutions of the equations stating the equality, for alii = 1,2,..., n, of the expressions
§3. The Geometry of Geodesies 347 Сг + 2^1<г<п агз\Ч^З It follows that the geometries in which the curves С are geodesies are those for which У^ aij(v)v3 = Pt(v) +ντφ(υ), г = 1,...,п, (3.9) 1<3<п where φ G ^(^O 1S arbitrary. Thus, the collection of geometric isomorphism classes of geometries with the required geodesies is in one-to- one correspondence with the collection of solutions a of (3.9) satisfying Trace(a) = 0. Let us solve Eq. (3.9) for the forms агз. Write the homogeneous quadratic polynomials Рг as P%(vi,...,vn) = ΣΚ*3\ν)ν3ΐ where ϋτ3(υ) is some linear combination of the vts. The R{3 are determined by this equation up to addition with Sl3 = Y^AijkVk, where Λ is skew symmetric in the last two indices but is otherwise arbitrary. Now set агз = Rl3 for г φ j, 1 < г, j < η, and агг = Rti - £ ^x<fc<n Rkk for 1 < г < п. Then Trace(a) = 0, so the π = ( J constitutes a projective gauge. Moreover, for 1 < i < n, we have3 Σ aijdxj=Pt(dxi,...,dxn)-dxtl- ^ Rkk J . 1<3<п \П l<k<n J Thus, Eq. (3.9) is satisfied by this choice of π no matter how the block ν is chosen and no matter how the a is varied by the matrix S. Moreover, by Proposition 2.2 each of these variations leads to a distinct geometric isomorphism class of projective Cartan geometry. This finishes the proof of (0· (ii) By Lemma 2.3, a geometry is torsion free if and only if, for a gauge of the type given in Proposition 2.2 (as is the case in the proof of (i)), when we write агк — ^Γ^ί/, then Гг^ is symmetric in the last two indices. For a geometry with geodesies C, the choice of the агк is determined up to the replacement where Агзк is skew symmetric in the last two variables. Clearly, we may take Агзк — —\(Yi3k — ^гкз) to be minus the skew-symmetric (in the last two indices) part of Tiki and so replace π by a torsion free connection. It is 3The product of 1-forms is the symmetric product here.
348 8. Projective Geometry clear from this that there is a unique block a such that π = ( , J has geodesies С for which this is true. The only remaining freedom in choosing the connection comes from the choice of the block u, which may be chosen arbitrarily and which therefore parametrizes the geometric isomorphism classes of torsion free geometries with geodesies C. (iii) Now we show that among the geometries constructed in (i) is a unique normal geometry. Since a normal geometry is torsion free, by (ii) only the freedom to choose the block ν remains. Since a normal geometry is of type I)i, the (1,1) block of the curvature vanishes. Let us write υ = θ1 a for some function a:U —> Mn(R), so that the curvature has the form T7 f° Y\ A _l. л ίυΑθ * ^ 11 = л Λι— απ-\-πΑπ=[ Λ . Λα \0 AJ \ 0 άα + θΛυ + αΛα] θιαΛθ * 0 άα + θΛθ*α + α Λ α In particular, the vanishing of the (1,1) block is equivalent to the matrix a's being symmetric since 0 = θιαΛθ = ^2 Qiaij л fy = Σ, (av ~ a3^i л @з ■ l<i,j<n l<i<j<n Finally, we study the effect on the curvature function of varying the block v. We see that the change υ *-* v + 0lb (where b is a symmetric nxn matrix) has the following effect on A: Α = άα + θΛυ + αΛα^άα + θΛ(ν + вьЬ) +аЛа = А + 0Лв*Ь so that Apq(di,d3) .-> Apq(di,d3) + (0 Л вьЪ)ря(дг,д3) = Ард(д{,д3)+ Σ 0РА0кЬкд(д{,д3) 1<к<п = Аря(дг,д3) + 22 №Pihj - SPJ6kl)bkq l<k<n — ^pq 0piOjq 0pjbiq. Since by Proposition 2.5(ii) we have mcci(K)= Σ Σ л«(0„з,))е;®е;, l<i,j<n \l<i/<n it follows that
§4. The Projective Connection in a Riemannian Geometry 349 Ricci(K) h-> Ricci(K) + ^ Σ (^Aj - 5ggb»j) e* 0 e* l<2,j<n \ 1<ς<η / = Ricci(K) - (n - 1) ^ 6l:7e* 0 e*. l<i,j<n By Proposition 2.9, Ricci(K) takes values in the symmetric part of (β/ϊ))* 0 (β/l))*, so ^ follows that for η > 1 there is a unique choice of the symmetric matrix (bik) such that Ricci(K) = 0. Thus, among the projective geometries whose geodesies are C, there is a unique normal geometry. Corollary 3.10. Suppose we are given a manifold Mn on which there is a system Ε of second-order ODEs such that Μ is covered by coordinate systems in each of which the system Ε assumes the form of equation E(P) for some Ρ (с/. Eq. (3.1)). Then there is a unique normal projective Cartan geometry on Μ whose geodesies are the solutions of E. Proof. If we choose a coordinate system in which the system Ε has the form of equation E(P), then on each coordinate patch this corollary is just 3.8(iii). The uniqueness allows us to fit together the geometries on the various coordinate patches. ■ §4. The Projective Connection in a Riemannian Geometry In this study we study the geodesies in a torsion free Riemannian geometry on Μ and construct from them a normal projective Cartan geometry on M. Thus we have two model pairs, the Riemannian pair (дше, ()ше) with group i?Rie and the projective pair (g, \)) with group H. We shall regard gpue as a subalgebra of g and Нще as a subgroup of Η according to с А С fl = < ( а . J | α + trace A = О L Яше = Ш a) lceR",^eO„(R)| с я = PSin+i(R) = Uac ЬЛ е sin+1(R)\/ζ. The proof of the following theorem is really an application of the ideas in the proof of Theorem 3.8. However, for the sake of clarity we include all the details.
350 8. Projective Geometry Theorem 4.1. The equation for the geodesies in a Euclidean geometry is of the type E(P), so there is a canonical normal projective connection with these geodesies. If the geometry is Reimannian (i.e., a torsion free Euclidean geometry) with gauge and curvature given by ω={θ a)> П=[о А then the associated normal projective geometry has gauge and curvature given by 0 г/ where θ a and υ=^ΓΪ Σ { Σ А1Л^еЛек \<к<п ^l<i<n J _'0 dv + υΛθ" 10 Α + θΛυ Ι' Proof. By Proposition 6.2.7 the equation for the geodesies of a Euclidean geometry with gauge π is в^сУ+Р^с) = = θη(ό)'+Ρη(ό) θι(ό) '" θη(έ) where Since this has the form of equation E(P) (see Eq. (3.1)), it follows that there is a unique normal projective Cart an geometry on Μ whose geodesies are the Euclidean geodesies. Let us determine this projective geometry explicitly in the case when π is torsion free. We begin by studying a projective gauge of the form Ό ι/ π = ( й ) ^*е'' ^ anc^ a come fr°m the Riemannian gauge, ε = 0, and ν is to be determined) so that the corresponding curvature is ν Αθ dv + υ /\α 'ЧаЛ^ άα + θΛν + αΛα Since ω is torsion free, it follows immediately that π is also torsion free. The equation for the projective geodesies is (by Proposition 3.5) the equality, for 1 < г < η, of the expressions
§4. The Projective Connection in a Riemannian Geometry 351 ^(έ)· + Σκ3·<η%(%(^) Note that this is identical to the equation for the Riemannian geodesies and is independent of the choices for the block v. It remains to show how to choose the block υ to obtain a normal geometry. The projective curvature mod f)2 is ϊ1=(υΑθ * ^ί=Ω+ί^Λ* * ^ \ 0 άα + θΛυ + αΛα] \ 0 θΑυ]' Since a normal geometry is of type I)i, it follows that we must choose the block υ so that the (1,1) block of the curvature vanishes. If we write υ = 9lb for some function b: U —► Mn(R), the (1,1) block of the curvature is уле = еььле= γ^ е{ь{3-лθά = Σ (bv-Ы^лЬ· l<i,j<n l<i<j<n In particular, the vanishing of the (1,1) block is equivalent to the matrix 6's being symmetric. Let us assume this from now on. We have Π = (ΠΡ9)= Σ Пряеря = \ Σ Πρ,φ,^ΜΛ!^ l<i,j<n = - Σ (Apq(di,dj) + θρ ЛУд(дг,д^)в{ AOjEpq 0<ρ,ς<η 1<2,J<71 (where as usual Epq G Mn+i(R) is the standard elementary matrix). It follows that we can write the curvature function (using the notation of Exercise 1.10 and the calculation of Theorem 3.8) as l<i,j,p,q<n mod#ora(A2(g/f)),f)2). Applying the Ricci homomorphism Ricci : Hom(A2(g/f)), fj) —► (g/f))* 0 (β/Ι})* (Definition 1.9), the curvature function is mapped (Exercise l.lO(iii)) to Ricci{K)=l- Σ Σ (\-№,9д)+^Л^№,ад е*®е* = 5 Σ f Σ Aqj{dudq)\el®e) l<i,j<n \ l<q<n J
352 8. Projective Geometry e* ® e* + \ Σ Σ ь^л м&л) < ®е. ^ Σ ( Σ л«(йл)) l<z,j<n \ 1<ς<η / + Ο Σ/ Σ/ brj(6qi6rq - <Sgg<Sri) I ^ Σ ( Σ ^·φα)κ®*; e* ® e* 1<z,j <n Since Ricci(K) takes values in the symmetric submodule of (β/f))* 0 (g/ί})* (Proposition 2.9), it follows that Ricci(K) — 0 if and only if we take bn — ^ Aqj(di,dq). l<q<n The Configuration of Geodesies The following figures serve to remind the reader that central projections provide, locally, a geodesic-preserving diffeomorphism between the n-sphere and Euclidean space and also between η-dimensional hyperbolic space and Euclidean space. geodesic image of geodesic under imtral projection plane tangent to south pole Now we wish to show the converse of this, due to Beltrami, that a Rie- mannian manifold whose geodesic configuration is locally the same as Euclidean space must have constant curvature. Theorem 4.2. Let Μ be a Riemannian manifold such that each point has a coordinate system in which every geodesic is a straight line. Then Μ has constant curvature.
§5. A Brief Tour of Projective Geometry 353 Proof. By Proposition 6.4.5, it suffices to show that А — c0 /\θι for some constant c. The hypotheses imply that the normal projective geometry on Μ with the same geodesies as the Riemannian structure on Μ must be flat. Projective flatness means Π = 0, so in particular, by Theorem 4.1, A -f Q/\v = 0. Since A is skew symmetric, it follows that the diagonal terms of 0 Λ ν must vanish. Thus, Vi = λ^ for some functions λ^, 1 < г < п. Thus, Aij — —XjOi Л #j, and then the skew symmetry of A implies that λ^ is independent of г, 1 < г < η, say λ^ = —с. Thus, υ = —ев1 and A = c0 Л 0£, but we need to show that с is constant. For η > 2, the constancy of с is contained in Lemma 6.4.1 (it is a consequence of the Riemannian Bianchi identity). For η — 2, the projective gauge and curvature are /0 -cQx -οθ2\ π = 0i 0 012 \#2 -012 0 J ( 0 0 = Π = άθι+θ12Λθ2 \άθ2-θ12Αθι respectively. Thus, 0 = d(c0i) - c02 Л 0i2 = dc Л 0i + cd0i - с02 Л 0i2 = dc Λ 0i - C0i2 Λ 02 - c02 Λ 0i2 = dc Л 0i, 0 = d(c02) + c0! Л 0i2 = dc Л 02 + cd02 + c0i Л 0i2 = dc Л 02 + c0i2 Л 0i + c0i Л 012 = dc Л 02. Hence, dc = 0 and с is constant. ■ §5. A Brief Tour of Projective Geometry In this section we describe very briefly a small sampling of the literature related to projective Cartan geometries. Cartan's Notion of a Subgeornetry Cartan studied a general (and perhaps peculiar by the usual standard) notion for the induced geometry on a submanifold. In Cartan's version we have a submanifold Мш С Pn. We select for each point χ G Μ a "normal" space for Μ consisting of a projective subspace Рп~ш(х) transverse to TX(M) and use it to obtain a reduction of the principal bundle. (A more usual way is to try to define a projective normal vector to a sufficiently non- degenerate surface. Both methods are discussed in [M.A. Akivis and V.V. -{d(c0i)-c02A0i2} 0 -d0i2 + c0iA02) -{d(c02)+C0lΛ0l2}, d0i2 - C0i Λ 02 0
354 8. Projective Geometry Goldberg, 1993], pp. 173ff). The result of this is that we obtain an induced projective Cart an geometry on Μ, but it will depend on the choice of "normal" spaces. Chern proved [S.-S. Chern, 1937] that in the analytic category every projective Cartan geometry on Μ arises in this way, provided {h(m2 + 2ra — 1) if m is odd, \(m2 + 2ra -hi) if m is even. Families of Subrnanifolds In [S.-S. Chern, 1943], Chern generalized both Cartan's result on geodesies in §3 and the analogous results of Hatchtroudi on hypersurfaces by showing that to certain (k -f l)(n — k) parameter families of A:-dimensional subrnanifolds of Mn, there may be uniquely attached a normal projective connection. Yen [C.T. Yen, 1953] gave a geometric version of this. Chern and Griffiths [S.-S. Chern and P. Griffiths, 1978] showed that flat projective connections are characterized by having a two-parameter family of totally geodesic subrnanifolds, suitably distributed (cf. Chern's remark on page xxiv of [S.-S. Chern, 1978]). Differential Equations If P2 = G/H, then the bundle of contact elements (i.e., pairs (x,L) where L С TX(P2) is a one-dimensional subspace) is В — G/K (where К С Η is the subgroup of ϋΓ, of codimension one, fixing a given line in P2). Cartan showed that a general ordinary second-order differential equation on a two dimensional manifold M2 determines a canonical Cartan geometry, modeled on В — G/K, on the bundle B(M) of contact elements of M. There is a differential invariant arising from the curvature of this geometry which vanishes if and only if the equation is of the type considered in §3. In this case the geometry is the "pull up" of the projective geometry on M. Chern [S.-S. Chern, 1940] generalized this to a third-order ODE on a surface M2. He also found a different invariant I that can be calculated from the equation. When I = 0, he associates to the equation a Cartan geometry on the 3-manifold of solution of the ODE modeled on Gio/Hj, where Сю is the group of contact transformations in the plane (i.e., the Lie sphere group). On the other hand, if Ι φ 0, then he associates to the equation a Cartan geometry with fundamental group the five-dimensional conformal group. For some purposes (e.g., in control theory) the equivalence under change of variables of second-order ordinary differential equations needs to be studied with respect to a smaller group than the group of all diffeomor- phisms of the space of variables. Taking the case of an equation of the form
§5. A Brief Tour of Projective Geometry 355 y" — F(x, у, у'), (х, у) е R2, we may wish to consider changes of variables of the form χ ι—> ф(х), у ι—> ф(х,у). The advantage of this kind of change of variables is that it preserves the role of the independent variable. It is studied in [N. Kamran, K.G. Lamb, and W.F. Shadwick, 1985]. CR Geometry CR Geometry is the study of the invariants of real hypersurfaces M2n~l С Cn that are unchanged under complex analytic change of variables of Cn. For a recent treatment of CR geometry, see [H. Jacobowitz, 1990]. Chern [S.-S. Chern, 1975] showed that when Μ is real analytic, there is a canonical projective Cartan geometry induced on Μ which is such an invariant.
Appendix A Ehresmann Connections Charles Ehresmann introduced his definition of an infinitesimal connection on a principal bundle in his paper [C. Ehresmann, 1950]. It put the notion of a Cartan connection on a rigorous footing and generalized it at the same time (cf. Proposition 3.1). Ehresmann's definition was accepted as the definitive definition of a connection. The reader can find a full exposition of this point of view in [S. Kobayashi and K. Nomizu, 1963]. And yet there has been some dissatisfaction with this state of affairs, leading one well-known mathematician to say that the theory of connections was founded on the wrong definition. This is a consequence, it seems, of two factors. First, Ehresmann connections are extremely general and so include perhaps much more than is actually interesting.1 Secondly, Ehresmann's formulation hides the similarity that connections can have to the Maurer- Cartan form on a Lie group and so provides an unnecessary obstacle to one's understanding. On the other hand Ehresmann connections appear as important components of the Cartan connections (cf. 6.5.12 ii) and 7.4.21) as well as being the foundation for the notion of covariant differentiation, so they do have a job to do. Perhaps the relation between Cartan connections and Ehresmann connections is something like the relation between rings of algebraic integers *Ιη a different context, Chern [S.-S. Chern, 1966, p. 167] comments on what is perhaps the inevitability of this situation.
358 Appendix A. Ehresmann Connections and arbitrary commutative rings in that each degree of generality illuminates the other. We pass now to the definition given in Chapter 6 of an Ehresmann connection on a principal bundle. We repeat it here for convenience. Definition 6.2.4. Let Η —» Ρ —» Μ be an arbitrary principal bundle over Μ and let Ϊ) be the Lie algebra of H. An Ehresmann connection on Ρ is an (^-valued form 7 satisfying the conditions (i) Д^7 = Ас1(Л-1)7, (ii) 7(Xt) = χ for every Xef)m The curvature of the Ehresmann connection 7 is the (^-valued 2-form on Ρ given by άη + \ [7,7]. * In view of the apparent formal similarity between this definition and the definition of a Cartan connection, it is appropriate to remark on an important difference between them. The form 7 takes values in the Lie algebra of the fiber of the principal bundle, so its restriction to any tangent space of P, 7P: TP(P) —» I), must have a kernel whose dimension is at least the dimension of M. In fact, condition (ii) implies that 7 is onto, so it follows that dim ker 7P = dim Μ for all ρ Ε P. This circumstance is at variance with that of a Cartan connection, which has no kernel at all. What is the relation between Ehresmann connections and Cartan connections? There are two ways of considering this issue. For a first-order reductive Cartan geometry, we show in §2 that the notion of a Cartan geometry is equivalent to the notion of a bundle of frames together with an Ehresmann connection on it. An example of this correspondence is given in Exercise 6.3.6 in the case of Riemannian geometry. On the other hand, we shall see in §3 how an arbitrary Cartan geometry (P,cj) modeled on the Klein pair (G,H) determines, and is determined by, a special kind of Ehresmann connection on the principal G bundle PxhG. This correspondence is different from—and perhaps less geometric than— the one described in §2. §1. The Geometric Origin of Ehresmann Connections Perhaps the motivating reason for Levi-Civita's search for what is now called the Levi-Civita connection was the need to find an analog of the vector gradient of Euclidean geometry (cf. Exercise 5.3.50) in the context of Riemannian geometry. In fact, Levi-Civita called his discovery the absolute derivative. The idea was generalized in 1917 by Η. Weyl to arbitrary vector bundles and was later developed by a number of people, including Schouten,
§1. The Geometric Origin of Ehresmann Connections 359 T.Y. Thomas, E. Cartan, and finally С Ehresmann. While this discovery had primarily analytical roots, it could also have resulted from certain simple geometric considerations, which I will sketch now. Given a vector field Υ on a manifold Μ and a vector Xx e TX(M), what should XX(Y) G TX(M) mean? We may of course formally write XJY) = \imYx+tXx~Yx. There is a small difficulty with giving meaning to the expression Yx+txx, but this may be solved using a Taylor expansion in t. More serious is the fact that the difference involves two vectors at different points of the manifold (i.e., in two different vector spaces), and so it has no meaning. In Euclidean space we can canonically identify any two tangent spaces by parallel translation, so the difference makes sense there, and in that case XX(Y) is just the usual vector gradient given by ^(Λ^τ+-+/-έ)=χ*(/ι)^+-+χ-(/-)έ· In a general Riemannian manifold, however, we would need to have an "infinitesimal connection," that is, an aifine isometry2 between the tangent spaces of the two "infinitesimally close" points χ and "x -f tXx" This will lead to aifine isometries TX(M) и Ty(M) whenever χ and у can be joined by a path, with the isometry being obtained by integration along the curve, that is, by composing the infinitesimal isomorphisms step by step along the curve. Of course, the isometry may well depend on the curve so chosen. This leads us to ask, "is the notion of an infinitesimal connection a natural one in differential geometry?" We are going to give some examples to show that the answer is yes. Involutes of Curves The first examples of this refer to the case of a curve Μ in the Euclidean plane. This is a trivial case, but it serves to help us orient ourselves. Up to translation, there is a unique orientation-preserving isometry between the tangent spaces of M. The geometry suggests two ways of identifying the tangent spaces corresponding to the Ehresmann and Cartan connections, respectively. The first is to slide the tangent line Ε a at A along the curve to В so that the point of contact on the moving line is fixed on that line. 2We do not wish to assume at the outset that the isometry preserves the origin. Of course, the Levi-Civita connection does preserve it. Even more generally, we can consider an infinitesimal connection on a manifold without any Riemannian structure, which would be a diffeomorphism between the tangent spaces of two "infinitesimally close" points.
360 Appendix A. Ehresmann Connections We might imagine the tangent line to be fastened rigidly to a small bead fitted around the curve but free to slide along it as in the figure. In this scheme, parallel translation preserves the origin. The second way is to allow the tangent line to roll without slipping along the curve. If a thread is wound along the curve and the part not in contact with the curve is pulled tight, then this part is tangent to the curve, and its successive positions as the thread is unwound identify the various tangent lines. In this scenario the points of the rolling tangent line trace out the involutes of the curve in the plane as in Figure A.l. The origin is not preserved under this version of parallel translation. FIGURE A.l In each of these cases the identification between tangent spaces at nearby points on the curve may be regarded as a one-dimensional distribution on the tangent bundle T(M) of the curve. This is perhaps clearest in the case of involutes, which, although they lie in the plane, may also be regarded as lying on the image of the canonical map T(M) -» R2. (χ,ι>) ι—► x+v As curves in T(M), the involutes are integral curves of the distribution on T(M). Although the involutes have cusps, as seen in Figure A.l, this is
§1. The Geometric Origin of Ehresmann Connections 361 merely an artifact of the mapping from the tangent bundle into the plane. There are no singularities in the corresponding distribution in the tangent bundle itself (as in the following picture). fibers ^ Μ к г ч 'I—^к—τ*—τ*—η ЦМ) This distribution gives an identification between the fibers of T(M) and indicates how, for an arbitrary manifold Μ of dimension n, a distribution ϊ( of dimension η on T(M) that is transverse to the fibers of T(M) describes an infinitesimal connection. In none of these examples have we been dealing with an Ehresmann connection on a principal bundle, so the reader has the right to ask, "what, if any, is the relationship between these notions and the definition of an Ehresmann connection?" We provide the answer to this question in the following two exercises. Exercise 1.2. (i) Let 7 be an Ehresmann connection on the principal bundle Η —> Ρ —> Μ. Show that the distribution <D on Ρ given by <DP — ker 7P satisfies the conditions (a) TP(P) = Tp(pH) <g> Ц, for all peP, (b) Rh*2)p = <Dph for all ρ G P, ft G H. (ii) Let Η —» Ρ —» Μ be a principal bundle and let CD be a distribution on Ρ satisfying conditions (a) and (b). Show that there is a unique Ehresmann connection 7 on Ρ giving rise to Ю as in (i) (cf. [S. Kobayashi and K. Nomizu, 1963], pp. 63-64). □ Exercise 1.3. Let Η —» Ρ —» Μ be a principal Η bundle and let Ю be a distribution on Ρ satisfying conditions (a) and (b) of Exercise 1.2(i). Let V —» Ε —» Μ be the vector bundle associated to the principal bundle Ρ by the representation (V,p) (i.e., Ε1 = Ρ χ я V^; cf. p. 39). Show that there is a unique distribution (De on Ε such that the derivatives of the canonical maps fv:P—*E = Ρ Χχ V sending ρ ι—> (ρ,ν) yield isomorphisms in the following diagram.
362 Appendix A. Ehresmann Connections 4 1- Show also that the distribution T>e is transverse to the fiber (i.e., no nonzero vector in T>e is tangent to the fiber). □ Rolling Without Slipping or Twisting Along a Surface in 3-Space A procedure generalizing the notion of an involute can be carried out for a surface Μ in 3-space. Given two points А, В G Μ and a curve σ: ([α, b], α, b) —» (Μ, Л, Б), we may allow the tangent plane to "roll without slipping or twisting" along the curve in order to identify the tangent planes at A and B. If we then translate the final position of the tangent plane so that its origin coincides with the point B, this procedure describes the Levi-Civita parallelism (in the induced metric) on the surface.3 The new ingredient here is that the distributions on the tangent bundle that arise in this way are generally no longer integrable, so the parallel translation definitely depends on the choice of the path. In Appendix В the notion of one manifold rolling without slipping or twisting on another in Euclidean space and its relationship to Levi-Civita connection and the Ehresmann connection on the normal bundle are studied in detail. §2. The Reductive Case In this section our aim is to show that in the case of a first-order reductive Cartan geometry, we have the following equivalence: first-order Cartan geometries on Μ with reductive model (f) Θ p, f)), group Я, and Cartan connection ω = ωρ -j- ω^ (principal Η frame bundles Ρ over Μ with fundamental form ωρ on Ρ and Ehresmann connection ω^ on Ρ Lemma 2.1. Let (Ρ,ω) be a reductive Cartan geometry on Μ modeled on (g, f)) with group Η so that $ = ϊ) Θ ρ is an Η module decomposition. Let ω^ 3If we take the raw, untranslated identification, we get the Cartan connection on the surface. See also the remarks at the end of §3.
§2. The Reductive Case 363 be the Ϊ) projection of ω. Then ω^ is an Ehresmann connection on Ρ and Ρ has a canonical {up to a choice of basis for p) interpretation as a frame bundle with group Η. Proof. Let us decompose the Cartan connection as ω — ω^ -f ωρ. Proof of Property 6.2.4(i) of an Ehresmann connection: By Definition 5.3.1(c)(ii), for a Cartan connection we have R^u = Ad(h~l)u. Hence R*hub + Д^р = Ad(h~l)ub +Ad(/T1)u;p, or R*hub - Ad(h~l)ub = -R*hup + Ad(h~l)up. Since the left side takes values in I) and the right side takes values in p, both sides vanish, and so R*hub = Ad{h-l)u>b. Proof of Property (U) of an Ehresmann connection. By Definition 5.3.1(c) (iii), for a Cartan connection we have ω(Χ^) = X for every X e f). Hence Χ=ω(Χ*)=ω)(Χ*)+ωρ(Χ*), or X-wb(Xt)=(Jp(xt). Since the left side takes values in \) and the right side takes values in p, both sides vanish so To interpret Ρ as a frame bundle, we refer to the procedure outlined in Exercise 5.3.21. ■ There is a kind of converse to this result that depends on the following notion of the fundamental forms on the principal frame bundle of a manifold. Definition 2.2. Let G/„(R) -» Ρ -» Mn be the principal G/„(R) bundle of frames on M. The fundamental forms #i,#2, · · · ,θη are the forms on Ρ defined as follows. A point ρ Ε Ρ consists of a pair (x,e), where e = (ei,e2,·.. ,en) is a frame at χ (i.e., a basis of TX(M)). Thus, for ν G TP(P), we may write π*ρ(ι/) = Oi(is)ei + Q<i(y)e<2. Λ Η- On(v)en, and this defines 4Such forms have also been called "canonical" or "solder" forms. The terminology depends on the context. When we are given the frame bundle of a manifold, the correct terminology is "fundamental" or "canonical" forms since, as described in the definition, they are canonically determined on the frame bundle. On the other hand, if we are given a principal bundle, then the "solder forms" are not canonical, but once they are given they allow us to uniquely identify this principal bundle with a bundle of frames so that the solder forms correspond to the restriction of the canonical forms. Thus, the solder forms allow us to "solder" (in the sense of "identify") the two bundles.
364 Appendix A. Ehresmann Connections the fundamental forms. Moreover, given any reduction Η —» Q —» Μ, the restrictions of the fundamental forms to Q are called the fundamental forms of the frame bundle Q. Φ Proposition 2.3. Let Η —* Ρ —* Mn be any reduction of the principal GZn(R) frame bundle associated to the tangent bundle of M. Let Ϊ) be the Lie algebra of Η and let 7 be an Ehresmann connection on P. Let ρ = Rn be the standard reprsentation of G/n(R) restricted to H, and let g = ί) Θ ρ be the Lie algebra for which i) is a subalgebra, the action of Ί) on ρ is the standard adjoint action, and [p,p] = 0. Let θ — (#i,02» · · · »^n)* denote the p-valued 1-form on Ρ determined by the fundamental forms. Then is a Cartan connection on Ρ modeled on the pair (g,f)) with group H. Proof. It suffices to verify condition 5.3.1(c). First we verify the condition 5.3.1(c)(iii). By Definition 6.2.4(ii), we have ί(Χ^) = X for every X el). Since X^ is tangent to the fiber, by Definition 2.2, ΘΡ(Χ1) = 0. Thus, ωρ(Χ*) = X for all X ef)- Next we verify the condition 5.3.1(c)(i), which says that ωρ:Τρ(Ρ) —» g is an isomorphism for each ρ G P. Since the dimensions are the same, it suffices to show surjectivity. By the verification of condition 5.3.1(c)(iii) above, the algebra i) lies in the image of ωρ. On the other hand, ker 7p has dimension η — dim Μ and is complementary to the tangent space of the fiber at p. Thus, by Definition 2.2, θρ | ker 7P: ker 7P —» p is an isomorphism and hence ωρ:Τρ(Ρ) —» g is surjective. Finally we verify the condition 5.3.1(c)(ii), which says that (Rh)*u — Ad(/i_1)u; for all h G H. The right action of Я on Ρ is given by the formula (ei,e2,.·. ,en) · h = I ^ Λαέ», ^Z ^«ei,..., J^ Λ<ηβ<) \ 1<г<п 1<г<п 1<г<п where h Ε Η С Gln(R). Now, writing ρ = (еь ег, · · ·, еп) Ε P, the defining equation π*ρ(ζ/) = #i(^)ei Η Η- вп{и)еп for the fundamental form yields 0i(i/)ei+--- + 0n(i/)en = 7T*pZ/ = (πο Дл)#р1/ (since π ο Rh = π) = K*ph(Rh*v) = 0i(Rh*v)(ei ■ ft) + ■ ■ ■ + 0n(Rh*is)(en · ft) = J^ ej(Rh*i>)hijei l<i,j<n
§3. Ehresmann Connections Generalize Cartan Connections 365 \<j<n 1<.7<™ from which it follows that θι(ν) = Σκκη^^*17)^ anc^ hence that R*h0 = h~le. Thus, = Ad(/i_1)7 + /i_10 (the first term comes from Definition 6.2.4(i)) = Aa(h~l)u. Ш Exercise 2.4. Verify that the correspondences given in Propositions 2.1 and 2.3 are inverse to each other and so complete the verification of the equivalence given at the beginning of §2. □ §3. Ehresmann Connections Generalize Cartan Connections Suppose we are given a Cartan geometry (Ρ,ω) on Μ modeled on the Klein geometry (G,H) with Lie algebras (g,i)). From the right principal Η bundle Η -м> Ρ —> Μ and the action of Η on G by left multiplication, we obtain the associated right principal G bundle G —> Q —> Μ and we have the canonical inclusion Ρ С Q — Ρ Хя G sending ρ ι—► (ρ, e). We are going to show that certain Ehresmann connections on Q restrict to give all Cartan connections on P. Proposition 3.1. Let (G,H) be a Klein geometry and let Ρ and Q be principal Η and G bundles, respectively, over a manifold M. Assume that dim G = dim Ρ and that φ: Ρ —> Q is an Η bundle map. Then the correspondence (Ehresmann connections λ m ( Cartan on Q whose kernels do > -^-> < connections not meet φ*(Τ(Ρ)) J | onP is a bisection of sets. Proof. The correspondence w ι—> ω — φ*ζυ. We must show that ω is a Cartan connection. Since φ*(Τ(Ρ)) Πker w — 0, it follows that ω — φ*νσ is a g-valued 1-form on Ρ with no kernel. First we verify condition 5.3.1(c)(i). Since dim Ρ = dim g and ωρ: Tp(P) —» g is injective, it follows that it is an isomorphism. Next we verify condition 5.3.1(c)(iii). Since φ: Ρ —» Q is an Η bundle map, it follows from Exercise 3.2.15(b) that for all ν G I), the vector fields
366 Appendix A. Ehresmann Connections X = z/t on Ρ (calculated via the right Η action on P) and Υ — ι/t on Q (calculated via the right G action on Q) are φ related (i.e., φ*ρ(Χρ) — Υψ(ρ) for all ρ e P). Thus ω(Χρ) — φ*π(Χρ) = τσ(φ*(Χρ)) = τσ(Υφ^) — ν for all ν G 0. Finally, we verify condition 5.3.1(c)(ii). We have Κ\ω = R*htp*w = (tpRh)*w = (Rhip)*w = φ* R*hw = tp*Ad(h-l)w = Αά(Η~ι)φ*ω. Thus, ω — φ*τυ is indeed a Cartan connection. Tfte correspondence ω *-> π = j(u). Given a Cartan connection ω on P, we may extend it to a form w — j(u) on Ρ χ G by means of the formula G7(PiP) = Ad(#_1)7r*u; + 7г£и;с, where πρ: Ρ χ G —> Ρ and πο'· Ρ x G —> G are the canonical projections. We are going to show first that w is the pull-up of an Ehresmann connection on Ρ χ# G whose kernel does not meet ф*(Т(Р)). This will show that j is a map (Cartan λ ( Ehresmann connections Ϊ connections > -?-> < on Q whose kernels do > , on Ρ J [ not meet φ*(Τ(Ρ)) J which means that it is a possible inverse for the map φ*. Note that the form w satisfies the condition νσ(0 χ z/t) — ν for all i/Gg. This verifies condition 6.2.4(H) for (the pull-up to Ρ χ G of) an Ehresmann connection on Ρ χ я G. We also see that w restricts to ω on Ρ χ e. In particular, zu does not vanish on T(P χ e). Now we study the effect on w of right multiplication by 7 Ε G on the second factor of Ρ χ G, which we denote by id χ Αγ. Under this map, zu pulls back by (id χ ^7)*^(р,р7) = ^7(р,<гу) ° (id x Ay)* = (Adtei)-1^ + π£ωσ) ο (id χ Д7)* = Ad(#7)_1u;^p* о (id χ Αγ)*) -h^ciTTG* ° (id x Ay)*) = Ad(#7)-1)u;^p*) + ^с(Ау* ottg*) = Ad(^7)_1^a; + Ad^-1)^^*) = Ad(7)-1(AdЫ-1π;u; + π^G) = Ad(7)-1n7. This verifies condition 6.2.4(i) for (the pull-up to Ρ χ G of) an Ehresmann connection on Ρ χ я G. Finally, let us verify that w is basic for the Η bundle map Ρ χ G —> Ρ xя G (i.e., that it is the pull-up of a form on Ρ Хн G). By Lemma 1.5.21, it suffices to show that w is invariant under the maps
§3. Ehresmann Connections Generalize Cartan Connections 367 / ah: PxG-> PxG (P,9) ·-»· (ph,h-lg) and vanishes in the fiber directions of PxG —> Ρ Хн G. For the former condition, we calculate (ahw){p,g) = w{ph,h-*g) ° <*h* = Ad(h~lg)~ln*Puj о ah* + π*0ωο ο ah* = Ad(/i_1#)_1u; ο πρ* ο αΛ# + cjg ο πο* ο αΛ# = Aa(h~lg)~luo Rh* ο πΡ* + uq ° Lh-i* ofG, = Ad(/i~1^)~1Ad(/i~1V ο πΡ* +wGo πα* = Аа(д)~1прш + тг^с = ^(P,<?)· Now consider the latter condition. Fix ν G \) С g. Let ι/t be the vector on PxG corresponding to ν with respect to the right Η action on Ρ χ G given by (^W) = ((a*p x /^G)o(id χ id x 6 x id) о (id χ Δ χ id) о p)^^(0,0, ι/) = (μρ χ μ^)* ο (id χ id χ t x id)* о (id χ Δ χ id)* о р*(р,д,е)(0,0, ν) = (μρ х μσ)* ο (id χ id χ 6 χ id)*^^^^)(0,г^,ζ^,0) = (μρ χ μσ)*(Ρ,β,β,ρ)(0,ι/,-^,0) = (μρ*(Ρ,β)(0,ι/),μσ*(β,ρ)(-^0)) = (u;-1Mp,-w-1(Ad(^)i/)!l), so that G7(p,p)(I/t)=a7(p,p)(w"1(I/)»-wG1(Ad(^-1)i/)) = Αά{3-ι)π*Ρω{ω~ι{ν), -^(Adfo"») = Ad(<T>-(Ad(<T» = 0. Since the vector fields z/t span the space of fiber directions, this finishes the proof of the second condition. Thus, w does pass down to yield a g-valued form on Ρ Хн G. The resulting form obviously satisfies conditions (i) and (ii) for an Ehresmann connection as well as the condition that ker w meets φ*{Τ{Ρ)) only in zero. The correspondences φ* and j are inverse to each other. To see this, we first calculate
368 Appendix A. Ehresmann Connections = Αά(β~ι)φ*π*Ρωρ + φ*π^ω0β = (πρ οφ)*ωρ + 0 showing that the composite φ* j is the identity. On the other hand, φ* is injective, for if φ*(πι) — φ*(τυ2), then w\ and ΖΣ72 agree on the image φ*(Τ(Ρ)) and hence on all the right translates Rg*4>*{T{P)). But they also agree on the vector fields ι/t for all i/Gg, and these two kinds of vectors span the whole of T(Q). Thus, φ* and j are inverse to each other. ■ Two Notions of Development Suppose the manifold Μ is equipped with a Cartan geometry modeled on the Klein pair (G,H). Let σ: 7 —» Μ be a path on M. In §4 of Chapter 5, we showed how to develop the path σ to a path σ: 7 —» С/Я" on the model space G//Z". The same methods showed how to associate to a loop σ on Μ the (Cartan) holonomy of σ which is an element of H. For an Ehresmann connection on a principal bundle over Μ, there is, in general, no analog of the development of a path on M. However, there is always an analog of the holonomy of a loop. It is this concept that we study here. Our study depends on the following generalization of the notion of horizontal vectors given in Definition 5.3.44. Definition 3.2. Let Η —» Ρ —> Μ be a principal Η bundle over Μ, let i) be the Lie algebra of 77, and let 7 be an Ehresmann connection on P. A vector ν G TP(P) is horizontal if 7(1/) = 0. A path σ: 7 —» Ρ is horizontal if cr*7 = 0 (i.e., the tangent vectors of σ are all horizontal). * Exercise 3.3.* Let Μ be equipped with a principal bundle Q and an Ehresmann connection on it. (i) Show that the map π*ρ:Τρ(Ρ) —» X^M) induced by the projection π: Ρ —> Μ restricts to give an isomorphism between the horizontal vectors at ρ e Ρ and TX(M) (cf. Lemma 5.3.45). (ii) Fix xq £ Μ and fix a point qo G Q lying over it. Show that every path on Μ starting at xq is covered by a unique horizontal path on Q starting at qo (cf. the proof of Proposition B.2.3). □ Definition 3.4. Let Q —» Μ be a principal G bundle equipped with an Ehresmann connection and let c: (7,97) —» (M, xo) be a closed path on Μ. Fix qo € Q lying over xq? and let c: (7,0,1) —» (Q,#o»(7i) be the unique horizontal lift of с starting at qo. Since qo and #ι lie in the same fiber of
§3. Ehresmann Connections Generalize Cartan Connections 369 Q, we may write qo — q\g for some g G G. The (Ehresmann) holonomy of с with respect to qo is д. Ф The behavior of this notion of holonomy is somewhat different than that of the notion of Cartan geometries given in Definition 5.4.18. Here is a table comparing these behaviors of holonomies of loops on M. Comparison of Holonomy of Loops Cartan Holonomy Ehresmann Holonomy • Has values in G · Has values in Η • Varies by conjugacy according · Varies by conjugacy according to the starting point of the lift. to the starting point of the lift. • Only applies to loops on Μ · Applies to all loops on M. that lift to loops on P. Proposition 3.5. Assume the context of Definition 3.4. If the Ehresmann holonomy of с with respect to qo is g, then the Ehresmann holonomy of с with respect to qoh is h~lgh. Proof. Let c: (7, 0,1) —> (Q, go? Qi) be the unique horizontal lift of с starting at go· We claim that the unique horizontal lift of с starting at qoh is Rh о с: (7, 0,1) —» (Q, qoh, q\h). Clearly, this is a lift of c, but it is also horizontal since (Ял о c)*7 = с*Я£7 = c*Ad(/Tx)7 = Ad(/rx)c*7 = 0 (since с is horizontal). But qoh — qigh — q\h(h~lgh), so h~lgh is the holonomy of с starting at qoh. Ш Since the Ehresmann connection may be regarded as a generalization of the Cartan connection in two distinct ways, we may compare the notions of holonomy in each of these two ways. We study first the relation between the holonomy of a Cartan connection and the Ehresmann connection derived from it via the procedure of Proposition 3.1. Proposition 3.6. Let (Ρ,ω) be a Cartan geometry on Μ modeled on (G,H) and let (Q,w) be the corresponding Ehresmann connection as given in Proposition 3.1. Fix a point xo G Μ and a loop c:(l,dl) —» (M,x0) that lifts to a loop based at qo G Po· Then the Cartan and Ehresmann holonomies of с are the same. Proof. Let c: (7,0,1) —» (P Хя G, (po,e), (po,g)) be any horizontal lift of с and let с = (ci,c2): (7,0,1) —► (PxG,(p0,e),(po^)) be a lift of с to Ρ χ G. By Definition 3.4, the Ehresmann holonomy of с is g~l. On the other hand, according to the proof of Proposition 3.1,
370 Appendix A. Ehresmann Connections 0 = c*W(p^g) =Ф- 0 = Ad(# )с π*Ρω -f с Kq — Ad(c^~ )c\u + c^cjg, so that by Proposition 3.4.10(ii) c[u = -Ad(c2)c^o;G = (c^1)*^· Since ci is a lift to Ρ of c, it follows (cf. Definition 3.7.2) that c^1 is the development of с in the Cartan sense, so that the Cartan holonomy of с is c2_1(l) — 9~1· Thus, the two holonomies are the same. ■ Finally, we study a simple case of the relation between the holonomy of a Cartan connection and an Ehresmann connection when the two connections are related as in §2. Proposition 3.7. Let (Ρ,ω) be such a Cartan geometry on Μ modeled on (G = Η xp L,H), where Η xp L is a semidirect product (with L the normal subgroup). Fix a point xq G Μ and a loop c: (7, dl) —» (Μ,χο) that lifts to a loop based at qo G Лэ (tyin9 over x$). Let g, \), and ρ be the Lie algebras of G, Η', and L, respectively. Decompose the Cartan connections as ω = ω^+ωρ, so that (by Lemma 2.1) ω^ is an Ehresmann connection on P. Then the Ehresmann holonomy of с is the Η component of the Cartan holonomy of c. Proof. On the one hand, we may fix a lift c: (7,0,1) —» (P, qo,qo) of с to a closed loop on P. On the other hand, by Exercise 3.3(ii), the loop с also has a unique horizontal lift C2: (7,0,1) —» (Q,qo,qi) starting at q^. Because c\ and c2 are both lifts of the same loop c, there is a unique map h: (7,0) -» (#,e) such that c2(t) = cx(t)h(t). Let h(t) = h(l - t). We have Qi — c2(l) = Ci(l)ft(l) = qoh(l). Thus, h(l)~l is the Ehresmann holonomy of с based at qo- On the other hand, the Cartan holonomy of с is obtained as the endpoint of the development, starting at the identity, of any loop on Ρ based at qo and covering the loop c; for instance, the loop (q^h) *C2. The development of this loop is obtained (cf. Exercise 3.7.5(b)) by path multiplication of the development of c2 with the development of q^h. Now c^u — (%ωρ (here we use the fact that с^иц — О, which holds since c2 is a horizontal curve). Thus, c^ui is a form with values in ρ С 0, so its
§4. Covariant Derivative 371 development on G, starting &texeeHxpL, lies entirely on the subgroup L and hence ends at an element of the form e χ I G Η xp L. On the other hand, the curve q^h on Ρ develops (cf. Lemma 5.4.12) to the curve, starting 8LtexeeHxpL, given by h(l)~lh and hence it also develops to the curve, starting at e χ I G Η xpL, given by (e χ l)h(l)~lh. Thus, the development of (qoh) * C2, starting at e x e G Я xp L, ends at (e χ l)h(l)-lh(l) = h(l)~l χ p(h(l)){l). This is the С art an holonomy and its Η component is the Ehresmann holon- omy. ■ It is interesting to compare this result to the case of rolling without slipping or twisting described in Appendix B. For a 2-sphere rolling on a plane along a closed loop on the sphere, the initial and final points of contact are the same on the sphere but differ by a translation on the plane. The total effect on the sphere is to undergo a two-dimensional rotation and translation. The rotation part is the Ehresmann holonomy of the loop on the sphere, whereas the rotation and translation are the Cartan holonomy of the loop on the sphere. §4. Covariant Derivative In Chapter 5 we discussed the covariant derivative associated to a reductive Cartan geometry (cf. Definition 5.3.46 through Exercise 5.3.52). In this section we show this notion generalizes to arbitrary Ehresmann connections (Proposition 4.4). As in Definition 5.3.46, the definition of covariant derivative depends on the notion of a horizontal vector given, in the present context, in Definition 3.2. Definition 4.1. Let (i) G —» Q —» Μ be a principal bundle over Μ, (ii) 7 be an Ehresmann connection on Q, (iii) Ε be the vector bundle associated to a representation (V, p) of G (i.e., E = Q xGl/), and (iv) У be a tangent vector field on Μ and Υ its horizontal lift to Q. The covariant derivative Όγ:Γ(Ε) —> Γ (Ε) associated to these data is defined by the following diagram.
372 Appendix A. Ehresmann Connections A\Q,p)-Ϊ-+A\Q,p) ψ ψ DYX Γ(Ε) > Γ(Ε) (Cf. Exercise 1.3.28 for a description of the isomorphism φ.) φ In preparation for the calculation given in Proposition B.3.7, it is useful to reformulate the definition of covariant derivative for the case where Q can be interpreted as a frame bundle of E. Proposition 4.2. When the representation (V, p) in Definition 4.1 is faithful, then (i) given a fixed basis ё\,..., e~v for V, there is a canonical interpretation of Q as the bundle of frames of Ε given by qeQ^ (ei (<?),..., e„(<7)), where ei(q) = (q,ei) eQ xGV = E, (ii) the covariant derivative ΏγΧ, where X G T(E), may be calculated as follows. Express the vector field X in terms of the basis ei(q) as X — ai(q)ei(q) Η + a„(q)ej;(q). Let Yq — Tq(Q) denote the horizontal lift ofYx. Then (DYX)X = Yq(ai)ei(q) + · · · + Yq{av)ev{q). Proof, (i) For this we refer the reader to the discussion on page 37. (ii) Define /: Q —► V by f(q) = Ei<j<i/ai(i)^'· We claim that / G A°(Q,p). To see this, first note that ej(qh) = (qh, ej) e Q xG V = (q,p(h)ej) = Y^p(h)jk(q,ek). Thus, Σ аз(я)ео(я) = x = Σ aj(Qh)ej(qh) = ^ aj(qh)p(h)jkek{q) \<j<v l<j<v l<J,fc<^ and hence ak{q) = Σ aMh)p(h)jk, l<j<v or аз{Ф) = Σ ak(Q)p(h~l)kj. \<k<v
§4. Covariant Derivative 373 It follows that f(Qh) = Σ aMh)ej = Σ ak(q)p(h~l)kjej l<j<v l<3,k<v> = p(h~1) Σ MQ)ek = p(h-1)f(q). \<j,k<V Next we claim that ψ(/) — X, which follows from the fact that ψ(/) £ T(E) — T(Q xgV) is induced by the map Q —» Q χ V sending q^(qJ(q))= U, Σ аЛя)ёз\ = Σ азШ^ёз) \ 1<3<" J 1<3<" l<j<v Now, according to Definition 4.1 we have (ΌγΧ)χ — ψ(Ϋς(/)), and the latter is induced by the map Q —> Q χ V sending Q ·-> (Q,Yq(f)) =\Q>Yq[ Σ aJeJ )) = \q> Σ ^я(аз)ёз We note that in the case described in Proposition 4.2, it is clear that the covariant derivative (ΌγΧ)χ depends only on the value of У at x. This is also true in general. Exercise 4.3. (i) Verify that the properties of Proposition 5.3.48 remain true for the covariant derivative of Definition 4.1. (ii) Verify that, for each χ G Μ, the vector (ΌγΧ)χ G Ex (= the fiber of Ε over x) depends only on the value of У at χ and the values of X along any curve in Μ tangent to Υ at X. (iii) Strengthen the result of (ii) by showing that the dependence on the curve is actually only a dependence on the first two terms of the Taylor series5 for the curve at x. □ Proposition 4.4. The covariant derivative associated to a reductive Car- tan geometry is the same as the covariant derivative associated to the Ehres- mann geometry obtained from it by the procedure of §2. Proof. Let us decompose the Cartan connection as ω — ω^ -f ωρ as in Lemma 2.1. Then ω^ is an Ehresmann connection, and the procedure for associating a covariant derivative to it described in Definitions 5.3.46 and 5.3.47 is identical to that of Definitions 3.2 and 4.1. ■ 5 In any coordinate system.
Appendix В Rolling Without Slipping or Twisting In this appendix we study the most classical of all nonholonomic differential systems, the system describing how one manifold may roll without slipping or twisting on another in Euclidean space. A sphere rolling along the central line of a helicoid without slipping or twisting. Such a system plays a role, for instance, in robotics, where one wishes to use robotic fingers to manipulate some object. Our interest, however, is to
376 Appendix B. Rolling Without Slipping or Twisting show the relationship of this concrete physical notion to the idea of Levi- Civita connection of a Riemannian manifold and the canonical Ehresmann connection on its normal bundle in Euclidean space.1 In §1 we give a definition of the notion of rolling without slipping or twisting in an "elementary" form which appeals directly to the intuition in Euclidean space. In §2 we reformulate this definition in terms of a differential sysem (i.e., a distribution) on a state space. In this formulation we easily obtain the existence and uniqueness of rolling without slipping or twisting. In particular, this yields isometries between the ambient tangent spaces at each point of the rolling. In §3 we show how this notion gives rise to both the Levi-Civita connection and the normal connection for a submanifold of Euclidean space. The section ends by relating the notion of rolling a manifold Mn on Rn to the notion of development described in Chapter 5. In particular, we show that a curve on Μ rolls without slipping on a straight line in Rn if and only if the curve is a geodesic. We also show that a vector field along a curve in Μ is parallel if and only if its image in Rn is parallel. This result (Proposition 3.5) makes the notion of the Levi-Civita connection and the normal connection for a submanifold of Euclidean space quite concrete. Finally, in §4, we study what happens when we roll one manifold on a second while the second is itself rolling on a third. §1. Rolling Maps In order to study rolling maps we need to have a way to represent elements of Еисдг(К), the group of isometries of Euclidean TV space. If g G Eucat(R), then we may write g:Kn -> Rn, ν ι—► Av+p where A G On(R) and ρ G R^. It is easy to verify that the map Euciv(R) -> | (I JJg GlN+i{R) | A G On(R),p G Rn| is an isomorphism. A rolling map will be a certain kind of one-parameter family g: I —> Eucat(R). We write g(t)v = A(t)v +p{i). For fixed t G /, we have g(t): KN -» RN with derivative g(t)* = A(t): KN -» RN, here "*" refers to the partial derivative with respect to the space coordinates. We write the ^oth of these connections are described in Chapter 6. See also [C. Dodson and T. Poston, 1991], pp. 207ff., for a treatment of the tangential part of this result. We also refer the reader to [R.L. Bryant and L. Hsu, 1993], p. 456, for a more advanced study of aspects of this kind of differential system.
§1. Rolling Maps 377 partial derivative of g(t) with respect to the parameter t as g(t): Rn —» Rn. In particular, the composite mapping g(t)g(t)~l:HlN —» RN makes sense. The following definition formalizes the intuitive notion of one manifold rolling on another in Euclidean space without slipping or twisting. The subsequent interpretation is meant to explain the sense in which it does so. Definition 1.1. Let M£, M? С RN (where TV = η + r) be submanifolds. Then a map g: I —> Еисдг(К) satisfying the following properties for each t G / is called a rolling of Μχ on Mq without slipping or twisting. More briefly, we shall call g a rolling map. The "rolling" condition: (1) There is a piecewise smooth "rolling curve on Μι" σχ:Ι —* Mi such that (a) g(t)ai(t) Ε Μ0, and (b) TgdwMtWi) = Tg(t)triW(M0) for all t. The curve gq'.I —» M0 defined by σο(£) = g(t)o~i(t) is called the development1 of σ\ on Mq . The "no-slip" condition: (2) g{t)g{t)-lvo{t) = о for ail t. The tangential part of the "no-twist" condition: (3) {g{t)9{t)-l)*TMt){M0) С Τσο(ί)(Μ0)χ for all t. The normal part of the "no-twist" condition: (4) {g(t)g{t)-%TMt){M^ С Τσο(£)(Μ0) for all t. * Interpretation Condition (1) says that M\ moves under the action of g(t) so as to be tangent to Mo at the point ao(t) — g(t)a\(t) at time t. Condition (2) says that the infinitesimal isometry g(t)g(t)~l fixes σο(£)> namely, that it is an infinitesimal rotation about the point ao(t). This is the "no-slip" condition. Next we consider conditions (3) and (4). Each tangent vector w G Tq(Mi) (respectively, normal vector w G Tq(M\)L), as it is carried along by the motion g(t), determines a one-parameter family of vectors wt = g(t)*w e Tg(t)q(g(t)Mi) (respectively, Tg^t)q(g(t)Mi)L) with velocity wt. Now suppose ν e Τρ(ίο)σι(ίο)(^(ί0)Μι) (= Τσο(£ο)(Μ0)). Setting w = g(to)^lv, we have υ = wto — g(to)*w and 2See §5 for the relationship of this to the definition of development given in §4 of Chapter 5.
378 Appendix B. Rolling Without Slipping or Twisting ™t\t=to = g(t0)*w = g(t0)*g(t0)*lv = (g(t0)g(to)~l)*v. Thus, condition (3) says that the motion of a vector ν e Tg^(Tl^to)(g(to)Mi) (= Τσο(ίο)(Μο)) that is "stuck to g(t)Mi" has no component of velocity in the tangential direction. This is the tangential part of the "no-twist" condition. Condition (4) says that a vector ν e Τσ^^(ί)Μι)Σ, as it is carried along by g(t), has no component of velocity in the normal direction. This is the normal part of the "no twist" condition. §2. The Existence and Uniqueness of Rolling Maps The interpretation given in §1 makes it clear that the conditions of Definition 1.1 should be required of a rolling map. What is not clear is the existence and uniqueness, for each curve σ(ί), of a rolling map with a(t) as its rolling curve. The most convenient3 way to deal with the existence and uniqueness question is to reformulate the equations of Definition 1.1 as a differential system, that is, as an η-dimensional distribution on a certain configuration space Σ. Roughly speaking, the configuration space Σ is the space of all positions of Mi in which it is tangent to Mq. More precisely, we define the configuration space to be Σ = {(p,A,q) G M0 x Олг(К) х Мг \ ATq(Mx) = Tp(M0)} (where we are identifying TP(M0) and Tq{M\) with subspaces of KN in the usual way). Why should Σ be regarded as the configuration space? First note that there is a canonical smooth map φ: Σ ->Eucn(R), (p,A,q) н-> д where g = φ(ρ, A, q) is defined by g:Rn-> Rn. χ »—► A(x — q)+p Since gq = ρ and g* — A, it follows that g*Tq(Mi) — Tp(Mq). Thus we see that a point in Σ determines a point ρ G Mq and an isometry g G Еисдг(К) such that gM\ is tangent to Mq at ρ = gq. This explains the sense in which Σ may be called the configuration space of all positions of M\ in which it is tangent to M0. 3This is instructive as well as convenient, as it introduces the reader to a beautiful application of differential systems.
§2. The Existence and Uniqueness of Rolling Maps 379 The first order of business is to determine the tangent bundle of Σ. It is not difficult to verify that the canonical projection Σ —» Mq χ Μχ is a fiber bundle with fibers On(H) χ Or(R). It follows that Σ is a manifold of dimension 2n + \n{n — 1) + ^r(r — 1). We shall also need to use the second fundamental form as it is described in Exercise 1.2.24. In particular, we shall interpret the second fundamental form as having values in the normal bundle of the manifold on which it is defined. Proposition 2.1. Let Bo and Βχ denote the second fundamental forms for M0 and Mi, respectively. Let V С TP(M0) x Ta(On(R)) χ Tp(M0) be given by V={(p, A, q) | AA~lu = ΑΒχ(</, Au) - B0(p, u) mod Tp(M0), Vu e TP(M0)}. ThenT(PtAt4)(E) = V. Proof. We know that (p, Л, q) G T(p5α,9)(Σ) if and only if it is the derivative at (p,A,q) of a curve (p(t), A(t),q(t)) on Σ. Let u\{t) be a vector field tangent to Mi defined along q(t). Then, by the definition of Σ, the vector field uo(t) = A(t)ui(t) is tangent to Mq alongp(t). Differentiating produces u0(t) = A(t)u1(t) + A(t)ui(t). By Exercise 1.2.24, we have uo(t) = -В0(р,и0)тоаТр(М0) and ui(t) = -Bi(q,ui)modTq(Mi). Thus, at (p, Л, q) we have -B0(p,u0) = Aui - ABi(q,ui)modTp(M0), or -B0(p, u) = AA~lu - ABx(q, A~lu) mod TP(M0) for all и G TP(M). (2.2) Thus, Т(рЛд)(Е)сУ. On the other hand, we claim that dim V < 2n + ^n(n — 1) -l· |r(r — 1), from which the proposition follows. To see this, we note that AA~1 is skew symmetric (since AAl = I => ААг+АА* = 0 =ϊ АА~1 + (АА~1У = 0). If we express the skew-symmetric matrix AA~l with respect to an orthonormal basis of ΈΙΝ that is adapted to Tq(Mi) (i.e., the first η basis elements span Tg(Mi), which implies that the last r basis elements span Tq(Mi)L), then Eq. (2.2) determines the (2,1) block of AA~1 in terms of ρ and <j, so we have (in the given basis) Aa~1 = I unknown η χ η \ ( known \ skew-symmetric matrix J у by symmetry J /i \ ( unknown r χ r ' I skew-symmetric matrix
380 Appendix B. Rolling Without Slipping or Twisting Thus, we see that specifying the variables ρ G Tp(Mo) and q G Tq(Mi), each with at most η degrees of freedom, forces AA~l to lie in a subspace of dimension \n{n — 1) -f \τ{τ — 1). It follows that we have the inequality dim V < 2n + -n(n - 1) + -r(r - 1). ■ A smooth curve c: I —» Σ determines a smooth curve фос: I —» Еисдг(К). The next order of business is to determine the conditions on с that will ensure that φ о с is a rolling map. Lemma 2.3. фос is a rolling map if and only if с is tangent to the n- dimensional distribution Q on Σ given by (i) ρ = Aq, (ii) AA~lu = ABi(A~lp, A~lu) - B0(p,u) for allue TP(M0), (iii) AA~lu = -AB[{A-lp,A-lu) + B$(p,i/) for all ν e TP{M0)L. {See Exercise 1.2.24(iii) for the definition of Вг.) Moreover, the derivative of the canonical map Σ —» Μ induces the isomorphism shown in the following diagram. \/ TPiM) Proof. Let c(t) be any curve on Σ. We write с = (ρ, A, q) and с = (ρ, A, q) and we try to express conditions (2), (3), and (4) of Definition 1.1 in terms of c. (Condition (1) is, of course, built into the very definiton of Σ.) We have gv = A(v — q) +p so gv — A(v — q)+p — Aq and g~lv = A~l(v -p) + q, so that gg~lv = А(А~1(у -p))-\-p-Aq and (gg~l)*v = AA~lv. Since gg~la — gg~lp — ρ — Aq, it follows that condition 1.1(2), namely, gg~la = 0, is equivalent to ρ = Aq. Now (gg~1)* — AA~l implies that condition 1.1(3), namely,
§2. The Existence and Uniqueness of Rolling Maps 381 (^-1).T(7o(t)(Mo)cT(ro(t)(Mo)x (and, respectively, condition 1.1(4), i.e., (дд~1)*Тр(М0)1- С TP(M0)), is equivalent to AA~1Tp(M0) С Tp(M0)L (respectively, AA~1Tp(Mq)l с Tp(Mo)), which by Proposition 2.1, is equivalent to the equation А А"1 и = ΑΒι(Α~ιρ, A~lu) — B0(p,u) for all и G TP(M0) (respectively, to the equation AA~lv = -AB\(A~lp,A~lv) + Blfav) for all υ e TP(M0)L). Of course, by Exercise 1.2.24, this latter equation is always true mod Tp(M0) (respectively, mod Tp(M$)L), and the right-hand side always takes values in the normal bundle (respectively, the tangent bundle) of M$. But here the left side also lies in Tp(Mo)1- (respectively, mod TP(M0)), so we get equality. Thus, φ о с is a rolling map if and only if с is tangent to the distribution To calculate the dimension of ^, we again fix, as in the proof of Proposition 2.1, an orthonormal basis of Rn+r that is adapted to Tq(M\) and consider the matrix AA~l which is skew symmetric in this basis. By Definition 1.1(3) and 1.1(4), AA~l must, in this basis, have the form (s ?)■ Thus AA~l is completely determined by its value on Tp(Mo), which is known by equation 2.3(H) at the point (p, A, q) once either ρ or q is known. By equation 2.3(i), q determines p, so it follows that dim ^ < n. On the other hand, for any choice of q G Tq(Mi) equations (i), (ii), and (iii) can be solved for ρ and A so that dim ^ = n. Finally, this same argument shows that the canonical projection π: Σ —» Μ inducing the map π*(ρ,Α,9):Τ(ρ,Α9)(Σ) -* Τρ(μ) nas tne property that 7Γ#(ρ>Λ>ς) | ^ is an isomorphism. ■ Now we are in a position to obtain the existence and uniqueness of rolling maps. Proposition 2.4. Let Μβ,Μ™ С Rn+r be submanifolds and let (po, A), <7o) G Σ. Assume we are given a piecewise smooth curve σι: (7,0) —» (Mi,go) (respectively, σο'(1,0) —» (Μο,Ρο))· Then there is a unique rolling map g:(I,0) —» (EuCn(R.), id) with rolling curve σι (respectively, development Proof. We may as well assume that the curve is smooth since we can patch the smooth pieces together to obtain the general case. We will deal only with the case of σχ: (1,0) —» (Mi,go)? as the other case is similar in the configuration space setting we are using. Pulling back the bundle M0 x On(R) χ Or(R) -» Σ -» Μι
382 Appendix B. Rolling Without Slipping or Twisting over I along σι yields a bundle M0 x 50n(R) χ SOr(K) -> Σσι -> 7 on which the restriction !^7l of the distribution ^ has dimension 1. Since a one-dimensional distribution is always integrable, there exists a unique curve in Σσι through (po> Abtfo) £ Σσι which is tangent to ^, The image of this curve in Σ is tangent to ^ and covers σ\. By Lemma 2.3, this proves the existence and uniqueness of the rolling map given in Definition 1.1. §3. Relation to Levi-Civita and Normal Connections In this section we relate our discussion of "rolling without slipping or twisting" to the Levi-Civita connection and to the Ehresmann connection on the normal bundle. Some of this material necessarily covers ground similar to that of §6.5. We consider the case of the standard η-plane Rn (= Μχ in the notation above) С HN rolling on a manifold Mn (= M0 in the notation above) С ΈΙΝ. Then Σ = {(p,A,q) eMx Одг(К) χ Rn | ARn = TP{M)} is the configuration space for this pair of manifolds, and we again have the distribution ^ on it described in Lemma 2.3. We are going to study the following diagram of fiber bundles. The three spaces -Ftan? ^nor and PT are the bundles over Μ that we have already considered in §6.5. We are going to describe the images of ^ under these maps and show in particular that its images in T(Pt3in) and T(Pnor) are (the kernels of) the Levi-Civita and the normal connections. Since, by Exercise A. 1.3, these kernels determine the connections, we see that for a manifold in Euclidean space the notion of rolling without slipping or twisting includes the notions of the Levi-Civita connection on the tangent bundle and the Ehresmann connection on the normal bundle. We define ΐ8θ(Κ*,ΚΝ) = {φ e Нот(К*,К") Ι (φ[μ),φ[υ)) = {u,v)\/u,v e Rk} and set
§3. Relation to Levi-Civita and Normal Connections 383 PT = {(p,A) eMx On(K) I A(Kn) = rp(Af)}, ^tan = {(ρ,φ) eMx Ъю(Кп,К") I φ(ΈΙη) = Tp(M)}, Pnor = {(ρ,φ) eMx Ъю(Кг,К") I ^(Rr) = Tp(M)L]. Lemma 3.1. The map PT —> Μ (respectively, Ptan —> M, Pnor —» M) induced by canonical projection is a principal On(R) x Or(R) (respectively, On(R), Or(R)) bundle map. Proof. We deal with the map PT —> Μ only, the others being similar. Consider the right action Ρτ χ On(R) x Or(R) -> PT. This action is free and proper (note that On(R) x Or(R) is compact) and acts transitively on the fibers of the projection PT —» Μ. It follows from Theorem 4.2.4 that PT —»> Μ is a principal bundle with group On(H) x Or(R). ■ Exercise 3.2. Verify that Ptan and Pnor are the tangent and normal or- thonormal frame bundles associated to the inclusion Μ С ΈΙΝ as in §6.5. Show also that PT corresponds to the bundle of the same name in §6.5. О Lemma 3.3. For ρ e Μ let V С TP(RN) χ Τα(Ον(Κ)) be defined by (p, A) G V <=ϊ ρ and A satisfy the following conditions (i) AA~lu = -B(p,u) mod TP(M) for all и е TP(M). (ii) AA~lv = Bl(p,v) mod TP(M)L for alive TP(M)L. ThenT{vA)(PT)GV. Proof. Let (p,A) G PT, and consider the vector space V. In an orthonor- mal basis of TP(HN) adapted to TP(M) (cf. the proof of Proposition 2.1), conditions (i) and (ii) imply that the matrix AA~l has the block form a a-i _ ( arbitrary determined by ρ у determined by ρ arbitrary In particular, the dimension of V is dim Μ + dim On(R) + dim Or(R) = dim PT. Thus, it suffices to verify the inclusion T^p^(PT) С V. Suppose that (p,A) e T(P)i4)(PT). Let (p(t),A(t)) be a curve on PT with tangent (p,A) at
384 Appendix B. Rolling Without Slipping or Twisting (p, A). Let u{t) be a tangent field (respectively, normal field) to Μ along the curvep(t). Then A~l(t)u(t) e RnxO (respectively, OxRr). Differentiating, we get -A-l(t)A(t)A-l(t)u(t) + A~l(t)u(t) e Rn x 0 respectively, 0 χ Rr) or -A(t)A-l(t)u(t) + ώ(ί) e TP(M) (respectively, TP{M)L)). But we know from Exercise 1.2.24(i) (respectively, (Hi)) that u{t) = -B(p(t), u{i)) mod TP(M) respectively, u(t) = JBi(p(i),tx(i))modTp(M)-L), so A(t)A-\t)u(t) + B(p(t)Mt))tTp(M) respectively, A(t)A-l(t)u(t) - в*(р(0,и(0) е Tp(M)L). This verifies the inclusion C. ■ Exercise 3.4. Show that T{p^(Ptan) = {(ρ,Φ) e TP(M) χ Hom(Rn,R^) | φ = -Β(ρ,φ(—)) moaTp(M)}, T(p^)(Pnor) = {faΦ) Ε ΤΡ(Μ) χ Hom(Rn,R") Ι ф = -Β*{ρΜ—)) modTp(M)±}. ■ It is convenient now to regard the group of Euclidean isometries Eucat(R) as the matrix group (with lxl and Ν χ Ν blocks down the diagonal in the block form) EuCiV(R) = i(l Jje GlN+l(R) | ρ e RN, A e On(R)\ with Lie algebra cncN(K) = U°p γ\ EMN+1(R)|pERN, ieo;v(R)}. In this picture, the isometry given by gv = Αυ+ ρ corresponds to the matrix ί .1, and the (left invariant) Maurer-Cartan form on Eucat(R) may be expressed as
§3. Relation to Levi-Civita and Normal Connections 385 ι o\_1/o o\ / ι о \ /о о ^EucN(R)-lp Aj yp AJ-\-A~lp A~lJ\p A 0 0 \ A~lp A~lAJ' Let us express the Maurer-Cartan form on Еисдг(К) (with Ιχΐ,ηχη, and r χ r blocks down the diagonal in the block form) as ^Eucn+r(R) Then we have Proposition 3.5. The inclusion Μ С Rn+r is covered by the right On(R) χ Or(R) equivariant map l:Pt С Eucn+r(R). Let us define Η = ker(a | Ρτ) Π ker(6 \ PT). Then, for each (p,A,q) e Σ, the derivative of the canonical map Σ —» PT induces the isomorphism in the following diagram. Т(рЛя)(Е) z> %jpA4) Proof. Since ('■"ь'Еисл, (R)) (ρ,Α) (Ρ,Α) = (wBuca,(R))/1 0\(».(РД))=(Д А-1 а)' x: :r using the block form of the Maurer-Cartan form, we have 1λ_(α(ρ,Α) -β* (ρ, A) Λ-\β(ρ,Α) Ί(ρ,Α) The proof of Lemma 3.1 shows that Рт С Eucn+r(R) is a right On(R) x Or(R) equivariant map. Now the derivative of the map Σ —» PT is given by (p, Л, q) ι-* (p, A), and so, according to the description of ^ in Lemma 2.3, (and in the present case, B0 = В and Βλ = 0) the image of ^p,A,q) under this map is W(p,a) С ТР(М) х Тд(Оп(К)) where (ρ, A) e W(PiA) & the following two conditions hold:
386 Appendix B. Rolling Without Slipping or Twisting (i) А А'1 и = -B(p,u), for all и G TP(M) (ii) AA~lv = -B»(p,i/), for all ν G TP{M)^. Thus, / · k\ c w _ / ^4-%(M) С Гр(М)-1-, and Г A-lA(Rn χ 0) с 0 χ Rr, and t^-^ioxR7·) cR"xO a(p,A) = 0, and .6faA) = 0 =>(p,A)€!H(PtA). {: Thus W(p^a) С ^(ρ,Α)· On the other hand, by Lemma 2.3, the composite %(p,A,q) —* Τ(ρ,Α)(Ρτ) —» TP(M) is an isomorphism, so that ^р>л,д) —* T(p ^4)(PT) is injective. Since PT is a principal On(R) x Or(R) bundle over Μ and α Θ 7 restricts to the fiber of PT to yield the Maurer-Cartan form on On(R) x Or(R), it follows that dim Η = η = dim ^, and so the image of %,A,g) is%,A)· ■ Corollary 3.6. For each (p,A) G PT, the canonical projections PT —» Ргап and PT —» Pnor induces the isomorphisms in the following diagram. Uplift 3 ^ T^tf) 3 ^ ^(ρ,ΑΐΗ^ί^ιι) => ker (a)(p,A|H^ ^1^(0 ^ ^Г (y)(p^|Rr) Proof. The proofs are similar, so we consider only the projection PT —» Ptan· Now the form α on PT is basic for this projection (Proposition 6.5.12) so the distribution ker a makes sense as a distribution on Ptan· Since Η = кег(а I Рт)Г)кег(<5 | PT), the canonical projection maps 0{ to ker a. Finally, since the derivative of the composite map PT —» Ptan —* Μ is injective on #", it follows that #" —» ker α is injective, and since the dimensions are the same, it is an isomorphism. ■ Proposition 3.5 establishes the link between "rolling without slipping or twisting" and the usual connection forms on a submanifold of ΈΙΝ since we know from Proposition 6.5.12 that ker(a | Ρτ) Π ker(<5 | PT) projects to the Levi-Civita connection on Ptan and to the canonical normal connection Proposition 3.7. Let t \~* (ao(t),A(t),a\(t)) G Σ be an integral curve for the distribution H^.
§3. Relation to Levi-Civita and Normal Connections 387 (i) t ι-* σ\{ί) G Rn is the development (in the sense of Definition 5.4.15) oft ι-» a0(t) G M. (ii) If v(t) is a vector field along t ι-» ao(t) G Μ which is tangent (respectively, normal) to Μ, then DMt)v{t) = A(t)(A(t)-lv(t)), where Όσο^ν(ί) is the Levi-Civita (respectively, the canonical normal) covariant derivative. (Hi) A vector field v(t) along 11-* ao(t) G Μ which is tangent (respectively, normal) to Μ is parallel if and only if A(t)~lv(t) is constant. Proof, (i) Exercise 5.4.14 tells us how to develop a curve t ι—► ao(t) G Μ to a curve on the model space of a reductive geometry. It says we must first take the horizontal lift σ: (/, a) —» (Ptan,p), which in the present case is the projection of 11-* (ao(t),A(t),a\(t)) G Σ, to Ptan- Then we must solve the equation σ*α;ι&η = da (i.e., ut3in(a) = σ) for the development σ: (/,α) —» (Rn,0). But this equation is, in the present circumstance, the equation A(t)~la0(t) = a(t). But since, by the definition of ^, A(t)~l&0(t) = &i(t), the uniqueness of development implies a(t) = a\(t) for all t. (ii) Case (1): v(t) is a tangent vector field. According to Definition A.4.1, the covariant derivative ΌχΥ', for X G ТЖ(М) and У a tangent vector field on M, is calculated as follows: express У as a function /y: Ρ —» g/(); lift X to a horizontal vector Xp = TP(P)\ compute Xp(fY) G g/ί); and then set DxY = v-l(xp(fY))eTx(M). First consider the tangent vector field v(t) on M. It determines the function t ^ φ{σ{ί),Α{ι))(ν(ΐ)) e g/i). Next, since the curve t \~* ao(t) has the lift t \~* (ao(t),A(t),a\(t)) G Σ, which is an integral curve for the distribution ^, it follows by Corollary 3.6 that the curve t ι-* (ao(t),A(t) \ Rn χ 0) is a horizontal lift to Ptan, that is, (&o(t),A(t) | Rn χ 0)) is a horizontal vector covering &o(t). Putting this all together, we have Ψ(σ(ί)1Α(ί))(Οσο(ί)υ(ί)) = (^(σ(ί),Α(ί))(ν(ί))) = (A(t)~\v(t))) = ^{(T(t),A(t)M(t)(A(t)-l(v(t)))). Since 4>(a{t),A(t)) is injective, this verifies case (1). Case (2): v(t) is a normal vector field. This time the covariant derivative is taken with respect to the Ehresmann connection on the principal Or(R) bundle Or(R) —» Q —» Μ of normal frames on M. According to Definition A.4.1, the covariant derivative ΌχΥ, for X G TX(M) and У a normal vector field on M, may be calculated as follows: express У as a function fy.Q —» Rr (i.e., write У in terms of the basis
388 Appendix B. Rolling Without Slipping or Twisting given by the frame q £ Q); lift X to a horizontal vector Xq £ Tq(Q); compute I9(/y) G Rr. Then DXY is the normal vector to Μ which, when expressed in the basis q £ ζ), is Xq(fy). First consider the normal vector field v(t) on M. It determines the function t ь-> A{t)-lv(t) £ Rr with derivative 11-» (A(t)-lv(t)) £ Rr. Since the curve t ι—> σο(0 has the lift t ι-* (ao(t),A(t),a\(t)) £ Σ, which is an integral curve for the distribution $t, it follows by Corollary 3.6 that the curve t ι—> (ao(t),A(t) | 0 χ Rr) is a horizontal lift to Pnor> that is, for each £, (&o(t),A(t) | 0 χ Rr) is a horizontal vector covering &o(t). Putting all this together, we have DMt)v(t) = A(t)(A(t)-1(v(t))). (iii) This is an immediate consequence of (ii). ■ §4. Transitivity of Rolling Without Slipping or Twisting in this section we will study the "composite" of two "rollings without slipping or twisting." Theorem 4.1. Let three submanifolds Mo,M1?M2 С ΈΙΝ be given which are tangent to each other at some point ρ £ M0 Π Μγ Π Μ2, and let a path σ0: (/,0) —► (M0,p) be given. Suppose that (a) M\ rolls on M0 along σ\: (1,0) —» (M\,p) without slipping or twisting via Q\: (1,0) —+ (Eucn(H.),I), with development σο: (1,0) —► (Mo,ρ), (b) M2 rolls on Mi along σ2 without slipping or twisting via g2. (1,0) —» (Eucn(R),I), with development σ^ (/,0) —» (Мьр). Then it follows that (c) M2 rolls on Mq along σ2 without slipping or twisting via #i#2: (/, 0) —» (Eucn(R),I), with development σο: (/,0) —» (Mo, ρ). Proof. It suffices to show that <7i#2 verifies properties (1), (2), (3), and (4) of Definition 1.1. (1) By (a), a0(t) = si(f)*i(0 and Τ^φ,^Μ,) = Τσο(ί)(Μ0). By (b), σι(ί) - 02(0*2(0 and Τσι(ί)(02(θΜ2) = Τσι{ί){Μχ). Thus, σ^0 = g\(t)g2(t)a3(t) and Та1(0ЫО£2(0^з) = 0ι(Ο*Τσ2(ί)(02(θΜ3) = ^(0*τσ2(ί)(Μ2) = Tai{t)(gi(t)M2) = Τσι{ί)(Μι).
§4. Transitivity of Rolling Without Slipping or Twisting 389 In the proof of the remaining parts, we shall need the (easily proved) formula (SiWfciOJ'iffiiOfcit))"1 =Λ(ί)5ι(ί)_1 +ft(<)fls(Oft(t)"1Si(<)"1· (2) {9i{t)92{t))-{9x{t)g2{t))-lax(t) = 9ι(ί)9ι(ί)-ισι(ή + 9i{t)92{t)92{t)-l9i{t)-^x{t) = 0 + 9l(t)g2(t)g2(t)-la2(t) = 0. (3) ((Λ(ί)52(ί))·(5ι(ί)№(ί))-1),Τσι(ί)(Μ1) = (^(«Ы*)"1).^)^) + ЫОЫгЫ*)-1^)-1)*^)^) С Τσι(ί)(Μ1)χ +9i(tU92(t)92(t)-l),9i(t):1T<Tl(t)(Ml) с т^м^ + Ы0*(Ы%2(*Г1)*т(Г2(4)(м2) с T<ri(t)(M1)-L + fll(i).Toa(t)(M2)-L сг^^смо-1-. (4) is similar to (3). Η Corollary 4.2. Suppose that two manifolds Мь M2 С RN are given which are tangent to each other at some point ρ £ Μχ Π M2, and let a path σ\: (/,0) —► (Μι,ρ) be given. Suppose that (a) M2 rolls on M\ along σ\ without slipping or twisting via g\\ (/,0) —> (Eucn(R),I), with development σ2: (/,0) —► (M2lp). Then it follows that (b) M\ rolls on M2 along σ2 without slipping or twisting via g2 = pj"1: (7,0) —» (Etxcn(R, 7), w/ii/i development σ\\ (7',0) —► (Mbp). Proof. Apply the theorem to the case Ms = M\ to see that M\ rolls on Mi along σι without slipping or twisting via g\g2. (7,0) —► (Eucn(R), 7), with development σ3: (7,0) —► (M2,p). But obviously pxp2 = -f and σ3 = σχ. Β Corollary 4.3. The "rolling data" consisting of a fixed manifold Μ and a curve σ on it depend only on the curve in space and the tangent spaces to Μ along the curve. Proof. Construct a manifold M\ out of σ and the tangent spaces along the curve by letting Vt be the subspace of Τσ^(Μ) orthogonal to af(t) and setting Μγ — UVJ. It is easy to see that M\ is a manifold in some neighborhood of σ which is a smooth curve on M\. Moreover, the rolling
390 Appendix B. Rolling Without Slipping or Twisting map of Μγ on Μ is the identity. Thus, by multiplication, if we wish to find the rolling map of Μ along (Μ,σ), it is enough to roll it along (Μχ,σ) or indeed any other manifold tangent to Μ along σ. ■ Exercise 4.4. Verify that the picture given at the beginning of this appendix is correct. That is, show that a sphere rolling on the central line of a helicoid rolls along a line of longitude. [Hint: use the fact that the central line of a helicoid is a geodesic] □
Appendix С Classification of One-Dimensional Effective Klein Pairs The classification of the one-dimensional effective Klein pairs was first given by Lie in the nineteenth century. In §1 we give a modern proof. In fact, the method we follow forms the basis for the "rough" classification of the primitive effective Klein pairs of any dimension over any field of characteristic zero (cf., e.g., [K. DePaepe, 1996]). More information on the classification of Klein pairs over the real and complex fields may be found in [K. Yam- aguchi, 1993]. §1. Classification of One-Dimensional Effective Klein Pairs Proposition 4.3.19. Let (g, fj) be an effective Klein pair with dim g/f) = 1. Then (g, fj) is isomorphic to one of the three Klein pairs described in Definition 4.3.18. Proof. Assume first that the adjoint action ad: f) —» End(g/f)) is injective. Since dim g/f) = 1, we get End (g/f)) = R. This means that either (i) t) = 0, in which case (g, f)) = (R,0), or (ii) f) = R and the adjoint map is an isomorphism. In this latter case we may choose a unique /iGf) such that ad(ft) = idfl/j>. Thus, if gi e g represents a basis element of g/f), then we have [h,g\] = Q\ +h\ for some h\ G i). But then the element g = g\+h\ also represents the
392 Appendix C. Classification of One-Dimensional Effective Klein Pairs same basis element of g/f), and moreover [h,g] = [h,gi H-fti] = [h,g\] -fO = gY -}- h\ = g. The map sending ι о J and ^(o ι establishes the isomorphism with the Klein pair of the affine line. This leaves us with the most complicated case where the adjoint action ad: f) —» End (g/f)) has a non zero kernel, which we assume for the rest of the proof. Define a sequence of subalgebras U-\ D Uq D U\ D Ό<ι D ... by setting U-i = g, U0 = f), and then, inductively, £/s+1 = {x e Us \ [g,x] с Us} for s = 0,1, In particular, we note that Ui = ker(ad: f> -» End(fl/f>) ^ 0. Sublemma 1.1. (i) [UuU^CUt+j. (ii) For some μ > 0, we have dim(£/s/£/s+1) = 1 for — 1 < 5 < μ and Us = 0 /ОГ 5 > /i. (iii) [£7-b£7e] ££/β/οτ·0<*<μ. Proof, (i) The result is obviously true if г < 0 and j < 0. Fix s > 1, and assume by induction that the result is true for —1 < г, —1 < j, and i+ j < s. Suppose г Η- j = s. Then ad(fl) [£/*,£/,·] С [adWbUj] + [C^ad^] С [i/t_i, ί/j] -l· [17», i/j_i] С t/t+j-i (by induction). Thus, [£/*,£/,] С £/*+,. (ii) Since dim g < oo and Uo D U\ D Ό<ι..., it follows that, for some minimal μ, ϋμ+\ = £/μ+2· But then [g, Ε/μ+ι] С Ε/μ+ι. Thus, ϋμ+ι is an ideal in g contained in I) and hence vanishes by the hypothesis of effectiveness. Since U\ φ 0, μ > 0. Next consider the maps ad: Us —► Hom(g, U3-\). These induce maps £7β/£7β+1-> Hom(fl, £/β-ι/^Λ which are injective because, by definition of the C/Ss, if χ G U3 and [x, g] С t/s, then χ G f/s+i. Since [t/s,f)] С U3, these maps induce injective maps £7β/£7β+1-> Hom(fl/f>, £7β_ι/£7β). Since dim g/f) = 1, it follows that dim([/s/17s+i) < dim(t/s_i/17s). But dim U-i/Uo = dim g/f) = 1. Assuming inductively that dim(U3-i/U3) = 1, it follows that dim(C/s/{7s+i) < 1. Either dim(Us/Us+i) = 1, in which case
§1. Classification of One-Dimensional Effective Klein Pairs 393 s < μ and the induction continues, or dim(U3/U3+\) = 0, in which case U3 = U3+i so that s = μ -f1. (iii) [U-i,U3] С U3 => Us С Us+i => Us = Us+i => s > μ. Μ Sublemma 1.2. The associated graded Lie algebra U-\/Uo®Uq/U\Q>· · -0 υμ/υμ+\ has a basis es G Us/Us+\,s = — 1,0...,μ, such that (i) [eo, e3] = —2se3 for s = —1,0,1,... and (ii) [e_bes+1] = e3 for s = 0,1,.... Proof. Part (ii) of the Sublemma 1.1 tells us that the associated graded Lie algebra U-\/Uq Θ Uq/U\ Θ · · · Θ υμ/υμ+\ has every grade of dimension one, and part (iii) says that [C/_i/C/o, U3/U3+\] Φ 0 for 0 < s < μ. Fix basis elements e_i G U-\/Uo and (temporarily) eM G £/μ/£/μ+ι, ancj define es = [e_i, es+i] for s = μ — 1,..., 0 inductively. By Sublemma 1.1 (iii) none of the e3s vanish. Now define Xj by [eo, e3] = X3e3 for s = — 1,0,1,..., μ, and note that Xses = [e0,e3] = [e0, [e_bee+i]] = [[eo»e_i],ee+1] -h [e_b [e0, es+1]] = [A-ie_i,ee+i] -h [e_b As+1es+1] — ^_ies -f- \3+\е3. This yields As+i = As —λ_ι. Thus, As = —sA_i. Since [C/_i/{7o, Uq/U\] has —A_ie_i = [e_i,eo] as a basis, it follows that λ_ι φ 0. If we replace our choice of eM by 2eM/A_i, this replaces es by 2es/A_i for s = μ — 1,..., 0 and also replaces A_i by 2 so that A_s = —2^. However, the equations e3 — [e_i, es+i] for s = μ — 1,..., 0 continue to hold. Ш Sublemma 1.3. The Lie algebra g has a basis x3, — 1 < s < μ, with x3 G U3 a lift of e3 G U3/U3+\, such that (i) [xq,x3] = —2sxz for s = —1,0,1,... ,μ and (ii) [χ_ι,χ3+ι] = xz for s = 0,1,... ,μ - 1 and (iii) [xt,x3] G {xt+3) {the span of xt+3) Proof, (i) Let Xq G f) be any lift of eo- We know from linear algebra that the characteristic polynomials of ad(xo) and ad(eo) are the same. Since, by Sublemma 1.2, ad(eo) is diagonalizable, with all eigenspaces of dimension one and eigenvalues 2,0, —2,..., — 2μ, the equality of the characteristic polynomials shows the same is true of ad(xo)· For s φ 0, let x3 denote an eigenvector of ad(xo) with eigenvalue — 2s. Replacing each xs, — 1 <s</i,5^0, by some multiple, we may assume that x3 G U3 is a
394 Appendix C. Classification of One-Dimensional Effective Klein Pairs lift of es G Us/Us+i for — 1 < s < μ. It is clear that xs — 1 < s < μ forms a basis for g. This proves (i). Now note that 8Ld(x0)[xt,Xs] = [did(xo)xt,Xs] + [xt,&d(x0)xs] = -2t[xt,xs] - 2s[xuxs] = -2(t + s)[xt,xs]. (ii) Now [x_bxs+1] G Us covers [e_i,es+i] = es G Us/Us+\ for all s. Since, by the preceding formula, 8Ld(x0)[x-i,xs+i] = -2s[x-i,xs+i], we see that [x_bxs+1] is the unique eigenvector of ad(xo) lifting es. Thus [ж_ьж5+1] = xs for s = 0,1,...,μ - 1. (iii) Since ad(xo)[#i,#s] = —2(t + s)[xt,x3] we see that [жг,ж5] lies in the —2(t -f s)-eigenspace of ad(xo) which, by (i), is (xt+3). Ш Next we use the Killing form В on g defined by B(x,y) = Trace(ad(x)ad(?/)). Note that В(ж0, ж0) = Trace(ad(x0)2) = (~2)2 -f 02 + 22 + 42 + · · · + 4/i2 > 0. Sublemma 1.4. B([x, y], z) = B(x, [τ/, ζ]) for all x,y,z G g. Moreover, the radical of В is an ideal of g. Proof. We have аа([ж,г/]) = [ad(x),ad(?/)] G End(g), so, setting X = ad(x), Υ = ad(?/), Ζ = ad(z), we have B([x, y], z) — B(x, [г/, ζ]) — Trace(ad[x, y\&d{z)) — Trace (ad (ж) ad [г/, ζ]) = Trace([X, Y]Z) - Тгасе(Х[Г, Ζ]) = Trace(XrZ - YXZ - (XYZ - XZY)) = Trdice(-YXZ + XZY) = Trace([XZ,r]) = 0. By definition Rad(£?) = {x G g \ B(x,g) = 0} is the radical of B. If χ G Rad(£?) and у G g, then В([ж,г/],д) = £?(ж, [г/, д]) = 0 by the first part of the lemma. Thus, Rad(£?) is an ideal. Ш Now we are in a position to complete the proof of the proposition using the basis ж_ь Жо,... for g constructed in Sublemma 1.3. Sublemma 1.3(iii) shows us that ad(xs)ad(xi)xw G (xs+t+u)- It follows that for s -f t Φ 0,
§1. Classification of One-Dimensional Effective Klein Pairs 395 &ά(χ3)&ά(χι) is nilpotent. In particular, B(x3,xt) = 0 unless s -f t = 0. Thus, Rad В D U2. Suppose U2 φ 0. We derive a contradiction as follows. Since 0 φ U2 С Rad(£?) and Rad(£?) <£_ f) (since (g, fj) is effective) it follows that there is an element of the form αχ-γ + βχ0 + 7#ι G Rad(£?) with α φ 0. Since i?(:ro> #o) 7^ 0? we have 0 = B(ax-i + βχ0 +7Χι, x0) = βΒ(χο,χο)=ϊβ = 0. Since Rad(£?) is an ideal we also have [χ-ι,αχ-ι +7X1] = 7x0 £ Rad(£?) so 7 = 0. Finally we have [χι,αχ-i] = axo G Rad(£?) so a = 0, which is a contradiction. Thus, U2 = 0 and so μ < 1. But by Sublemma l.l(ii), μ > 1. Thus, /i = 1 and the map sending χ-^(! ο) χο -(ν ;) Xl "(2 J) establishes an isomorphism between (g, fj) and the Klein pair of the projective line. ■
Appendix D Differential Operators Obtained from Symmetry In this appendix we apply the ideas about geometric differential operators described toward the end of §5 of Chapter 3 to show how curl, divergence, and the Cauchy-Riemann operator arise from symmetry considerations on a two-dimensional Riemannian manifold. This appendix may be regarded as an introduction to [E. Cartan, 1966]. The main aim of that book is to obtain the Dirac operator on a four-dimensional Lorentzian manifold. We present the simplest case of Cartan's method, with the aim of clarifying the principles involved in obtaining geometric differential operators from symmetry considerations in a Cartan geometry.1 Throughout this appendix we are considering an oriented Riemannian surface with model (β, f)) = (euc2(R), so2(R)) and group Η = S02(R). In §1 we give a very pedestrian introduction to the "plethysm" of representations of the circle group. In §2 we apply this to decompose the covariant derivative on Riemannian surfaces. §1. Real Representations of 50г(К) Let Vk = (R2,pfc), к G Z, denote the real two-dimensional representation of S02(R) given by 1On p. 55 of [E. Cartan, 1966] one can find a very brief description of the analogous material in the case of Euclidean 3-space.
398 Appendix D. Differential Operators Obtained from Symmetry (cos# — sin#\ (cosкв — sinA;#\ сгл ,— . sinfl cosi? ) = [sinke созкв J e 5°2(R)- The isomorphism SOy(K) - Uy(C) (cos θ — sin#\ .a ) k+ егв sin θ cos θ J allows us to make the reinterpretation Vk = (С,р^), with рк(егв) = егкв. Let ек, Д denote the standard real basis for Vk and let e£, /£ G V£ be the dual basis.2 Under the identification R2 и С we have the correspondence Хквк + 2/fc/fc ^ ^fc = %k + ij/fc- Exercise 1.1. Show that the representations V±k are isomorphic (over R). Note that Vo = U Θ Ε/, where С/ is the trivial one-dimensional representation. In fact, the representations Vk, к > 0, together with [/, constitute a complete set of isomorphism classes of all the irreducible representations. Every finite-dimensional representation is isomorphic to a direct sum of copies of these representations. However, we don't really need these facts. What is important for our study is the Clebsch-Gordan (or plethysm) problem of describing the decomposition of the tensor products of the representations Vfc. This is given in Lemma 1.2. Further information on plethysm may be found in [W. Fulton and J. Harris, 1991]. Lemma 1.2. (i) For k,l ^ 0, there is an isomorphism of representations φ: Vk 0 Vi = Vk-ι Θ Vib+j defined by ( ф(ек 0 ei) = \ek-i + \ek+u I Ф(1к 0 fl) = \zk-l - \tk+U I Ф(ек 0 fi) = -\fk-i + \fk+i, { Φ{ίк ® et) = \fk-i + \fk+i or, equivalently, ( <jrl{ek+i) = ek 0 ei - fk 0 Д I </>-l(ek-i) = ek 0 ei + fk 0 //, | 0_1(/fc+z) = ek 0 fi + Д 0 ez, (ii) T/iere is an isomorphism of representations r: V£ —» Vfc defined by 2Note that the left 50n(R) module structure on Homj^(Rn,R) is always taken to be (h · y?)(v) = ip(h~1v). While it would be possible to drop the inverse in the present abelian (n = 2) case, there would eventually be conflict with the other cases.
§1. Real Representations of 50г(К) 399 jr(el) = ek, \r(/fe*) = /fc. Proof, (i) It is clear that φ is a linear isomorphism. Let Zk £ T4, ^ G Vj, regarding these as complex numbers. Then the formulas above imply that ф(гк <g) ζι) = {\zkZi, \zkz{) e Vk-i Θ T4+j, which makes obvious the fact that φ is an 502(R) module isomorphism. (ii) Clearly, r is a linear isomorphism. We must see that it satisfies r(eie · φ) = eie · τ(φ) for all φ e V£ = HomR(Ffc,R). We may write ^ = ^(efc)4+^(/fc)/fc so that we have eie-V = (eie-V)(ek)el + (eie^)(fk)rk Thus, r{eie ■ ψ) = тЫе-мек)е*к + V{e~ike' Д)Я) = ip(e-ikt)ek)ek + ψ{ε~ίΜ fk)fk = v?(cos(fc#)efc - sm(k9)fk)ek + ip(sm(k6)ek + cos(k6)fk)fk = (p(ek)(cos(ke)ek + sin(ke)fk) + <р(Д)(- sin(fc<9)efe + cos(fc<9)/fc) = <p(ek)(eu,-ek)+<p(fk)(ei°-fk) = eie ■ Ыек)ек + р(А)Л) = ei0 ■ τ(φ). Ш Let /о о о \ /о о о\ /о о о\ So = 0 0 -1,^=1 0 0 , Е2 = 0 0 0 е в, \0 1 О / \0 О О/ \1 О О/ and let Eg*, £?, Щ G β* be the dual basis. Corollary 1.3. Vk <8>β* и 14-ι θ^θ Vfe+1, where the isomorphism sends ek®E£^ek, fk®E^^fk, ek®E\ y-> iefc_! +^efe+i, /fc®£i ·->· i/fc_! + ±/fc+1, ek ® E^ ι-»· -^/fc-i + l/fc+i, fk®E^^ \ek-i - \ek+l. Proof. Since f) ~ U and ρ и V^, it follows from Lemma 1.2(ii) that
400 Appendix D. Differential Operators Obtained from Symmetry g* = i)* ep* «t/e Vi. e* k+ и (a generator of U) El ι—► ei The rest is a direct application of Lemma 1.2(i). ■ Exercise 1.4. Find analogs of Exercise 1.1, Lemma 1.2, and Corollary 1.3 for the group 02(R). □ §2. Operators on Riemannian Surfaces Let Μ be an oriented Riemannian surface and let Ρ be the oriented or- thonormal frame bundle over Μ with canonical projection π: Ρ —» Μ. Namely, Ρ = {(х,еье2) I x e M,ebe2 G TX(M)} where the pairs (ei,e2) are oriented orthonormal bases of TX(M) and 7г(ж, ei, e2) = x. The right action of S02(R) on Ρ is given by (x,ei,e2) · #(s) = (x,ei cos 5 -fsinse2, —ei sins -f cosse2), where „, ν /cos5 —sins Я(5) = v ' \sms cos 5 We are going to link up the representation theory of §1 with the derivative on P. We begin by expressing the Cartan connection form on Ρ as /0 0 0 \ ω= \θι 0 -a) eAl(P,zuc2(R)). \θ2 α 0 / The forms α, θ\, θ2 are defined (cf. Chapter 6) by the equations π*Μ = 0i(y)el + 02(v)e2 for all ν е Т(ж>М1>М2)(Р), άθι-αΑθ2 = Ο, d6>2 + α Λ 6>! = 0. Let Xq, Χχ, X2 denote the vector fields on Ρ dual to α, θ\, θ2. Lemma 2.1. (i) 7Γ*(-ΧΊ) = еь π*(Χ2) = e2. (ii) If X is an arbitrary vector field on M, its horizontal lift {i.e., the unique lifl on which a vanishes) at (x,ei,e2) G Ρ is given by X — (X ■ e1)X1 + (X ■ e2)X2.
§2. Operators on Riemannian Surfaces 401 Proof, (i) This follows from the equation π*{ν) = #i(^)ei + #2(^)e2 and the fact that Χχ,Χ2 are dual to #i,#2. (ii) Since α(Χγ) = a(X2) = 0, it follows that (X · ex)Xi + (X · e2)X2 is horizontal, and by (i) it is a lift of X. ■ Lemma 2.2. (i) Fix ρ G P, and define c:I —► Ρ by c(s) = ρ · R(s). Then c'(0) = Xqp. (ii) Lei F: Ρ —* Vk transform according to F(ph) = Pk(h~l)F(p) for h e SOa(R). Then (X0F)(p) = * ( Д j) F(p). (iii) Lei X be a vector field on Μ', and consider the functions X · ef. Ρ —» R, г = 1,2. T/ien Jf0<-y,e1> = (Jf,e2>, Xo<-y,e2> = -<Jf,ei>. Proof, (i) Factor the curve с as the composite. Then c*cj = (μ ο (id x R) ο υ2)*ω = 6* о (id* хй>^ = б* о (id* x Д*)^^)-1)^ + тг*^я) = Ad(R(s)~l)(np ο (id x R) ο *,2)*u; + (тгя о (id x R) ο υ2ΥωΗ = Ad(R(s)-l)(np о 62)*cj + #*и;я = Д*ЗД, since 7Гр о б2 is constant. Thus, "Mo))=4-(s)LH4s = Д*и;я s=0 б?5 ■■(JH\R* ' ds s=0 -л'г s=0 1 -1 1 о s=0 Thus, c'(0) = Jf0. (ii) Fix ρ G P, and take с as in (i). Then by (i) we have (XqF)(p) = F(c(5))'|s=0. Now F(c(s)) = F(p ■ R(s)) = pk(R{a))-lF(p) = R(-ks)F(p), so (X0F)(p) = R(-ks)'\s=0F(p). But *(-M'L=o cos ks sin ks — sin ks cos A;s 0 к -к 0 s=0 -fcsinfcs к cos ks -k cos ks —к sinks s=0
402 Appendix D. Differential Operators Obtained from Symmetry (iii) Define F:P -* R2 by F(p) = (j*'^j J, where ρ = (x,eue2). Then F(p.R(s))=(/^C0SSe'+Sinse2\) v v " \\X, -smsei +cosse2)) = fC0SS SinS)({/Xyei\)=R(s)F(P). \-sms cossj \(X,e2) у Thus, F transforms as a function F: Ρ —> V\. Hence, (X0F)(p)=^1 l^)F(p), or X0(X,el)\_f (X,e2) ^Xo(X,e2)J \-{X,ei) Proposition 2.3. LetT(k) denote the space of smooth functions F: Ρ —> Vk transforming according to F(ph) = pk(h~l)F(p) for h e Η = S02(R).3 Write F e T(k) as F=(^)=Fiek + F2fk. (i) The universal covariant derivative DF: Ρ —» Vk 0 £J* is given by OF = XY[F) 0 E\ + X2(F) 0 £* + Х0(Л 0 £0*· (ii) Decomposing DF in terms of the components of Vk 0 g* = 14_ι Θ Vfc Θ Vfc+i г/ге/ds (when к φ 0) ί/ie ί/iree operators D: г(к) {д^д) Г(Л -1)0 Γ(Λ) Θ Γ(Λ + 1), where I f Χι Χ2\ Ύ , ( 0 l\ δ Ι (Χλ -Χ2 9=2\-Χ2 Χι)' Jk = k{-1 Ol· 9=2V^2 * Proof. For v e g, we have (DF)v = ω l(v)F:P —► Vfc. In particular, (DF)Ei = Xi(F), 0 < г < 2, which verifies (i). Now calculate 3Of course, Г(&) may also be interpreted as the space of smooth sections of the associated 2-plane bundle Ρ x# Vk-
§2. Operators on Riemannian Surfaces 403 DF = X0{FYek + F2fk) ® К + Xi(Fiek + F2fk) ® F* + X2(Fyek + F2fk) ® EZ = XoiF,)^ ® F0* + X0(F2)fk ® F0* + X^e* ® F* + Xy(F2)fk <g> F* + X2{Fx)ek ® F* + x2(F2)/fc®F2*. Breaking up this expression according to the decomposition of Vk ® β* indicated in Corollary 1.3 yields £г = Хо(^1)ек + .ВДг)/к + X1(F1) Qefc_! + lefc+i) + лда) Q/fc_! + i/fc+1 + X2(F0 (-^/fc-i + |/fc+i) + *2(F2) Qefe_! - ^efc+1 + ^(№(F0 + X2(F2))efc_1 + (A-!(F2) - Χ2(^))Λ-ι) + i((X1(F1)-X2(F2))efc+1 + (X1(F2) + X2(F1))/fc+1). This, together with Lemma 2.2, yields the result. Η Corollary 2.4. T/ie covariant derivative DF: Ρ —> Vk <8> p* is given 6?/ DF = Xj (F) <g> F* + X2(F) ® F2* and decomposes as T(k){^T(k-l)®T{k+l). Proof. This follows directly from the proposition. Η Definition 2.5. The operator В is called the Cauchy-Riemann operator. σο The following lemma prepares the way for the calculation of д and д in a gauge. Lemma 2.6. Fix a local section σ(χ) = (x,ei,e2) of Ρ (so e\,e<i is an orthonormal frame on M). (i) At points of Ρ in the image ofa, we have Xi = a*(ei) — a(a*(ei))Xo, г = 1,2. Now let 0 0 0 θχ 0 -a <92 a 0 denote the infinitesimal gauge corresponding to σ.
404 Appendix D. Differential Operators Obtained from Symmetry (й) а{Х · ei) - a(ei)X · e2 if j = 1, -№ * ei) - Ί e.(X . e2) _ a(e.)X . ei if j = 2. Proof, (i) By Lemma 2.1(i), π*(Χ.) = e», i = 1, 2, so 7Γ*(Χ; - а*(е»)) = 0. Thus, Xi = a*(ei) + λΧο for some function λ. Applying a to this equation yields 0 = а(а*(е.)) + λ. (ϋ) X,(X · e,·) = (a - а(а)Х0)(Х · ea) by (i) = et(X · ej) - a(ei)X0(X · e,·) = f е»рГ · ei) - a(ei)X · e2 if j = 1, \ ei(X · e2) - a(et)X · ei if j = 2, where the final equalities come from Corollary 2.2(Hi). ■ Theorem 2.7. (i) If we interpret the operator <Э:Г(1) —» Г(0) as a map from vector fields on Μ to pairs of functions on M, it assumes the form, with respect to the gauge above, dF = 1 f ex(X - ex) + e2(X · e2) - a(e2)X - ex - а(ех)Х · e2\ 2 Vei(^ * Ы - e2(X · eY) + a(e2)X · e2 - a(ei)X · eY )' In particular, in the case when Μ is the Euclidean plane with the standard coordinates X\, x2, we may take e\ = д/дх\, e2 — д/дх2 (in which case, a = 0), and the last expression reduces to ι /EL· + UL· f)F — _ ( dxi dx2 x дх\ дх2 whose components are half of the usual expressions for divergence and curl (ii) If we reinterpret д: Г(0) —» Г(1) as a map from R2-valued functions on Μ to vector fields on M, this operator takes the form df = |(M/i) - e2(/2))ei + (e2(/x) +ei(/2))e2). The right side is the same no matter what orthonormal basis (ei, e2) is used. In particular, when Μ is the Euclidean (x\,x2) plane and e\ = д/дх\, e2 = д/дх2, the equation df = 0 becomes the Cauchy-Riemann equations dj\=dh df1 = _df1 дх\ дх2 дх2 дх\' Proof, (i) The vector field X on Μ corresponds (cf. Corollary 5.3.16) to the function F:P —► V\ given by F(x,ebe2) = (X · eb X · e2). Thus,
§2. Operators on Riemannian Surfaces 405 1 ( Χι X2\ {FA _ 1 / Xi(Fl)+X2(F2) dF 2\-X2 Xi)\F2) 2\-X2(Fl)+Xl(F2) 1{ Χλ(Χ - ei) + X2(X - e2) 2 \-X2(X-el)+Xl(X.e2) By Lemma 2.6 we get aF = 1 { ei(X · ei) - a(ex)X · e2 + e2(X · e2) - a(e2)X · б! 2 V ~e2(^ · ex) + a(e2)X · e2 + eY(X · e2) - a{eY)X · б! which yields the formula. Thus the final remark is clear. (ii) Reinterpreting / = ( у j : Μ -» R2 as the element F = π* / = ( ^ J) ) Ξ Γ(0), we calculate α 2 V ^2 JfJ Wb) 2 ν^2(π*/ι)+^ι(π*/2)7 ~2 \(**(X2))fi+(**(Xi))f2j 2 W/O+^W = 2^,e1>e2)((ei(/i) - e2(/2))ei + (e2(/i) + ex(f2))e2). Interpreting this as a vector field on Μ gives the result. Η We remark that the Cauchy-Riemann operator on a Riemannian surface is a special case of the Dirac operator on an even-dimensional manifold.
Appendix E Characterization of Principal Bundles We give a proof of the following result stated in Chapter 4. Theorem 4.2.4. Let Ρ be a smooth manifold, Η a Lie group, and μ: Ρ χ Η —* Ρ a smooth, free, proper right action. Then (i) P/H with the quotient topology is a topological manifold (dim P/H — dim Ρ - dim #), (ii) P/H has a unique smooth structure for which the canonical projection Ρ —» P/H is a submersion, (iii) ξ — (Ρ, π, Μ, Η) is a smooth principal right Η bundle. Proof, (i) Step 1. The orbits of Η yield a foliation on P. Since the composite μο \h:P —* Ρ χ Η —+ Ρ sending ρ ι—► (ρ, Κ) ι-* ph is a diffeomorphism for all /i £ H, it follows that the action μ: Ρ χ Η —► Ρ is a submersion and hence induces a foliation on Ρ χ Η with leaves of the form Lp = {(q,h)ePxH\qh = p}. The leaves of this foliation are permuted under right action of Η on the second factor of Ρ χ H. This foliation projects, via the first factor projection Ρ χ Η —» Ρ, to a foliation on Ρ whose leaves are the orbits of H. Step 2. The leaves of the foliation on Ρ are properly embedded subman- ifolds.
408 Appendix E. Characterization of Principal Bundles Define φρ: Η —» Ρ by ft —» ph. Every leaf is the image of one of these maps. Since the map φρ factors as φρ = μορρ, where pp:H —> PxH sends h —» (p, ft), it is smooth; since the action is free, φν is injective. If К С Ρ is a compact set, then φ~ι{Κ) = {ft G Я | {p}ft ПК/Й}, which is compact since {p} and if are compact and the action is proper. Finally, we show that φρ: Η —» Ρ is an immersion. If it is not an immersion, then for some ft G Я the map (pp*h:Th(H) —► Tph{P) is not injective. But then the commutativity of the diagrams (where ra\P-+P sends g ι-* да for α G Я) гл(Я) —> ^(P) shows that (pp*e:Te(H) —» Tp(P) is not injective (take а = ft-1) and that if ^ G ker (^p*e, then Pa*^ G Ker φρ*α for all a e H. Choose such a ^ 7^ 0, and let ^:R -> Я be the one-parameter subgroup satisfying ^*e(l) = v. Let 5 G R, and set ft = ^(5). Now the derivative of the composite φρ ο ψ at 5 G R is ipp*ip*h-Ts(R) —► Th(H) —» Tph(P), which vanishes. Thus the composite ^p ο ^ is constant, and in particular <pp is not injective, which is a contradiction. 5iep 5. For each point ρ Ε Ρ there is a foliated chart (ΙΙ,φ), with <?: Wp) ~» (Rn x Rm_n,0) and a set Τ = φ~ι(0 χ P™"n), with ε > 0, such that the map a:T χ Η —> Ρ sending (£,ft) ь-» ift is a diffeomorphism onto its image. Fix ρ Ε Ρ, and let X be the leaf through p. Since X is proper, we may choose a foliated chart ([/, <p), with φ: (Ε/, Е/П£,р) -» (Rn χ Rm_n, Rn x 0, 0 χ 0). Now we set Τ = Τε = φ~ι(0 χ Ρ™"71), where B™~n is a ball of radius ε about the origin in Rm_n, so that Τ has compact closure. We claim that there is a neighborhood of the identity V С Η such that TV С U. For if this fails, we can find a sequence of neighborhoods of the identity Vn С Я, п = 1,2,..., which all have compact closures and satisfy Я - я - % % -> Ρ
Appendix E. Characterization of Principal Bundles 409 oo Vi D V2 D V2 D V3 D · · · D f] Vn = e n=l such that each set TVn meets the complement of U. Now choose sequences tn G Τ and vn e Vn with int;n in the complement of U. Since Τ and V^ have compact closures, we may, by passing to subsequences, assume that tn and vn converge to t G Τ and e G H, respectively. Thus, tnvn —* t, which again lies in the complement of U. But this is impossible since t G Τ С U. This proves the existence of V С Η such that TV С U. Replacing 1^сЯЬу its identity component, we may assume that it is an open path-connected neighborhood of the identity satisfying TV С U. Set Κε = К = {h e Η \fhnf φ 0}. Now since f is compact and the action is proper, it follows that К is compact. We claim that e G К is an isolated point of К so that К — {e} is again compact. It suffices to show that Vf\K = {e}. Let h e VC\K, and choose a path σ: (7,0,1) -> (V,e,ft). Since the path ρσ:(/,0,1)->(Ε/,ρ,ρΛ) lies in a single plaque and both p,ph G f, it follows that ph = ρ so that /i = e. Next we claim that for some ε > 0, the map α = αε:ΤεχΗ —► Ρ sending (ί, /ι) —* ί/ι is injective. Suppose not. Then we may choose, for each n, two unequal pairs of points (in, /in), (un, kn) €Τι/η χ Η with tnhn — unkn for all n. Thus, tn = unXn (where λη = knh~l) and hence λη G Κχ/η С ΚΊ· Now if we were to have tn — un, it would follow that λη = e, since the action is free, and then the pairs {tn,hn), (un,hn) would be equal. Since we assume this fails, it follows that we must have λη G K\ — {e} for all n. Since Κ ι is compact, a subsequence of the Ans converges to some λοο φ е. But tn = un\n and un both converge to p; therefore, taking the limit, we get ρ = ρλοο, which implies that λ^ — e, since the action is free. This is a contradiction. Since for the choice of ε in the last paragraph the map a is a smooth injection, and the dimensions of its domain and range are the same, we need only show that it is an immersion to finish step 3. Because of the commutative diagram TxH a ) Ρ idx/Ц |g TxH a ) Ρ it suffices to show that a*: T(i>e)(Tf χ Η) —► Tt(P) is injective for all t eT. But Γ(ί|β)(Τ χ Η) = Tt(T) ®Te(H), and the restriction of a* to the first factor is the inclusion-induced map Tt(T) —» Tt(P), while the restriction to the second factor is a map (pt:Te(H) —» Tt(P). These maps are both
410 Appendix E. Characterization of Principal Bundles inclusions, and the images lie in subspaces having only 0 in common, so a* is injective. Step 4- The quotient space P/H is Hausdorff. Let p, q G Ρ lie in distinct Я orbits. Choose a transversal Τ containing ρ as in step 3. We may assume that the oribit of q does not meet the closure of T, for it meets the coordinate system about ρ in at most one plaque, which we may avoid by shrinking the transversal T. Similarly, let U be a transversal containing q whose closure does not meet the orbit of p. Now the orbits UH form a closed set that meets Τ in a closed set T0. Since ρ e Τ — T0 and the latter is open in T, we may further shrink Τ so that Τ Η Π UH = 0. In particular, Τ Η and UH are disjoint, so their images separate the images of ρ and q in P/H. Step 5. The quotient space P/H is locally Euclidean of dimension equal to dim P — dim H. If Τ is one of the transversals through ρ guaranteed by step 3 and π: Ρ —» P/H is the canonical projection map, then (π(Τ), (π | T)~l) is a coordinate chart for P/H. It suffices to show that the induced map π | Τ: Τ —» π(Τ) is a homeo- morphism. It is certainly continuous, and by step 3 it is bijective. But it is also open, for if U С Τ is open, then UH is open and hence n(U) is open in тг(Т). Finally, dim P/H = dim Τ = dim Ρ - dim Я. (ii) The charts (π(Τ), (π | Τ)-1) arising from the transversals of step 3 yield a smooth structure of G/H. It suffices to show that two such charts, (π(Τ), (π | T)_1) and (π(17), (π Ι [Ζ)"1) sav> are smoothly compatible. But (π | Τ)-χ(π(Τ) Π тг(Т)) = Τ HUH and (π | U)~l(n(T) Π тг(Т)) =THnU. Moreover, the change of coordinate mapping is just the "sliding" map Τ Π UH —► ТЯ П С/, which is known to be smooth from Chapter 2. (iii) ξ = (Ρ, π, Μ, Я) is a smooth principal right Η bundle. Let Τ be a transversal as in step 3. Then π(Τ) is an open set in P/H. Define ф: тг(Т) х Я -> π-χ(π(Τ)) by V>M0> Λ) = ^· This was shown to be a diffeomorphism in step 3. Given two transversals T\ and T2, we get two maps φι and V>2 and the associated transition function φ- Vi: π(Γι) Π тг(Т2) х Я -> π(Τχ) Π π(Τ2) χ Я. Define /:Τι -> Я by setting ?/>2"Vi(7r(0>e) = (*"(*),/(*))· Then / is smooth and ^νι(π(ί),Λ) = (π(ί),/(ίΓ1Λ). Thus, the maps φ constitute an Η atlas for the bundle Ρ —► P/H. Ш
Bibliography Μ.A. Akivis and V.V. Goldberg, Conformal Differential Geometry and Its Generalizations, Wiley Interscience, 1996. M.A. Akivis and V.V. Goldberg, Projective Differential Geometry of Sub- manifolds, North-Holland, Amsterdam, 1993. J. Beem and P. Ehrlich, Global Lorentzian Geometry, Dekker, New York, 1991. M. Berger, Geometry II, Springer-Verlag, New York, 1987. W. Boothby, An Introduction to Differentiable Manifolds and Riemannian Geometry, 2nd ed., Academic Press, San Diego, 1986. R.L. Bryant and L. Hsu, Rigidity of integral curves of rank 2 distributions, Invent. Math. 114, 435-461, 1993. R. Bryant, S.-S. Chern, R. Gardner, H. Goldschmidt, and PA. Griffiths, Exterior Differential Systems, Springer-Verlag, New York, 1991. W.K. Buhler, Gauss, A Biographical Study, Springer-Verlag, New York, 1981. G. Cairns and R. Sharpe, The inversive differential geometry of plane curves, VEnseignement Math. 36, 175-196, 1990. G. Cairns, R. Sharpe, and L. Webb, Conformal invariants for curves and surfaces in three dimensional space, Rocky Mountain Journal of Mathematics, 24, no. 3, 933-959, 1994. E. Cartan, Les systemes de Pfaff a cinq variables et les equations aux derivees partielles du second ordre, Ann. Ec. Norm. 27, 109-192, 1910. E. Cartan, Les espaces a connexion conforme, Ann. Soc. Pol. Math. 2, 171-221, 1923.
412 Bibliography Ε. Cartan, Sur les varietes a connexion projective, Bull. Soc. Math. France, 52, 205-241. See also the collected works, vol. Ilia, 825-861, 1924. E. Cartan, Le calcul tensoriel en geometrie projective, C.R. Acad. Sci. 198, 2033-2037, 1934. E. Cartan, La Methode du Repere Mobile, la Theorie des Groupes Continus et les Espaces Generalises, Exposes de Geometrie, V, Paris, Hermann, 1935. E. Cartan, Le calcul tensoriel projectif, Recueil Soc. Math. Moscow, 42, 131-148, 1935. E. Cartan, L'extension du calcul tensoriel aux geometries non affines, Ann. Math. 38, 1-13, 1937. E. Cartan, Legons sur la Theorie des Espaces a Connexion Projective, Paris, Gautier, 1937. E. Cartan, Families de surfaces isoparametriques dans les espaces a cour- bure constant, Annali di Mat. 17, 177-191, 1938. Also in the collected works. E. Cartan, Lecons sur la Theorie des Spineurs I, II, Exposes de Geometrie, XI, Hermann, Paris, 1938. E. Cartan, Selecta Jubile Scientifique, Paris, Gauthier, 1939. E. Cartan, La notion dOrientation dans les differentes geometries, Bull. Soc. Math. France, 69, 47-70, 1941. E. Cartan, The Theory of Spinors, Hermann, Paris, 1966, translation of Lecons sur la Theorie des Spineurs I, II, Exposes de Geometrie, XI, Hermann, Paris, 1938. E. Cartan and A. Einstein, Letters on Absolute Parallelism 1929-1932, Princeton University Press, Princeton, 1979. T. Cecil, Lie Sphere Geometry, Springer-Verlag, New York, 1992. T. Cecil and S.-S. Chern, Tautness and Lie sphere geometry, Math. Ann. 278, 381-399, 1987. S.-S. Chern, Sur la possibilite de plonger un espace a connexion projective donne dans un espace projectif, Bull. Sci. Mat. 61, 234-243, 1937. S.-S. Chern, The geometry of the differential equation y'" = F(x, y, yf', y"), Science Reports Tsing Hua Univ. 4, 97-111, 1940. S.-S. Chern, A generalization of the projective geometry of linear spaces, Proc. Nat. Acad, of Sci. USA, 29, 33-43, 1943. S.-S. Chern, The geometry of G-structures, Bull. Amer. Math. Soc. 72, 167-219, 1966. S.-S. Chern, On the projective structure of real hypersurfaces in Cn+1, Math. Scandinavia, 36, 74-82, 1975. S.-S. Chern, Circle bundles, in Geometry and Topology, III Latin Amer. School of Math., Lecture Notes in Math., Springer-Verlag, 597, 114— 131, 1977. S.-S. Chern, Selected Papers, Springer-Verlag, New York, 1978. S.-S. Chern, From triangles to manifolds, Amer. Math. Monthly, 86, 339- 349, 1979.
Bibliography 413 S.-S. Chern, Finsler geometry is just Riemannian geometry without the quadratic condition, Notices of the AMS, 43, no. 9, 959-963, 1996. S.-S. Chern and C. Chevalley, Elie Cartan and his mathematical work, Bull. AMS, 58, 217-250, 1952. S.-S. Chern and P.A. Griffiths, An inequality for the rank of a web and webs of maximal rank, Annali Sc. Norm. Sup.-Pisa, Serie IV, 5, 539- 557, 1978. S.-S. Chern and J. Wolfson, A simple proof of Frobenius theorem, Manifolds and Lie Groups, Papers in Honor of Y. Matsushima, Birkhauser, 347- 349, 1981. K. DePaepe, Primitive Effective Pairs of Lie Algebras over a Field of Characteristic Zero, Ph.D. Thesis, University of Toronto, 1996. J. Dieudonne, Treatise on Analysis, 4, translated by LG. Macdonald, Pure and Applied Mathematics, Vol. 10-IV, Academic Press, New York- London, 1974. C. Dodson and T. Poston, Tensor Geometry, 2nd ed., Springer-Verlag, New York, 1991. S. Donaldson and P.B. Kronheimer, The Geometry of Four-Manifolds, Oxford Science Publications, 1990. J. Dugundji, Topology, Allyn & Bacon Inc., Boston, 1966. C. Ehresmann, Sur les espaces localement homogenes, L'Enseignement Math. 35, 317-333, 1936. C. Ehresmann, Les connexions infinitesimales dans un espace fibre differ- entiable, Collogue de Topologie, Bruxelles, 29-55, Thone, Liege, 1951. С Ehresmann, Structures feuilletees, Proc. Vth Can. Congress, Montreal, 109-172, 1961. A. Einstein, Ether and Relativity, in Sidelights on Relativity, translated by G.B. Jeffrey and W. Perrett, Methuin, 1922. L.P. Eisenhart, Non-Riemannian Geometry, AMS Colloquium Publications VIII, fifth printing, 1964. D. Ferus, H. Karcher, and H. Munzner, Cliffordalgebren und neue isopara- metrische Hyperflaschen, Math. Z. 177, 479-502, 1981. A. Fialkow, Conformal differential geometry of a subspace, T.A.M.S. 56, 309-433, 1944. B.L. Foster, Would Liebniz Lie to You? Math. Intelligencer, 8, no. 3, 34-40, 1986. D. Freed and K. Uhlenbeck, Instantons and Four-Manifolds, Springer- Verlag, Berlin, New York, 1984. W. Fulton and J. Harris, Representation Theory, A First Course, Springer- Verlag Graduate Texts in Mathematics, New York, 1991. R.B. Gardner, The Method of Equivalence and Its Applications, CBMS- NSF Regional Conference Series in Applied Mathematics, 58, Society for Industrial and Applied Mathematics, Philadelphia, 1989. M. Golubitski, Primitive actions and maximal subgroups of Lie groups, J. Diff. Geometry, 7, 175-191, 1972.
414 Bibliography R. Guenon, The Reign of Quantity and the Signs of the Times, Penguin Metaphyscial Library. This is a reprint of the 1952 translation by Lord Northbourne of the French original published by Edition Gallimard, Paris, 1945. D. Howard and J. Stachel, ed., Einstein and the History of General Relativity, Birkhauser, Boston, 1989. J. Humphreys, Introduction to Lie Algebras and Representation Theory, Springer-Verlag, New York, 1972. D. Husemoller, Fiber Bundles, McGraw-Hill, New York, 1966. H. Jacobowitz, An Introduction to CR Structures, Math. Surveys and Monographs, 32, Amer. Math. Soc, Providence, RI, 1990. N. Jacobson, Lectures in Abstract Algebra, 2, Van Nostrand, London, 1953. N. Kamran, K. Lamb, and W. Shadwick, The local equivalence problem for y" = F(x, y, yf) and the Panleve transcendants, J. of Diff. Geometry, 22, 139-150, 1985. B. Kamte, Cercles generalises et completude dans la geometric de Cartan, Master's thesis, University of Quebec at Montreal, 1995. R. Kirby, The Topology of Four Manifolds, Springer-Verlag, New York, 1989. F. Klein, from Klein's Erlangen Program of 1872, English translation "A Comparative Review of Recent Researches in Geometry" translated by M.W. Haskell, Bull. A MS (formally Bull, of the New York Math. Soc), July 1893. S. Kobayashi and K. Nomizu, Foundations of Differential Geometry, Vol 1, Interscience, New York, 1963. B. Komrakov, A. Churyumov, and B. Doubrov, Two Dimensional Homogeneous Spaces, Preprint Series, Inst, of Math., Univ. of Oslo, 1993. A. Kosinski, Differential Manifolds, Academic Press, Boston, 1993. Ursula K. Le Guin, Excerpt from The World is a Forest, ©1972 by Ursula K. Le Guin; first appeared in Again, Dangerous Visions; reprinted by permission of the author and the author's agent, Virginia Kidd. L. Loomis and S. Sternberg, Advanced Calculus, Addison-Wesley, 1968. Y. Manin, Strings, Math. Intelligencer, 11, no. 2, 59-65, 1989. P. Meyer, Qu'est ce qu'une differentielle d'ordre n? Expo. Math. 7, 249-264, 1989. J. Milnor, Topology from the Differentiable Viewpoint, Univ. of Virginia Press, Charlottesville, Virginia, 1965. P. Molino, Riemannian Foliations, Birkhauser, Boston, 1988. M. Monastyrski, Riemann, Topology and Physics, Birkhauser, Boston, 1987. H. Ozeki and M. Takeuchi, On some types of isoparametric hypersurfaces in spheres I, Tokohu Math. J. 127, no. 2, 515-559, 1975. H. Ozeki and M. Takeuchi, On some types of isoparametric hypersurfaces in spheres II, Tokohu Math. J. 128, no. 2, 7-55, 1976.
Bibliography 415 Η. Poincare, Science and Hypothesis, Dover, 1952, page 50. This is a translation of the French original published in 1902. W.A. Poor, Differential Geometric Structures, McGraw-Hill, New York, 1981. M. Postnikov, Lectures in Geometry, Semester V, Lie Groups and Lie Algebras, Mir, Moscow, 1986. E. Ruh, Cartan connections, Proc. Sympos. Pure Math. 54, part 3, A.M.S., Providence, RI, 1993. P. Samuel, Projective Geometry, Undergradute Texts in Mathematics, Springer-Verlag, New York, 1988. I. Singer and J. Thorpe, The curvature of 4-dimensional Einstein spaces, 355-366, in Global Analysis, Papers in Honor of Kunihiko Kodaira, edited by D. Spencer and S. lyanaga, Princeton Math. Series No. 29, 1969. M. Spivak, A Comprehensive Introduction to Differential Geometry, Vol. 1 & 2, second edition, Publish or Perish Press, Houston, Texas, 1970 and 1979. N. Tanaka, On the equivalence problem associated with simple graded Lie algebras, Hokkaido Math. J. 8, 23-84, 1978. C.-L. Terng, Recent progess in submanifold geometry, Proc. of Symp. in Pure Math. 54, Part 1, 1993. G. Thorbergson, Isoparametric foliations and their buildings, Ann. Math. 133, 429-446, 1991. H. Weyl, Space, Time, Matter, Dover Publications, New York, 1952 (a reprint of Weyl's book, the first edition of which was originally published in German in 1918). J.E. White, The Method of Iterated Tangents with Applications in Local Riemannian Geometry, Pitmann, Boston, 1982. J.H.C. Whitehead, Locally homogeneous spaces in differential geometry, Ann. of Math. 33, no. 2, 681-687, 1932. H. Whitney, Differentiable Manifolds, Annals of Math. 37, no. 3, 645-680, 1936. T. Willmore, Note on embedded surfaces, An Stiint. Univ. ((Al I. Cusa" Iasi s. I. a. Mat. 11, 493-496, 1965. T. Willmore, Riemannian Geometry, Clarendon, Oxford, 1993. C.N. Yang, Magnetic monopoles, fibre bundles and gauge fields, An. New York Acad. Sci. 294, 86-97, 1977. H. Yamabe, On an arcwise connected subgroup of a Lie group, Osaka Math. J. 2, no. 1, 13-14, 1950. K. Yamaguchi, Differential systems associated with simple graded Lie algebras, in Progress in Differential Geometry, 413-494, in the series Advanced Studies in Mathematics, 22, ed. by K. Shiohama, Math. Soc. of Japan, 1993. C.T. Yen, Sur la connexion projective normale associee a un feuilletage du deuxieme ordre, Ann. Mat. Рига Appl. 34, no. 4, 55-94, 1953.
Index Action effective, 138 free, 145 proper, 145 transitive, 138 Ado's theorem, 127 affine coordinate system, 5 affine plane, 141 atlas, 5 Cartan, 174 g-codimensional foliated, 82 Bianchi identities first and second, 193, 236 Bianchi submodule, 232 bundle, 29 G atlas, 31 G bundle, 31, 32 G structure, 32, 33 automorphism, 30 base space, 29 construction of G bundles, 35 effective G bundle, 31 exterior power, 38 flat, 32 induced or pullback bundles, 32 isomorphic G bundles, 33 isomorphism, 30 principal, 36, 37 product bundle, 32 quotient bundle, 38 reduction, 147 section, 38 smooth, 29 tangent, 46 total space, 29 transition function, 31 trivial or product, 30 trivialization, 30 vector bundles, 37 Whitney sum, 38 Calculus, fundamental theorem, 66 canonical line bundle, 31, 268 Cartan atlas, 174 Cartan connection, 181 Cartan gauge, 174 Cartan geometry, 176, 184 admissible frames, 191 Bianchi identity, 192, 193 connection, 181 constant curvature, 221 covariant derivative, 198 curvature function, 191
418 Index Cartan geometry (continued) curvature, 184 curvature section, 192 effective, 176, 184 first-order, 191 flat, 177 generalized circle, 210 geometric orientation, 203, 207 geometrically oriented, 207 higher-order, 191 holonomy, 210 isomorphism, 176, 185 local isomorphism, 185 locally symmetric, 225 moduli space of, 217 normal, 202 of g-curvature type V, 201 of curvature type V, 201 primitive, 184 reductive, 184, 197 space form, 221 special curvature, 201 tangent bundle of, 188 tensors of type (V, p), 194 torsion free, 184 torsion, 184 Cartan structure, 176 Cartan's formula, 263 Cauchy-Riemann operator, 403 charts, 5 compatible, 6 compatible along a path, 7 Clebsch-Gordan problem, 398 collinear points in Pn, 332, 337 complete 1-form, 129 vector field, 69 complex projective line, 144 conformal arclength, 320 curvature, 320 metric, 277 transformation, 271 conformally flat, 236 constant curvature, 236 constant of integration, 115 contact elements, 354 covariant constant, 198 covariant derivative, 371 base version, 199 universal, 194 CR geometry, 355 curl, 404 curvature, 176 Darboux derivative, 115 derivation, 50 at p, 42 development, 368, 377 of a g-valued 1-form, 119, 120 diffeomorphism, 11 differential equation, 331, 354 differential form 1-form, 52 p-forms, 54 basic and semibasic, 62 transforming according to a representation, 195 differential operators, 195 Dirac operator, 405 directional derivative in Rn, 42 distribution, 74 associated to a foliation, 82 integral submanifold, 75 involutive or integrable, 74 local basis, 74 divergence, 404 dual principal curvature, 256 Ehresmann connection, 235, 358 curvature of, 358 Einstein manifold, 236 elliptic curve, 24 elliptic plane, 141 embedding, 16 weak, 17 equivalence problem for conformal metrics, 285 for Weyl structure, 292 Euclidean n-space, 242 Euclidean geometry, 227, 234 Euclidean group, 228 Euclidean model, 228 Euclidean plane, 140 exterior derivative of a p-form, 56 of a function, 52 exterior product of p-forms, 55
Index 419 Families of submanifolds, 354 fiber, 29 flow generated by a vector field, 73 foliation, 82 characterization of simple, 90 leaf, 82 production or trivial, 83 simple, 90 vertical foliation of a bundle, 83 frame bundle fundamental forms, 363 Frobenius theorem, 76 theorem, differential form version, 81 fundamental theorem of calculus, 124 Gauge, 165 gauge symmetry, 169, 178 generalized circle, 337 geodesies, 236, 344 configuration of, 352 geometric rigidity, 216 geometry of similarity, 141 Grassmannian, 147 group of periods, 122 Holonomy group, 210 restricted, 211 holonomy Cartan 210 Ehresmann, 369 of a leaf, 86 homogeneous space, 153 homomorphism, 13 Hopf fibration, 33 horizontal curve, 368 distribution, 198 vector, 269, 368 hyperbolic n-space, 242 hyperbolic plane, 140 Infinitesimal gauge, 167 integral of a 0 valued 1-form, 115 involutes of curves, 359 isoparametric, 259 Jacobian variety, 136 Klein geometry, 151 associated effective, 151 effective, 151 first order, 164 fundamental property, 160 geometric isomorphism, 153 geometrically oriented, 151 higher order, 164 infinitesimal, 156 kernel, 151 Lie theoretic properties, 153 locally effective, 151 one-dimensional effective geometries, 157 primitive, 151 principal group, 151 space, 151 tangent bundle, 163 Klein pair, 156 effective, 156 reductive, 156 Klein space, 153 Leaves, space of, 90 Levi-Civita connection, 235, 358 parallelism, 362 Lie algebra, 103 representation, 127 subalgebra and ideal, 103 Lie group, 12 adjoint action and representation, 110 characterization, 130 completeness of left-invariant fields of, 106 left and right translation, 96 left invariant field v\ 102 left invariant vector fields, 101 Lie group-Lie algebra correspondence, 107, 126 one-dimensional subgroups, 105 product and quotient rules, 113 standard examples, 63 structural equation, 108 vector field X* induced by action of, 107 Lie sphere plane, 143 light cone, 142
420 Index lightlike subspace, 268 vector, 268 line in Pn, 337 locally ambient geometry, 247 locally Klein geometry, 154 locally, flat, 18 Lorentz plane, 141 Mobius band, 30 Mobius geometry, 277 locally ambient, 305 normal, 280 normal submodule, 275 Ricci homomorphism, 275 special curvature, 278 three-dimensional normal, 281 Mobius model, 269 Lie algebra for, 272 Mobius plane, 142 Mobius group, 269 manifold G structure, 48 immersed, 26 orientable, 6 reduction of structure group, 48 smooth, 6 topological, 3 topological orientation, 7 Maurer-Cartan form, 98 mean curvature vector, 256 method of equivalence, 331 model algebra Mobius geometry, 271 Mobius subgeometry, 295 projective geometry, 332, 338 Riemannian geometry, 228 Riemannian subgeometry, 245 model geometry, 174 effective, 174 kernel, 174 primitive, 174 reductive, 174 moduli space of Cartan geometries, 217 monodromy and covering transformations, 128 monodromy, 122 monodromy group, 122 monodromy representation, 122 mutation, 154, 217, 243 n-sphere, 3, 5, 24, 242, 273 no-slip condition, 377 no-twist condition, 377 normal bundle, 247, 248, 305 relatively flat, 254 normal submodule, 202, 275, 336 One-parameter groups, 71, 72 pth exterior power, 38 parallel, 198 period group, 122 defined up to conjugacy, 122 period map, 115 classical, 136 plaque, 18 chart, 18 plethysm, 397 Poincare, first return map, 86 primitive models, 218 principal bundle, 36, 37 reduction, 147 principal curvature decomposition, 256 principal curvatures, 253 constant, 259 principal subspaces, 256 projective completion, 143 projective geometry, 338 curvature function, 340 normal, 342 normal space, 353 special geometries, 341 subgeometry, 353 symmetric, 342 torsion free, 339 projective group, 334 projective model, 334 projective model of Mobius η space, 268 projective, space, 4 proper, 16 Quotient bundle, 38 Rank, 12
Index 421 real projective plane, 142 reduction compatible with a distribution, 258 of principal bundle, 147 reductive, 220 Ricci curvature, 236 Ricci homomorphism, 229 Ricci submodule, 232 Riemannian geometry, 234 locally ambient, 246 Riemannian metric, 227, 238 Riemannian space form, 242 rolling map, 377 Scalar curvature, 236 scalar submodule, 232 second fundamental form, 27, 253, 310 section, parallel, 200 sectional curvature, 243 shape operator, 253 slide map, 84 smooth map, 11 structure, 6 solder form, 235, 363 sphere, 3, 24 submanifold described implicitly, 23 described parametrically, 25 immersed, 18, 26 proper, 23 regular, 22 smooth, 19 topology, 20 subspace, affine, 18 symmetric submodule, 336 Tangent bundle, 41, 46 tangent space, 27 tensor gauge expression of, 194 of type (V», 194 trace zero, 310 traceless Ricci, 236 Umbilic hypersurface, 260 umbilic point, 310 Vector bundle geometric, 190 isomorphism, 37 vector field, 49 u>constant, 129 algebraically involutive or integrable system of, 76 bracket, 51 complete, 69 horizontal lift, 198 linearization, 68 related by a map, 50 vectors analytic, 42 geometric, 39 Wedge product of p-forms, 55 Weingarten map, 253, 310 Weyl curvature, 236 Weyl geometry, 282 Weyl structure, 292 Weyl submodule, 232 Whitney sum, 38 Whitney's embedding theorem, 26 Willmore curvature, 314