Cover
Title page
Date-line
Contents
Preface
Preface to the English Edition
Outline and Goal of the Theory
Chapter 1 Manifolds
6
6
1.2 Definition and examples of manifolds
12
20
22
22
1.3 Tangent vectors and tangent spaces
23
23
24
24
30
30
34
1.4 Vector fields
36
37
1.5 Fundamental facts concerning manifolds
48
Summary
Exercises
Chapter 2 Differential Forms
57
63
2.2 Various operations on differential forms
69
70
2.3 Frobenius theorem
88
2.4 A few facts
90
Summary
Exercises
Chapter 3 The de Rham Theorem
97
3.2 Integral of differential forms and the Stokes theorem
106
3.3 The de Rham theorem
111
111
3.4 Proof of the de Rham theorem
3.5 Applications of the de Rham theorem
Summary
Exercises
Chapter 4 Laplacian and Harmonic Forms
151
4.2 Laplacian and harmonic forms
4.3 The Hodge theorem
4.4 Applications of the Hodge theorem
Summary
Exercises
Chapter 5 Vector Bundles and Characteristic Classes
177
5.2 Geodesics and parallel translation of vectors
5.3 Connections in vector bundles and
185
188
5.4 Pontrjagin classes
199
5.5 Chern classes
208
5.6 Euler classes
211
5.7 Applications of characteristic classes
Summary
Exercises
Chapter 6 Fiber Bundles and Characteristic Classes
232
232
233
233
235
235
236
238
6.2 $S^1$ bundles and Euler classes
246
252
254
6.3 Connections
258
264
264
6.4 Curvature
265
6.5 Characteristic classes
276
278
6.6 A couple of items
Summary
Exercises
Perspectives
Solutions to Exercises
Chapter 2
Chapter 3
Chapter 4
Chapter 5
Chapter 6
References
Index
$\mathbb{C}P^n$
$C^r$ function
$C^r$ map
$\mathbb{C}^\infty$ atlas
$\mathbb{C}^\infty$ diffeomorphism
$\mathbb{C}^\infty$ differentiate homeomorphism
$\mathbb{C}^\infty$ differentiable manifold
$\mathbb{C}^\infty$ function
$\mathbb{C}^\infty$ manifold
$\mathbb{C}^\infty$ map
$\mathbb{C}^\infty$ singular $k$ chain
$\mathbb{C}^\infty$ singular $k$-simplex
$\mathbb{C}^\infty$ singular cochain complex
$\mathbb{C}^\infty$ structure
$\mathbb{C}^\infty$ triangulation
$\mathbb{C}^\infty$ vector field
Diff $M$
Exp $tX$
$G$-structure
$\mathbb{H}^n$
$k$ cochain
$k$-form
$l$-chain
$l$-simplex
$n$-dimensional numerical space
$n$-dimensional sphere
$n$-dimensional torus
$n$-dimensional vector space
$n$-sphere
$P^n$
$\mathbb{R}$, $\mathbb{R}^2$, $\mathbb{R}^3$
$\mathbb{R}^n$
$\mathbb{R}P^n$
$T_x\mathbb{R}^n$
$\epsilon$-neighborhood
A
action of a group
adjoint operator
admissible
Alexander-Whitney map
algebra
alternating
alternating form
anti-derivation
associated bundle
atlas
automorphism group
B
Betti number
Bianchi's identity
boundary
boundary cycle
boundary operator
bracket
bundle map
C
Cartan-Eilenberg theorem
Cech cohomology
cell
chain complex
characteristic class
characteristic number
E
Chem class
Chern number
Chern-Simons form
classes $C^r$, $C^\infty$
classifying space
closed form
closed manifold
cobordant
coboundary
cochain complex
cocycle
cocycle condition
coherent orientation
cohomologous
cohomology
commutative vector fields
compact
complement of a knot
complete
completely integrable
complex Lie group
complex manifold
complex projective space
complex vector bundle
complexification
conjugacy
conjugate bundle
connection
for a complex vector bundle
in a general bundle
in a principal bundle
connection form
contractible
contractible open covering
coordinate change
coordinate functions
coordinate neighborhood
cotangent bundle
cotangent space
covariant derivative
covariant exterior differential
covering
covering manifold
covering map
curvature form
cycle
D
algebra
group
de Rham complex
de Rham theorem
concerning the product
for triangulated manifolds
derivation
diffeomorphism
diffeomorphism group
differentiate manifolds
differential
differential form
coordinate-independent definition
differential ideal
directional derivative
discrete group
distance
distribution
divergence
double complex
dual bundle
dual space
embedding
Euclidean simplicial complex
Euclidean space
Euler characteristic
Euler class
Euler form
Euler number
Euler-Poincare characteristic
exact form
existence and uniqueness of the solution of ODEs
existence of partitions of unity
exterior algebra
exterior differentiation
exterior power bundle
exterior product
F
fiber
fiber bundle
flat connection
flat $G$ bundle
frame field
free
Frobenius theorem
fundamental class
fundamental vector field
G
Gaussian plane
general linear group
general position
geodesic
gradient
graph
Grassmann algebra
Green's operator
H
harmonic function
Hausdorff separation axiom
Hausdorff space
Hirzebruch signature theorem
Hodge decomposition
Hodge operator $\ast$
Hodge theorem
holomorphic mapping
holonomy homomorphism
homeomorphism
homogeneous coordinate
homologous
homology group
homology theory
of cell complexes
of simplicial complexes
homotopy invariance of de Rham cohomology
homotopy type
Hopf invariant
Hopf index theorem
Hopf line bundle
Hopf map
horizontal lift
horizontal vector
hyperbolic space
I
index
induced bundle
induced connection
integrability condition
integral curve
integral manifold
interior product
intersection form
intersection number
invariant polynomial function
inverse function theorem
involutive
isomorphic bundles
isolated singular point
J
Jacobian
Jacobian matrix
K
Kronecker product
Kiinneth formula
L
Laplacian
left-hand system
lens space
Levi-Civita connection
Lie algebra
Lie derivative
Lie group
lift
line bundle
link with two components
linking number
local chart
local coordinate system
positive
locally finite
M
mapping degree
Massey products
triple
Maurer-Cartan equation
Maurer-Cartan form
maximal atlas
maximal integral curve
metric connection
metric space
multilinear
N
Newton's formula
nonzero section
normal bundle
null cobordant
O
one parameter group of transformations
open covering
open neighborhood
open set
open simplex
open star
open submanifold
orbit
orbit space
ordered basis
orientable
orientation
orientation preserving
oriented manifold
orthogonal group
P
parallel along a curve
parallel displacement
partition of unity
Pfaffian
Poincare disk
Poincare' duality
Poincare' lemma
polar coordinates
polyhedron
Pontrjagin class
Pontrjagin form
Pontrjagin number
primary obstruction
principal bundle
principal $G$-bundle
product bundle
product manifold
projection
proof of the de Rham theorem
properly discontinuous
pullback
Q
quotient space
R
reducible
refinement
regular submanifold
restriction of a bundle
Riemannian manifold
Riemannian metric
in a vector bundle
right-hand system
S
section
self-adjoint
signature
simplicial complex
singular $k$-chain
singular $k$-simplex
singular chain complex
singular homology group
singular homology theory
singular point of the vector field
special orthogonal group
stabilizer
standard $k$-simplex
Stiefel-Whitney class
Stokes theorem
on chains
structure constant
structure equation
structure group
subbundle
submanifold
submersion
support
symbol
symplectic form
system of Pfaffian equations
T
tangent frame bundle
tangent space
tangent vectors
topological manifold
topological space
topologically invariant
torsion tensor
total Chern class
total differential
total Pontrjagin class
total space
transition function
triangle inequality
triangulation
trivial bundle
trivial connection
trivialization
U
unitary group
universal covering manifold
universal $G$-bundle
upper half space
V
vector field
vector space
velocity vector
vertical vector
volume element
volume form
W
Weil homomorphism
Whitney formula
Whitney sum
Whitney's embedding theorem
Z

Author: Morita Shigeyuki  

Tags: mathematics   geometry  

ISBN: 0-8218-1045-6

Year: 2001

Text
                    Translations of
MATHEMATICAL
MONOGRAPHS
Volume 201
Geometry of
Differential Forms
Shigeyuki Morita
'•' 11- American Mathematical Society


Translations of MATHEMATICAL MONOGRAPHS Volume 201 Geometry of Differential Forms Shigeyuki Morita Translated by Teruko Nagase Katsumi Nomizu , American Mathematical Society M Providence. Rhode Island
Editorial Board Shoshichi Kobayashi (Chair) Masamichi Takesaki BIBUN KEISHIKI NO KIKAGAKU (GEOMETRY OF DIFFERENTIAL FORMS) by Shigeyuki Morita Copyright © 1997, 1998 by Shigeyuki Morita Originally published in Japanese by Iwanami Shoten, Publishers, Tokyo, 1997,1998 Translated from the Japanese by Teruko Nagase and Katsurni Nomizu 2000 Mathematics Subject Classification. Primary 57Rxx, 58Axx. Abstract. This book is a comprehensive introduction to differential forms. It begins with a quick introduction to the notion of differentiable manifolds, and then develops basic properties of differential forms as well as fundamental results concerning them, such as the de Rham and FVobenius theorems. The second half of the book is devoted to more advanced material, including Laplacians and harmonic forms on manifolds, the concepts of vector bundles and fiber bundles, and the theory of characteristic classes. Among the less traditional topics treated is a detailed description of the Chern-Weil theory. The book can serve as a textbook for an undergraduate or graduate course in geometry. Library of Congress Cataloging-in-Publication Data Morita, S. (Shigeyuki), 1946- (Bibun keishiki no kikagaku. English] Geometry of differential forms / Shigeyuki Morita ; translated by Teruko Nagase, Katsurni Nomizu. p. cm. — (Translations of mathematical monographs, ISSN 0065-9282 ; v. 201) (Iwanami series in modern mathematics) Includes bibliographical references and index. ISBN 0-8218-1045-6 (softcover : alk. paper) 1. Differential forms. 2. Differentiable manifolds. I. Title. II. Series. III. Series: Iwanami series in modern mathematics. QA381 M67 2001 51.V.37—dc21 2001022608 © 2001 by the American Mathematical Society. All rights reserved. The American Mathematical Society retains all rights except those granted to the United States Government Printed in the United States of America (x) The paper used in this book is acid-free and falls within the guidelines established to ensure permanence and durability. Information on copying and reprinting can be found in the back of this volume. Visit the AMS home page at URL: http://www.ams.org/ 10 9 8 7 6 5 4 3 2 1 06 05 04 03 02 01
Contents Preface xiii Preface to the English Edition xvii Outline and Goal of the Theory xix Chapter 1 Manifolds 1 1.1 What is a manifold? 2 (a) The n-dimensional numerical space Rn 2 (b) Topology of Rn 3 (c) C°° functions and diffeomorphisms 4 (d) Tangent vectors and tangent spaces of Rn 6 (e) Necessity of an abstract definition 10 1.2 Definition and examples of manifolds 11 (a) Local coordinates and topological manifolds 11 (b) Definition of differentiable manifolds 13 (c) Rn and general surfaces in it 16 (d) Submanifolds 19 (e) Projective spaces 21 (f) Lie groups 22 1.3 Tangent vectors and tangent spaces 23 (a) C°° functions and C°° mappings on manifolds 23 (b) Practical construction of C°° functions on a manifold 25 (c) Partition of unity 27 (d) Tangent vectors 29 (e) The differential of maps 33 (f) Immersions and embeddings 34 1.4 Vector fields 36 (a) Vector fields 36 (b) The bracket of vector fields 38
viii CONTENTS (c) Integral curves of vector fields and one-parameter group of local transformations 39 (d) Transformations of vector fields by diffeomorphism 44 1.5 Fundamental facts concerning manifolds 44 (a) Manifolds with boundary 44 (b) Orientation of a manifold 46 (c) Group actions 49 (d) Fundamental groups and covering manifolds 51 Summary 54 Exercises 55 Chapter 2 Differential Forms 57 2.1 Definition of differential forms 57 (a) Differential forms on Rn 57 (b) Differential forms on a general manifold 61 (c) The exterior algebra 61 (d) Various definitions of differential forms 66 2.2 Various operations on differential forms 69 (a) Exterior product 69 (b) Exterior differentiation 70 (c) Pullback by a map 72 (d) Interior product and Lie derivative 72 (e) The Car tan formula and properties of Lie derivatives 73 (f) Lie derivative and one-parameter group of local transformations 77 2.3 Frobenius theorem 80 (a) Frobenius theorem — Representation by vector fields 80 (b) Commutative vector fields 82 (c) Proof of the Frobenius theorem 83 (d) The Frobenius theorem Representation by differential forms 86 2.4 A few facts 89 (a) Differential forms with values in a vector space 89 (b) The Maurer-Cartan form of a Lie group 90 Summary 92 Exercises 93 Chapter 3 The de Rham Theorem 95 3.1 Homology of manifolds 96
CONTENTS ix (a) Homology of simplicial complexes 96 (b) Singular homology 99 (c) C°° triangulation of C°° manifolds 100 (d) C°° singular chain complexes of C°° manifolds 103 3.2 Integral of differential forms and the Stokes theorem 104 (a) Integral of n-forms on n-dimensional manifolds 104 (b) The Stokes theorem (in the case of manifolds) 107 (c) Integral of differential forms on chains, and the Stokes theorem 109 3.3 The de Rham theorem 111 (a) de Rham cohomology 111 (b) The de Rham theorem 113 (c) Poincare lemma 116 3.4 Proof of the de Rham theorem 119 (a) Cech cohomology 119 (b) Comparison of de Rham cohomology and Cech cohomology 121 (c) Proof of the de Rham theorem 126 (d) The de Rham theorem and product structure 131 3.5 Applications of the de Rham theorem 133 (a) Hopf invariant 133 (b) The Massey product 136 (c) Cohomology of compact Lie groups 137 (d) Mapping degree 138 (e) Integral expression of the linking number by Gauss 140 Summary 142 Exercises 142 Chapter 4 Laplacian and Harmonic Forms 145 4.1 Differential forms on Riemannian manifolds 145 (a) Riemannian metric 145 (b) Riemannian metric and differentieal forms 148 (c) The *-operator of Hodge 150 4.2 Laplacian and harmonic forms 153 4.3 The Hodge theorem 158 (a) The Hodge theorem and the Hodge decomoposi- tion of differential forms 158 (b) The idea for the proof of the Hodoge decomposition 160 4.4 Applications of the Hodge theorem 162
x CONTENTS (a) The Poincare duality theorem 162 (b) Manifolds and Euler number 164 (c) Intersection number 165 Summary 166 Exercises 167 Chapter 5 Vector Bundles and Characteristic Classes 169 5.1 Vector bundles 169 (a) The tangent bundle of a manifold 169 (b) Vector bundles 170 (c) Several constructions of vector bundles 173 5.2 Geodesics and parallel translation of vectors 180 (a) Geodesics 180 (b) Covariant derivative 181 (c) Parallel displacement of vectors and curvature 183 5.3 Connections in vector bundles and 185 (a) Connections 185 (b) Curvature 186 (c) Connection form and curvature form 188 (d) Transformation rules of the local expressions for a connection and its curvature 190 (e) Differential forms with values in a vector bundle 191 5.4 Pontrjagin classes 193 (a) Invariant polynomials 193 (b) Definition of Pontrjagin classes 197 (c) Levi-Civita connection 201 5.5 Chern classes 204 (a) Connection and curvature in a complex vector bundle 204 (b) Definition of Chern classes 205 (c) Whitney formula 207 (d) Relations between Pontrjagin and Chern classes 208 5.6 Euler classes 211 (a) Orientation of vector bundles 211 (b) The definition of the Euler class 211 (c) Properties of the Euler class 214 5.7 Applications of characteristic classes 216 (a) The Gauss-Bonnet theorem 216 (b) Characteristic classes of the complex projective space 223
CONTENTS xi (c) Characteristic numbers 225 Summary 228 Exercises 229 Chapter 6 Fiber Bundles and Characteristic Classes 231 6.1 Fiber bundle and principal bundle 231 (a) Fiber bundle 231 (b) Structure group 233 (c) Principal bundle 236 (d) The classification of fiber bundles and characteristic classes 238 (e) Examples of fiber bundles 239 6.2 S1 bundles and Euler classes 240 (a) S1 bundle 241 (b) Euler class of an S1 bundle 241 (c) The classification of S1 bundles 246 (d) Defining the Euler class for an Sl bundle by using differential forms 249 (e) The primary obstruction class and the Euler class of the sphere bundle 254 (f) Vector fields on a manifold and Hopf index theorem 255 6.3 Connections 257 (a) Connections in general fiber bundles 257 (b) Connections in principal bundles 260 (c) Differential form representation of a connection in a principal bundle 262 6.4 Curvature 265 (a) Curvature form 265 (b) Weil algebra 268 (c) Exterior differentiation of the Weil algebra 270 6.5 Characteristic classes 275 (a) Weil homomorphism 275 (b) Invariant polynomials for Lie groups 279 (c) Connections for vector bundles and principal bundles 282 (d) Characterisric classes 284 6.6 A couple of items 285 (a) Triviality of the cohomology of the Weil algebra 285 (b) Chern-Simons forms 287
xii CONTENTS (c) Flat bundles and holonomy homomorphisms 287 Summary 291 Exercises 292 Perspectives 295 Answers to Exercises 299 Chapter 1 299 Chapter 2 302 Chapter 3 305 Chapter 4 308 Chapter 5 310 Chapter 6 311 References 315 Index 317
Preface As the title indicates, this book is an exposition of differential forms. What is a differential form? The aim of this book is to answer that question. To explain differential forms, we have to comment on the differentiable manifolds over which they are defined. In brief, a differentiable manifold is a modern representation of a figure as a geometric object, and is an important notion in modern mathematics. Therefore many textbooks on differentiable manifolds have been published. The reader may have seen or even already studied some of them. In these textbooks, without exception, differential forms are defined. However, in many cases only the definition and the fundamental properties are presented, while only a brief description is given of how they are used to analyze the structure of the differentiable manifold. This is because so many fundamental facts already take up a lot of space. Now that the notion of differentiable manifold is completely established, this situation is in a sense inevitable. However, an inconvenience arises. That is, the reader will be busy studying the fundamental facts of differentiable manifolds, and will be left with little time for practical manifolds. Also, as a theoretically reasonable description is frequently not in the order of historical development, the excitement of the discovery often gets lost. Modern mathematics is now progressing dynamically. In geometry a revolutionary change started in the 1980s, and it continues today with no sign of halting. In such an era of progress, it is very important to understand mathematics as a living system that is ready for new advances, rather than as a completely established system. In this series there is no book entitled "Manifolds". This may be because the series editors took the above facts into account, and desired to lead the readers to the scene of current active mathematical research as quickly as possible.
xiv PREFACE When I started to write this book, I found it much more difficult than I had expected. Since mathematics stands on logic, vague descriptions are not allowed. On the other hand, if I tried to base my explanations on the historical motivations, the book would quickly grow too long. I have tried to find a reasonable compromise, and will leave it to the reader to judge how close I have come to the goal. The contents of this book can be summarized as follows. In Chapter 1 we begin with the definition of differentiate manifolds and the fundamental ideas connected with them, such as tangent vectors, tangent spaces, etc. The description, although minimal, should be sufficient for understanding the rest of this book. In Chapter 2 we introduce differential forms, define their fundamental operations, and then prove the theorem of Frobenius. This theorem gives a necessary and sufficient condition for the integrability of "fields of directions" given at every point of a manifold, which are described by either differential forms or vector fields, and its importance has been increasing recently. The theme of Chapter 3 is the theorem of de Rham. The reader may have heard the name of this theorem. In fact it is a very important result, and we may even say that it serves as the basis of the theory of manifolds. In addition to the ordinary proof, we give an explanation to clarify its relation to the integration of differential forms. Also, several applications of this theorem are given at the end of the chapter. Although they may be somewhat difficult, the author hopes that they will show the reader some of the power of this theorem. The second part of this book begins in Chapter 4, in which we study the relationship between Riemannian metrics and differential forms. We then explain the beautiful theory of harmonic forms, due to Hodge and also to Kodaira and de Rham. Briefly, this theory may be said to give a refinement of de Rham's theorem in the context of Riemannian manifolds. In Chapter 5, we introduce the notion of vector bundles. This is the notion obtained by generalizing the tangent bundle of a manifold, and it is a crucial tool in modern mathematics. We also explain the concepts of connection and curvature, which are used to measure how vector bundles are twisted. In the last chapter, Chapter 6, we explain the theory of characteristic classes, which I would say is one of the highest summits of modern geometry. By virtue of this theory, the structure of figures, namely manifolds, can be expressed in terms of differential forms,
PREFACE xv which are local objects; and if we integrate them, the global structure appears as concrete numbers, called characteristic numbers. Here almost all of the previous material will be used. Those readers who want to know the details of manifolds or homology theory that are used in this book are invited to consult textbooks on those subjects. If this book leads the reader to such a study, or if it awakens an interest in deeper theories, the author will be very happy indeed. Shigeyuki Morita July 1996
Preface to the English Edition This is a translation of my book originally published in Japanese by Iwanami Shoten, Publishers. It aims at introducing the reader to the theory and practice of differential forms on manifolds, assuming only a minimum of knowledge such as linear algebra, calculus, and elementary topology. It also includes a quick introduction to the concept of differentiate manifolds. I hope that this book will provide the reader with a flavor of modern geometry and encourage him or her to proceed to a study of deeper theories. The original Japanese edition was published in two volumes. I am grateful to Mrs. Teruko Nagase for translating the first part (Chapters 1,2,3) while keeping close contact with the author during the work. The second part (Chapters 4,5,6) was translated by Professor Kat- sumi Nomizu, who also suggested several improvements on the text. I would like to express my deep gratitude to him. Finally, I would like to thank the American Mathematical Society for publishing this English edition, and their staff for providing excellent support. Shigeyuki Morita January 2001
Outline and Goal of the Theory Geometry is the science of figures. We study various properties of figures, and classify given figures according to the results. We have the notion of invariants, which can serve as the most effective method of classification. We may briefly say that invariants describe geometric structures in terms of numbers. For example, as is well known, the condition for congruence of triangles is described by such invariants as the length of edges and angles at vertices. However, in geometry we are not always studying given figures. Sometimes we ask what kind of figures can exist at all, and also enumerate conditions for their existence. It may be an eternal problem, both in physics and in mathematics, for humans to imagine the figure and shape of the universe where we live, and to study those conditions. In geometry, in some cases, we can even surprise people by constructing unknown figures practically. This is one of the pleasures of studying geometry. The appearance of non-Euclidian geometry is surely a typical example, which shows that the axiom of parallels is not true in general. The figures that are treated in modern geometry are called manifolds. It is usually said that the notion of manifolds was introduced by Riemann in his inaugural lecture at Gottingen University in 1854. In this talk, the geometry of manifolds with Riemannian metrics - namely, differential geometry - was also initiated. This was an epoch- making lecture, in advance of its time. We also owe a great deal to a series of work by Whitney, begun in the 1930's, for the formulation of manifolds that we are now using. Although they are all called "manifolds", there are various kinds of manifolds. The simplest are the topological manifolds, which we only require to be locally homeomorphic to a Euclidean space. However, a manifold usually means a differentiable manifold, which has smoothness; examples include curves and surfaces with beautiful xix
xx OUTLINE AND GOAL OF THE THEORY curved shapes. There are also complex manifolds and algebraic manifolds (or varieties), which have finer structures. Nowadays each kind of manifold is studied by its own methods. However, we should not forget the origin of all manifolds, back in the days when mathematics was not separated into branches like today. That is, we must consider those manifolds that were considered by our great predecessors such as Gauss, Riemann and Poincare. As a simple but very important example, let us consider ori- entable closed surfaces. While we shall give an exact definition later on, for the present just imagine a smooth surface in space which is bounded and has no boundary, as in Figure 0.1. The classification of Figure 0.1. Orientable closed surface these surfaces was already completed at the beginning of this century. There is an invariant which comes to mind from a first glance at the figure, namely the number of holes, which is called the genus. Then a necessary and sufficient condition that two closed surfaces be the same (in modern language, homeomorphic or diffeomorphic) is that their genera are equal. So if E9 denotes a genus g closed surface, then the infinite sequence Eo, Ei, E2,.... exhausts all the orientable closed surfaces. Eo is the sphere and Ei is the surface called a torus. Usually they are denoted by S2 and T2, respectively. It is not an exaggeration to say that all the essence of geometry is contained in the above classification of closed surfaces and in the Gauss-Bonnet theorem mentioned below. Actually, one goal of 20th century geometry was to try to extend these facts to general manifolds in higher dimensions. By the way, even if we consider only surfaces where the conclusion is extremely simple and clear, if we think about its meaning a little more deeply, we find that the problem is not so simple. It may appear that the meaning of the genus g is so obvious from the figure that there is no difficulty in defining it. However, this is only because, in this picture, the figure is positioned in such a way that its genus is clearly
OUTLINE AND GOAL OF THE THEORY xxi shown. In the case of a complicated surface, it is impossible to be sure of the genus at sight, no matter how small it is. Moreover, there is the intrinsically more important point that manifolds are not always located in the well-known Euclidean spaces. In fact, it is characteristic of modern geometry that manifolds are independent of the framework of Euclidean spaces, and became quite free objects. Therefore, when we study them, we cannot always utilize their relative relation to the whole space. Moreover, in the cases of higher dimensional manifolds, it is impossible to observe them directly with our eyes, no matter how much we try to stretch our imagination. So how can we manage them? One possible method would be a combinatorial one where we take certain items as units and decompose manifolds into these items. As the items, we can use points, lines, triangles, and what are called simplices, which are their generalizations to general dimensions. Also we could use cells, which have more flexible shapes. This method is very practical, and historically the first geometric invariant, the Euler number, was found by this combinatorial method. If we decompose a given figure into several triangles and take the alternating sum of the numbers of vertices, edges and triangles, then the total is independent of the method of decomposition, and it constitutes a specific quantity of figures. Many readers may know that in the case of a genus g closed surface, it is equal to 2 - 2g, and this fact in turn indicates that the genus could be defined by a combinatorial method. This is homology theory, introduced by Poincare about 100 years ago, which extended these ideas and became an important means to study figures by combinatorial methods. This theory enables us to count the number of "holes" in each dimension (called the Betti number). Thus the Euler number was given a theoretical basis for the first time by Poincare, and so it is sometimes also called the Euler-Poincare characteristic. In the 20th century, cohomology groups were defined as the dual of homology groups, and, with both of them, a branch of geometry called algebraic topology flourished. Another method originated from the theory of surfaces due to Gauss, as well as the Gauss-Bonnet theorem which followed it, and it uses differentiation and integration to study figures. Although we say simply a genus g closed surface, there are various ways of realizing it in the space. In more mathematical terms, there are various kinds of Riemannian structures on Eg, and we can bend it quite freely. What Gauss showed in his theory of surfaces is that how curved a surface is,
xxii OUTLINE AND GOAL OF THE THEORY now called the curvature, is an intrinsic quantity of the surface and can be defined apart from the space where it lies. Because of this, it may be said that he set the stage for the above-mentioned work of Riemann. The Gauss-Bonnet theorem claims that if we integrate the curvature K of a genus g closed surface S, which is curved arbitrarily, over the whole surface, then the result is a constant independent of how it is curved, and that constant is 2ir times the Euler number x(S)- If we express this in mathematical terms, we obtain the following beautiful equation: f Kda = 2nX(S). Now if we try to describe the goal of modern geometry in one sentence, we may say that this goal is to extend the classification of closed surfaces and also the Gauss-Bonnet theorem to manifolds of arbitrary dimension in various ways. Here differential forms played a fundamental role. First of all, the theorem of de Rham claims that the homology as well as the cohomology groups, which are defined by combinatorial methods, can be obtained using differential forms in the case of differentiable manifolds. But then, how does it go? Elements in the fc-dimensional homology group of a manifold are represented by so-called fc-dimensional cycles. A cycle is literally a figure without boundary which returns to itself. In the cases where k = 0,1,2, a cycle may be understood to be a point with ± signs, an oriented closed curve, and an oriented closed surface, respectively (see Figure 0.2). On the other hand, what is a fc-form on a manifold? When k — 0, it Figure 0.2. Cycles is simply a function. Therefore it takes a value on any 0-dimensional cycle. For general k > 0, it may be said that a fc-form is a kind of function that has a value on any ordered ^-directions (that is, tangent vectors) at each point of a manifold. Therefore we can integrate it over any fc-dimensional cycle, and we obtain a certain quantity. We can repeat the theorem of de Rham by saying that we can obtain the (co)homology groups, with coefficients from R. of any differentiable manifold completely by such an operation of integrating differential
OUTLINE AND GOAL OF THE THEORY xxiii forms on cycles. As above, since differential forms are something like functions defined on any ordered directions at each point, it might be easy to understand that they can describe various geometric structures on manifolds. In the case of surfaces, although the curvature, which tells how much it is curved, is a function on the surface, it will be more natural to consider it as a differential form Kda of degree 2 - that is, a combination of the curvature K and da which is called the areal element. It is this 2-form that can be generalized to the cases of higher dimensional Riemannian manifolds. It is called the Riemannian curvature form, and it expresses how a manifold is curved explicitly. Going back a little bit, we have the tangent space and the tangent bundle, which provide the most important tool to analyze the structure of manifolds. The collection of all the tangent vectors at a point is the tangent space, which gives the first approximation describing the state of neighborhoods of that point, and the collection of all these tangent spaces over the whole manifold is the tangent bundle. Therefore, we can say that the tangent bundle is a space made of vector spaces, which are flat spaces, over each point of a manifold. The way these spaces are connected to each other is controlled by the group of all the automorphisms of the vector space, which is a Lie group called the general linear group. Generalizing this idea, we obtain the notion of fiber bundles, which is motivated mainly by the great work of E. Cartan in the first half of the 20th century. Briefly speaking, a fiber bundle is a manifold obtained by tying together a family of manifolds, called fibers, which stand systematically over each point of another manifold (see Figure 0.3). In fiber bundles, the group (called the structure group) that controls connections between fibers is an infinite-dimensional group in general, but the cases where it becomes Figure 0.3. Fiber bundle
xxiv OUTLINE AND GOAL OF THE THEORY a Lie group are especially important. Then there arose an important method for studying the structure of differentiable manifolds, and that is to consider various fiber bundles over a given manifold with various Lie groups as their structure groups, and to investigate them from a synthetic point of view. Now, how many fiber bundles, with a given Lie group as the structure group, are there on a manifold? This is a fundamental question, and it is the theory of characteristic classes that answers it. Roughly speaking, characteristic classes are a certain description of how fiber bundles are twisted over a manifold in terms of its cohomology groups. The characteristic classes called Chern classes or Pontrj agin classes are typical examples. There are various approaches to this theory; among them the Chern-Weil theory is important. There is a general method, where we give a relation between the fibers of a fiber bundle in terms of a certain differential form of degree 1, called the connection, and then differentiate it to obtain a quantity called the curvature, which describes how the fiber bundle is curved. The Chern-Weil theory gives a beautiful framework for research by systematically applying this method to fiber bundles with arbitrary Lie groups as their structure groups. We cannot overestimate the important roles which Chern classes and Pontrjagin classes played in classifying and analyzing the structure of differentiable manifolds. For example, we have characteristic numbers that are obtained by integrating polynomials in them over manifolds, which are generalizations of the Euler number. They can express global structures of manifolds in terms of numbers quite explicitly. In modern geometry, the importance of characteristic classes is still increasing. Moreover, they are going to play a deeper role through detailed analysis of differential forms, rather than merely cohomology classes, which express the above classes. For example, in the 21st century, there will be grand attempts to generalize the theory of harmonic integrals, which describes the relationship between the de Rham cohomology and Riemannian metrics, in wider frameworks. In these new developments, it is not too much to say that, so to speak, differential forms play the role of water and air for life.
CHAPTER 1 Manifolds In this chapter, we give an exposition of differentiable manifolds, on which our leading characters, differential forms, are defined and act. Roughly speaking, a differentiable manifold is a smooth figure or space. For example, curves and surfaces are differentiate manifolds. Since a point on a curve can be described by a single parameter, it is called a 1-dimensional manifold. Similarly, every surface can be locally obtained by slightly rounding a small domain in a plane. Then a point can be denoted by the ordinary xy coordinates in the plane. Thus, a point on a surface can be described by 2 parameters. So, a surface is called a 2-dimensional manifold. In general, it is impossible to obtain the whole given surface, however skillfully we round a domain in the plane. To obtain it, we must glue several domains. In other words, the above-mentioned coordinates are not always defined over the whole surface, unlike the planar case. These coordinates are called local coordinates. Smoothness of a surface is reflected in the relationship between different local coordinates. Differentiable manifolds are defined by extending those properties of curves and surfaces to general dimension. That is, a differentiable manifold is a topological space such that any point on it has a neighborhood whose points can be described by local coordinates consisting of n independent parameters, and the relationship among different local coordinates can be described by differentiable functions. Now, unlike a curve or surface, a manifold does not always appear in a well known space. In many cases it is rather difficult to think of it as a geometric figure, since it is born in an extremely abstract framework. However, it often happens that when we find local coordinates on such a set and study the relationship among them, a hidden geometric structure gradually comes to light. Since we try to include as many objects as possible, it is inevitable for the definition of manifolds to be abstract. However, once an object is recognized to be a manifold by virtue of this abstract definition, it appears in
2 1. MANIFOLDS a known space through the coordinates, and turns out to be a very practical object. In the study of not only manifolds but also modern mathematics in general, it is important to connect abstract ideas with concrete examples so that they mutually enrich each other. We will carry out our exposition with this fact in mind. 1.1. What is a manifold? (a) The n-dimensional numerical space Rn. Before we define manifolds, we shall give some fundamental examples. First, let R be the set of real numbers. If we consider it geometrically as the real line, it is a 1-dimensional manifold. Next, the set R2 = {(x,y); x,yeR} of all points with coordinates (x,y) (that is, the xy-plane R2) is a 2-dimensional manifold. In the same way, the set R3 = {(x,y,z);x,y,zeR} of all points with coordinates (x,y,z) (that is, xyz-space R3) is a 3-dimensional manifold (Figure 1.1). Generally, the set Rn = {x= (x1,x2>--- ,xn); Xi eR} of all the n-tuples (xi,X2,--- ,xn) of real numbers is called an n- dimensional numerical space . Rn is the most fundamental n-dimensional manifold. As for its geometric image, it is a space extending boundlessly in n independent R3 S <*».«> w (x.y) Figure 1.1. R, R2,R3 directions. A general n-dimensional manifold is formed by smoothly glueing, one by one, some (in general, infinitely many) domains in Rn (we shall give the details later). Therefore, let us first review the geometric properties of Rn and the fundamental facts of differentiable functions defined on Rn which are to be used as the glueing maps.
1.1. WHAT IS A MANIFOLD? (b) Topology of Rn. For two given points x = (xi,X2, • • • , xn), y — (yi,y2, Rn, the distance d(x,y) between them is defined by d(xty) = >/(*! - ViJ + {X2 - y2J + ¦ • ¦ + (xn - ynJ- The distance d(x, 0) between the point x and the origin is sometimes written ||x|| for short. In the cases of n = 1,2 and 3, d(x,y) is the ordinary length of the line segment xy connecting 2 points x,y, and the above formula is the natural extension of this to general n. It is easy to see that d(x,y) satisfies the following three fundamental properties: (i) d(x, y) > 0 if x ^ y, and d(x, x) = 0. (ii) d(x,y) =d(y,x). (iii) For 3 arbitrary points x,y, z, d(x,y) + d{y,z) > d(x,z) (triangle inequality). Thus, Rn is a metric space with distance d. The distance d also defines a topology on Rn as follows, and with this topology, Rn is a topological space. For a point x in Rn and a positive number e > 0, the set t/(x,e) = {y GRn; d(x,y) < e) of all points whose distance from x is less than € is called the e- neighborhood of x (Figure 1.2). A subset U of Rn is called an open set, if, for an arbitrary point x in U, we can take a sufficiently small c-neighborhood of x so that it is entirely contained in U. For example, all e-neighborhoods are open sets. An open set containing x is sometimes called an open neighborhood of x. U(x,e) l v--'' \ Figure 1.2. e-neighborhood and open set Let U be the set of all open sets of Rn. Note that the empty set 0 is considered to be an open set. Then, it is easy to see that U satisfies the following three conditions:
4 1. MANIFOLDS (i) Rn, 0 € U. That is, the whole space Rn and the empty set 0 are open sets, (ii) If Uu #2, • • • , Uk € W, then CA n U2 n • • • n t/* € W. That is, the intersection of a finite number of open sets is an open set. (iii) For an arbitrary family {Ua}Q€A (Ua € U) of sets belonging to U, their union [Ja€A Ua G U. That is, the union of an arbitrary number of open sets is an open set. In general, when a family U of subsets of X satisfies the above three conditions (where Rn is replaced by X), we say that a topology is defined on X and call the sets in li open sets. Thus, our n-dimensional numerical space Rn turns out to be a topological space that is prerequisite to a manifold. (c) C°° functions and diffeomorphisms. When we overlap parts of two small pieces of paper and glue them together, we obtain a larger piece of paper. Also, when we paste two parts of a piece of paper nicely, we can make various surfaces. For example, when we roll up a rectangular sheet of paper and glue both ends together, we have either a cylindrical ring or a surface called a Mobius strip (Figure 1.3). As mentioned before, roughly speaking, a manifold is a figure obtained by repeating the above kind of operations with open sets in Rn as parts. We shall formulate these operations of overlapping and glueing mathematically. Figure 1.3. Forming surfaces by glueing "Overlapping and gluing" two open sets U, V in Rn is read as identifying U and V by a homeomorphism <p : U —> V. Here, <p being a homeomorphism means that it is a one to one and onto map, and both (p and <p_1 are continuous. Then U and V may be considered
1.1. WHAT IS A MANIFOLD? 5 to have the same shape, as topological spaces. Next, to construct a differentiable (that is, smooth) manifold, the glueing maps have to be smooth. This role will be played by diffeomorphisms, to be defined below. A function / : U —> R defined on an open set U of Rn is said to be of class Cr if all the partial derivatives up to order r exist and are continuous. Such a function is called a Cr function. Functions which are of class Cr for all r are said to be of class C°° and are called C°° functions. Thus, a C°° function is a function which can be differentiated freely. Next, consider a map <p : U —¦ Rm defined on an open set U of Rn and with values in Rm. This map can be described by m functions tpi : U -» R (i = 1, • • • , m) as follows: <p(x) = {<pl{x),--- ,v>m(x)) {xeU). If all <pi are of class Cr (or class C°°), <p is called a Cr map (or a C°° map). The composition of two C°° maps is also a C°° map. This follows from the chain rule of composite functions. Definition 1.1 (Diffeomorphism). Let U,V be open sets in Rn. A homeomorphism <p : U —* V from U onto V is called a C°° differentiable homeomorphism, or simply a diffeomorphism, if both (p and <p~l are of class C°°. It is the inverse function theorem that plays an important role in judging whether a given map is a diffeomorphism, and also in practical construction of diffeomorphisms. To describe it, we shall prepare a term. For a map ip = (</?i, • • • , ipm) defined on an open set U with values in Rm, the matrix few few ¦¦•few V^fW few -fe(»v is called the Jacobian matrix of the map <p at the point x € U. When m = n, the determinant of the Jacobian matrix is called the Jacobian.
6 1. MANIFOLDS Theorem 1.2 (Inverse function theorem). Let ip be a C°° map from an open set U of Rn to Rn. If the Jacobian at a point x in U is not 0, then there exists an open neighborhood V C U of x such that if{V) is an open set and if is a diffeomorphism from V onto <p{V)- Even if a given C°° map from Rn to Rn happens to be known, for some reason, to be a one to one map on an open set C/, it is generally difficult to get the inverse map practically. However, if we calculate the Jacobian of this map and find out that it is not 0 at each point on U, by virtue of the above theorem the inverse map is also a C°° map, and we can conclude that this is a diffeomorphism. In this sense, the inverse function theorem is important. Example 1.3. As a simple example, consider a map i/>: R2 -» R2 defined by <p(x,y) = (x2 - y2,2xy). Since the Jacobian matrix of <p is J, its Jacobian is 4(x2 + y2). Thus, the Jacobian of <p does not vanish at any point outside the origin. By the inverse function theorem, y? is a diffeomorphism if restricted to a sufficiently small neighborhood of that point. However, this map is not a one to one map on the open set obtained from R2 by removing the origin. This is because <p{x,y) = <p(-x, -y). This map over the above open set is a so-called double covering map, as will be explained later. (d) Tangent vectors and tangent spaces of Rn. Rn is also considered to be an n-dimensional vector space. In this case, an element x of Rn expresses a point on the numerical space and, at the same time, it is considered to be an (n-dimensional row) vector connecting the origin to this point. For two vectors x, y G Rn and a real number a € R, two operations, namely the sum x + y € Rn and multiplication by a real number ax € Rn, are defined, and some fundamental relations between them are satisfied. We recall that Rn with these structures is called a vector space. Here, to be more geometric, we will understand it as the set of all arrows (that is, n-dimensional vectors) emanating from the origin, and denote it by ToRn to distinguish it from the numerical space. Here the subscript 0 indicates the origin. We call TbRn the tangent space of Rn at the origin, and an element of it (that is, a vector starting from the origin) is called a tangent vector of Rn at the origin. For a general point x other than the origin, we denote the set of all vectors starting from x by TxRn, and call this the tangent space of Rn at
1.1. WHAT IS A MANIFOLD? x and its elements the tangent vectors at x. A structure of an n-dimensional vector space over R is induced on TxRn naturally. Figure 1.4. Tangent vectors We now choose a basis of the tangent space TqW1. The unit vector of length 1 starting from the origin in the direction of positive Xj is written _d_ dXi The reason for using this symbol should be made clear by the following explanation. By definition, -— is a tangent vector to Rn at the origin, axi that is, an element of TbRn. It is easy to see that -—, • • • , -— is a ox i dxn basis of ToRn. Then, an arbitrary tangent vector v at the origin can be uniquely written as a linear combination d d v = oi^— + ••• + an-—. oxi oxn For a general point x in Rn, the parallel translation of the tangent vector -— to x as the initial point is written as OXi \dxiJx This turns out to be a tangent vector at the point x, and it is obvious that ( -— ) >¦••,( ~— ) forms a basis of the tangent space TxRn at \OX\/x \oxnJx x. With this notation, the tangent vector -— at the origin should be OXi written precisely as (-z— ) . However in the case where the relevant V OXi ' o point is clear in advance, we sometimes simply write ^— instead of OXi (—) \dXiJx It may seem that the above description, which is meant to distinguish the two aspects of Rn, that is, Rn as a geometric figure and
8 1. MANIFOLDS Rn as a vector space, makes the matter difficult. However, this is because Rn is originally a "straight" space, so that the tangent space at an arbitrary point on it looks the same as the original space Rn. In the case of a general curved manifold, the tangent space (see §1.3 for the definition) at a point is often far different from the whole figure. However the tangent space is locally a good approximation of a neighborhood of each point on the manifold, and gives us an important foothold to study the structure of manifolds. Here, we shall cite two important roles played by a tangent vector. These two roles, each of them characterizing the tangent vectors, give a guideline to defining the tangent vectors of a general manifold (where it is not always possible to draw arrows like on Rn). The first role is as a velocity vector to a curve. A smooth curve on Rn is expressed by a C°° map c : R —> Rn. Then the velocity vector at a point c(t) (t € R) on the curve is denoted by S«-(?«••••¦?«)• where c = (ci, • • • , Cn). It is also written as c(t). This velocity vector is read to be a tangent vector at the point c(t) (Figure 1.5). When we move the point over the curve and also consider various curves, a variety of tangent vectors appear at each point of Rn. Figure 1.5. Velocity vector The second role of the tangent vector is as a directional derivative. For example, let a function /(xi,--- ,xn) of n variables be given. Then various partial derivatives -— (i = 1, • • • , n) are con- oxx sidered, and each of these is thought of as a partial derivative in the positive direction of each axis x». Generalizing this idea, for an arbitrary tangent vector A.1) v = a!— + ... + an__ oxx dxn at the origin of Rn, it will be natural to define the partial derivative v(f) of the function / at the origin in the direction of v by v(/)=ai^@)+*,+an^:@)-
1.1 WHAT IS A MANIFOLD? 9 This tangent vector v can be regarded as a tangent vector not only at the origin but also at a general point x in Rn. To distinguish this from the original v, let us write vx. Then vx(f) is thought of as a description of the partial derivative Vx(/)=a1^-(x) + .-- + an^-(x) of / at x in the direction of vx. If we let vf(x) — vx(f), then vf is a function on Rn, and its value at x is vx(f), that is, the partial differential coefficient of / at the point in the direction of v. Thus vf is nothing but the partial derivative of / with respect to v. From the explanation so far, the reason why the notations of the partial derivatives are used to describe the tangent vectors should be clear. Above, we cited two roles of the tangent vectors; and, in general, the velocity vector of a curve widely varies in both direction and size depending on the point. Also, when we consider the partial derivative of a function, we need not keep the direction of differentiation constant. In some cases it is convenient to consider the partial derivative of a function in a direction which depends on the place. Thus the notion of vector field arises. Practically it is defined as follows. A vector field X on Rn is an assignment of a tangent vector Xx € TxRn to each point' x in Rn. Here we use the usual notation Xf instead of v, to describe vector fields. Since the unit vectors -— {i = 1, • • • , n) in the positive direction of each Xj-axis form a OXi basis of the tangent space TxRn at each point x, an arbitrary vector field X can be written as (L2) * = />z?r+-+'«sl;- This formula is formally the same as A.1). However, while in A.1) each coefficient is constant, in A.2) fc is a function on Rn, and the direction and size of X change on each point depending on the value of fi. When all the /t are C°° functions, X is called a C°° vector field. We illustrate a simple example of a C°° vector field on the plane (Figure 1.6). So far we have seen that for a function / and a vector field X on Rn, the derivative Xfolf'va the direction X is defined.
10 1. MANIFOLDS Figure 1.6. Vector field on the plane (e) Necessity of an abstract definition. In the next section we shall give a practical definition of manifolds which may seem to be quite abstract. As we mentioned before, manifold is a notion obtained to generalize curves and surfaces in space to higher dimensions. Therefore one might think it enough to work only in the numerical space Rn. Actually, all manifolds, even if they are defined abstractly, can be realized as generalized surfaces (called sub- manifolds) in Rn after all (Whitney's embedding theorem). However, there is a reason to give an abstract definition. The reason is that even if it is embeddable in Rn, the embedding is not always a natural one and does not always reveal the structure of the manifold. Rather, there is a danger that it may conceal symmetry or other things that might have existed in the manifolds. For example, the sets of (i) all lines in the plane, (ii) all planes in space, (iii) some particular patterns on a surface, and (iv) special curvatures of a surface, turn out to be manifolds with a beautiful and rich structure. These manifolds may not come out if we stick only to Rn. As an example which is simple but full of interesting suggestions, we consider the set of all triangles. However, this is too vague to capture as it stands. For instance, for, say, equilateral triangles with sides 10 centimeter long, it is too much if we distinguish one on the blackboard from a copy on our notebook. We should identify congruent ones. Here we also identify similar ones, and consider the set T of all classes of similar triangles. An equivalence class of similar triangles is determined if three angles are specified. If a,0 and 7 are the angles, they satisfy the conditions a + /? + 7 = 7r, a,/?,7>0. Then T is realized as a domain such that x,y,z > 0 over the plane x + y + z = n inR3. This domain is the interior of an equilateral
1.2. DEFINITION AND EXAMPLES OF MANIFOLDS 11 triangle in that plane. Actually, we should consider the exchange of names of the three angles, for instance a and C. It is easy to see that this corresponds to the action of the group of congruences (consisting of rotations around the origin by 0°, 120° and 240° and reflections with respect to the three axes of line symmetry) of the equilateral triangle. We now realize T in another way. We apply a similarity to expand or shrink a given triangle so that the two vertices A, B are in the position of 0 and 1 in the complex plane and the third vertex is at the point z in the upper half plane H = {z — x + iy\ y > 0}. Thus an arbitrary point on H represents an equivalence class of similar triangles. If we put B, C in the position of 0 and 1 instead of A, B, then a brief computation shows that A is at . In the same way, if we put C, A in the position of 0 and 1, then B is at . Also if we exchange A and B, C is at 1 - z, and if we exchange two other vertices, we obtain similar formulae. Thus T is realized as a figure obtained by identifying complex numbers in H which are transformed into each other by the above operations (the number of identified points is 6 for a general point, less than 6 for special points, and just one for , which corresponds to the equilateral triangle). Although, because of this identification, T itself cannot be a manifold, we can represent T nicely using H. This example shows that various coordinates can be considered on the same object, depending on the purpose. 1.2. Definition and examples of manifolds (a) Local coordinates and topological manifolds. Let M be a topological space, that is, a figure in the broadest sense. We begin by listing, one by one, the conditions for M to be a manifold. First, let M satisfy the Hausdorff separation axiom. That is, for any two distinct pointsp,q e M, there exist an open neighborhood U of p and an open neighborhood V of q such that U and V do not intersect. Such a space M is called a Hausdorff space. Although there are some important topological spaces that fail to satisfy this separation axiom, it can be said that almost all ordinary figures which are the object of geometry satisfy it. For example, the numerical space Rn and all its subspaces are obviously Hausdorff spaces.
12 1. MANIFOLDS The second condition is that for an arbitrary point p of M, there exists an open neighborhood U of p homeomorphic to an open set V of Rn. Let <p : U —* V be such a homeomorphism. Then, as the image of each point q in U by </? is a point in Rn, it can be written as an n-tuple of real numbers: ?>(<?) = (*i (<?),••• ,xn(q)). This n-tuple is called the local coordinates of q, and U is called a coordinate neighborhood. Moreover, xi, • • • , xn are called coordinate functions defined on U. Thus, every point that is sufficiently Figure 1.7. Local coordinate system (chart) close to p can be uniquely described by n independent parameters, called local coordinates. The pair (U, <p) is called a local chart or local coordinate system. Instead of (U,(p), we sometimes write (U;xi,--- ,xn), using coordinate functions. Recall that chart is a word meaning a "map". By {U,<p) or (U;x\, ¦ ¦ • ,xn), a map describing a neighborhood of the point p is given. We assume one more topological condition on M. This condition, called the second countability axiom, claims that there exists a base of countably many elements for the system of open sets. That is, there exist countably many open sets U\, U2, ¦ • ¦ of M and, for an arbitrary open set U and a point p in it, there is some i such that p G Ui C U. If we consider all rational points (points all of whose coordinates are rational numbers) on Rn and all e-neighborhoods of them, where € runs through all positive rational numbers, we see that Rn satisfies this axiom. In the definition of manifolds, this axiom is not always assumed beforehand. However, as the manifolds we usually deal with satisfy this axiom without exception, we assume it from the beginning in this book. A topological space that fulfills the above three conditions is called a topological manifold.
1.2. DEFINITION AND EXAMPLES OF MANIFOLDS 13 Definition 1.4 (Topological manifolds). A Hausdorff space M which satisfies the second countability axiom is said to be an n- dimensional topological manifold if an arbitrary point on it has an open neighborhood that is homeomorphic to an open set of Rn. Example 1.5. Here is a simple example of a topological space where, among the above three conditions, only the Hausdorff separation axiom is not fulfilled. Define a subset M of the plane R2 by M = {(x,0); x<0}u{(x,l); x > 0 } U {(x, -1); x>0} (Figure 1.8). We give M the following topology, which is different Figure 1.8 from the topology as a subspace of the plane. For every point p € M except the two pointsp+ = @,1), p_ = @,-1), consider the ordinary e-neighborhood Ue{p) — {q € M; d{p,q) < e } and for the exceptional two points, let U?(p±) = {(x,0); -e<x<0}U{(x,±l); 0 < x < e} (with the convention that the double signs are taken in the same order). Now, we invest M with the topology defined by the base of open sets which consists of U€(p) for all points p on M and all positive numbers e. By definition, any open neighborhoods of the two points p+,p~ necessarily intersect, no matter how these neighborhoods are chosen. Thus M is not a Hausdorff space. However, it is obvious that M satisfies the other two conditions. (b) Definition of differentiable manifolds. Let M be a topological manifold. By Definition 1.4, at an arbitrary point on M, there exists a "map" which describes its neighborhood; that is, there exists a local coordinate system that gives an identification of its neighborhood with an open set of Rn. Now, there are rrfany maps, but not all of them are necessary for studying the structure of M. For example, consider the maps which describe the surface of the earth. While we can make all sort of maps according to the purposes in hand, from the viewpoint of studying the whole
14 1. MANIFOLDS surface of the earth, the important thing is whether the several sheets of maps cover all the points of the earth. We call a system of maps satisfying this condition an atlas of the earth. Once an atlas is given, (in theory) we can make all maps. The same is true for manifolds. Thus the following definition is natural. Definition 1.6. Let M be a topological manifold. A family of local coordinate systems <S = {(Ua, <pa)}aeA is said to be an atlas of M if {Ua}a?A is an open covering of M, that is, if the open sets Ua cover the whole of M. In the above setting, we write Va for the image of the homeomor- phism <pa from Ua into Rn. M is covered by coordinate neighborhoods Ua in the atlas «S, and on the other hand, each Ua can be identified with an open set VQ of Rn via the homeomorphism <pa. If we see this the other way, it can be said that a manifold M can be made of open sets VQ of Kn by glueing these parts one by one. We remark that even if we cut off from Rn the part Va used to make M, Rn fills the gap soon and is again complete, so that if the next part Vp intersects "the imaginary hole", Vp can be taken out with complete shape. Rn is, so to speak, a "fountain" from which the parts of manifolds spring. We now study how to glue parts together. Suppose two coordinate neighborhoods UayUp in the atlas S intersect. Then the corresponding two open sets Va,Vp in Rn are to overlap in part to make M. Look at Figure 1.9. In this figure, it is seen that the open sets ipa{Ua H Up) in Va and <pp{Ua n Up) in Vp are glued to each other by the homeomorphism fCa =^o <Pal : <Pa{Ua O Up) -+ fp(Ua fl Up) between them. Figure 1.9. Coordinate change
1.2. DEFINITION AND EXAMPLES OF MANIFOLDS 15 Now, as fpa is a map from an open set of Rn to Rn, it is described as fpa = (/jQ,--,/^a) by n continuous functions f0ct. If we write the local coordinates of an arbitrary point p on Ua 0 17/3, (xi(p),--- ,xn(p)) with respect to {Uay<pa) and (yi(p),--- ,yn(p)) with respect to {Up,tpp), then there is a relation Vi(p) = fpaiziiPh--- ,xn{p)) between them. Thus the homeomorphism fpQ describes the relationship between two local coordinates, so it is called the coordinate change. In the case of topological manifolds, it is enough that coordinate changes or glueing maps are homeomorphisms, and there are no other conditions. So as a figure, it may be said that they do not have a rich structure. Against this, differentiable manifolds are defined by claiming that all coordinate changes are diffeomorphisms. By virtue of this they become globally smooth figures, and it is possible to study them in detail using differentiation and integration. We state the definition. Definition 1.7. Let M be a topological manifold. An atlas S = {(^o» &*)}<*€ a of M is called a C°° atlas if all its coordinate changes fpa = <Pp ° <Pa1 are C°° maps. We also say that the atlas determines a C°° structure on M. A manifold with a C°° structure is called a C°° differentiable manifold or simply a C°° manifold. Although in the above definition the coordinate changes are only claimed to be C°° maps, by the inverse function theorem they are, of course, C°° diffeomorphisms. Two C°° atlases S, T given on a topological manifold M are said to be equivalent if the union SuT is also a C°° atlas. It is easy to see that the union of all C°° atlases equivalent to S is also a C°° atlas. This atlas is called the maximal atlas determined by S. A necessary and sufficient condition for two atlases to be equivalent is that the maximal atlases determined by them coincide. So it will be natural to identify C°° structures on M given by equivalent C°° atlases. To show that a given figure is a C00 manifold, it is preferable to construct an atlas of as small a number of local coordinate systems as possible. However, once this is done, it is convenient to use the maximal atlas collecting all the possible coordinate systems, so that we can exchange the local coordinate systems freely according to our purposes.
16 1. MANIFOLDS Let p be a point on M. A local coordinate system (U, y?) belonging to a maximal atlas of M is called a coordinate system around p, if p belongs to U. We present a simple example of a local coordinate system. Example 1.8 (Polar coordinates). Let U be the domain obtained by removing the nonpositive part of the x-axis {(x,0); x < 0} from the xy-plane R2. Let r be the distance of a given point p — (x,y) in U from the origin, and 0 (—•tt < 9 < -n) the angle from the positive direction of the x-axis measured in the counterclockwise direction. If we define a map y? : U —» R2 by y?(p) = (r,0), then (U,tp) is a local coordinate system of R2. y?(p) is called the polar coordinates of the point p. While here we consider the domain obtained by removing the nonpositive part of the x-axis from R2, we sometimes consider polar coordinates in other domains according to our purposes. Generally a C°° structure does not always exist on a topological manifold, and even if it exists it is not unique (except for the case of dimension 0). However, the failure of the uniqueness here is due to the uninteresting reason that an arbitrary difFeomorphism can be made to be a homeomorphism which is not a diffeomorphism by a small local perturbation. The essential classification of C°° structures is given in terms of diffeomorphisms, and will be mentioned in §1.4. Hereafter, the manifolds that we deal with in this book will be C°° manifolds. We sometimes call a C°° manifold simply a manifold. In the next subsection, we give some important examples of C°° manifolds. (c) Rn and general surfaces in it. Example 1.9. The numerical space Rn is an n-dimensional C°° manifold. As an atlas we can take only one local coordinate system, (Rn,id). Here id denotes the identity map of Rn. Note that, in this case, the coordinate has a meaning on the whole of Rn. We mention another way to look at Rn, namely, to consider Rn as the product of n copies of R. This is a special case of the next example. Example 1.10. Let M, N be C°° manifolds, and let <S, T be their atlases respectively. To begin with, it is easy to see that the product space M x N is a topological manifold. Let SxT = {{UxV1<px V); (?/,?>) € «S, (V,V) e T}. We can see that S x T determines a natural C°° structure on M x TV. We call this the product manifold of M and TV.
1.2. DEFINITION AND EXAMPLES OF MANIFOLDS 17 Example 1.11 (n-dimensional sphere). The set of points in Rn+1 whose distance from the origin is 1, Sn = {x = (xi,--. ,in+i)an+1; x2i + --- + xl+i = 1}, is called the n-dimensional sphere (n-sphere). S1 is the unit circle in the plane and S2 is the unit sphere in space. Let us check that Sn is naturally an n-dimensional C°° manifold. We consider the two points p+ = ((),••• ,0,1), v- = @, ¦•• ,0,-1) on Sn. Let U+ = Sn - p_, U- = Sn - p+; then U+ and U- cover the whole of Sn. The stereographic projection (p+ : U+ —* Rn from the point p_ is a homeomorphism, as we can easily see (Figure 1.10). Similarly, the stereographic projection ip- : U- —> Rn from the point p+ is also a homeomorphism. Then we see that the two local coordinates systems, (?/+,?>+), (*7-,?>-), form a C°° atlas of Sn. p. Figure 1.10. The stereographic projection </?+ Example 1.12. The product Sl x ; • • x Sl of n copies of S1 is denoted by Tn and is called the n-dimensional torus. T2 is the surface of a doughnut, shown in Figure 1.11. The n-dimensional torus is one of the most important manifolds. Figure 1.11. 2-dimensional torus EXAMPLE 1.13 (General surfaces in Rn). The (n - ^-dimensional sphere 5n_1 is defined by the equation x\ H h x\ — 1 = 0 in
18 1. MANIFOLDS Rn. It is natural to generalize this and consider a figure Z consisting of all the points that satisfy m equations fi(xi,--- >*n) =0 (t = l,--- ,m). Here, each fi is assumed to be of class C°°. We put / = (/i, • • • , /m)- Z takes various shapes according to the properties of /. In the extreme case, it may happen to be the empty set. Under what conditions will Z be a smooth figure, that is, a manifold in Rn? We will give one sufficient condition. We are given m equations, each of them considered as imposing a restriction on the n variables X, which are essentially independent parameters in Rn. The m equations give m restrictions, and the degree of freedom decreases by m. As a result, a point on Z is expected to be described by n - m parameters. However, if we consider the case of /i = /2, it is easy to see that this is not always true. The following condition is sufficient to guarantee this: The rank of the Jacobian matrix of / has the maximal value m at each point on Z. We see that, under this condition, Z is an (n - m)-dimensional C°° manifold as follows. By the assumption, for any given point p on Z we can choose Xir, • • • , Xim such that the matrix '&M - %?V>\ is regular. We remove x^,• • • , x,m from the n variables x\, • • • , xn, and let the remaining variables be Xj1, • • • , Xjn_m. Then we see that there exists an open neighborhood U of p such that the coordinates Xj1,..., Xjn_m of an arbitrary point q on ZnU can move freely around the corresponding coordinate values of p, while the other m coordinates X{x, • • • , Xim are determined uniquely as C°° functions of them by virtue of the inverse function theorem 1.2 as follows. To show this, for simplicity, we assume that Xjx = Xi,--- , Xjn_m = xn-m. Consider Rn = Rn~m x Rm and define a map F : Rn -* Rn by F(x) = (xi,-- ,xn_m,/(x)). Then by the assumption, the Jacobian of F at p is not 0. Thus, by the inverse function theorem, there exists an open neighborhood U of p such that F induces a dif- feomorphism from U onto V = F(U). If we write the coordinates
1.2. DEFINITION AND EXAMPLES OF MANIFOLDS 19 of the point p as p = (pi,P2) (pi G Rn_m, p2 G Rm), then there exist open neighborhoods U\, U2 of pi and p2 in Rn-m and Rm respectively, and we may assume that U = U\ x(/2. Then the inverse function F-1 : V —» U is described as F-1(x) = (x1,---,xn_m,/i(x)). Here /i : V —> U2 is a C°° map. Now let q € Z n L7 be an arbitrary point on Z near p\ then F(^) = (91,0). Therefore, Q = (9i,92) = F~l o F(9) = (gi,%i,0)). From this we obtain q2 = h(qi,0), and surely we see that the first n — m coordinates (previously Xjlt- ¦ ¦ ,Xjn_m) of q are free to move around pi, while the remaining m coordinates (previously ?*,,••• , Xim) are determined as their functions. From the above discussion, we can use Xjx,• •• ,?Jn_m as the coordinates of a point on Z n U (see Figure 1.12). That is, Z f\U has the shape of a graph of a C°° map /i(<7i,0) defined on an open set U\ containing the point p\ in Rn-m with values in an open set Ui containing the point p2 in Rm. When p moves over Z, we have to change the local coordinates. However, it is easy to see that the transformations among the local coordinates are of class C°°. While we leave the details to the reader, we see in this way that Z is an (n — m)-dimensional C°° manifold. Figure 1.12. A general surface (d) Submanifolds. Example 1.14 (Open submanifold). An arbitrary open subset U of a C°° manifold M is a C°° manifold in a natural way, because if
20 1. MANIFOLDS S = {(UQ,<Pa)}a€A is a C°° atlas of M, then S' = {((/Qn[/,^)}a6/, is a C°° atlas of U. Here ip'Q denotes the restriction of ipatoUanU. We call U an open submanifold of M. At first sight, these manifolds may look worthless; however, actually they are very important. We will present two examples. The first example is the general linear group GL{n\ R), defined as the set of all regular n x n real matrices. The set M(n;R) of all real square matrices of order n can be identified with Rn in a natural way, and thus it is a C°° manifold. Since the determinant is obviously a continuous function on it, the set of all matrices whose determinants are not 0, that is, GL(n; R), forms an open submanifold of M(n;R). Moreover, GL(n;R) also has a Lie group structure that will be mentioned later. The second example is the complement of a knot. A knot is a closed curve in R3 that does not intersect itself (Figure 1.13). Figure 1.13. Knot Since the 3-dimensional sphere S3 is considered as the space obtained by adding one point (the point at infinity) to R3, the knot is also considered to be contained in S3. The space S3\K, where a knot K is removed from S3, is called the complement of K. It is well known that these spaces have rich structures and are very important in the theory of 3-dimensional manifolds. We now give the definition of a general submanifold. Definition 1.15. Let M be an n-dimensional C°° manifold. A subset N of M is called a submanifold of M if it satisfies the following condition. For an arbitrary point p € N, there exist an open neighborhood U of p and coordinate functions Xi, • • • ,xn defined on U such that N n U = {qe U; xk+1(q) = • • • = xn(q) = 0}. Here A: is a constant integer > 0. Furthermore, if N is a closed subset of M, N is called a closed submanifold.
1.2. DEFINITION AND EXAMPLES OF MANIFOLDS 21 In this case, it is easily seen that N has the structure of a k- dimensional C°° manifold in a natural way, and the inclusion map N C M is a C°° map. Example 1.13, above, shows that Z is a submanifold of Rn. In some books, the definition of submanifold requires weaker conditions. In that case, a subset satisfying the above condition is called a regular submanifold. (e) Projective spaces. Example 1.16. We write Pn for the set of all lines in Rn+1 through the origin and call it the n-dimensional real projective space. To distinguish Pn from the complex projective space which will be defined below, we sometimes write RPn. On Pn, a C°° manifold structure is induced as follows. Given a point (xi, • • • ,xn+i) on the space Rn+1 - {0} (that is Rn+l with the origin removed), the line passing through this point and the origin defines an element of Pn. This defines a projection 7T : Rn+1 - {0} - Pn, which is obviously a surjection. We put the quotient topology on Pn. That is, a subset U of Pn is defined to be an open set if tt~1{U) is an open set of Rn+1 - {0}. For two points x = (xi,--- ,in+i), y = (yi,--- ,yn+i) in Rn+1 -{0}, their images by it are the same if and only if there exists a nonzero number a € R such that y^ = ax{ (i = 1, • • • , n+1). In this case, if we denote this by x ~ y, we see that ~ gives an equivalence relation on Rn+1 - {0}. In other words, Pn is the quotient space of Rn+1 - {0} by this equivalence relation. We denote tc(x\,--- ,xn+i) by [xi,--- ,xn+i]. This is called the homogeneous coordinate of Pn. For i = 1, • • • ,n + 1, let [/i = {[x1,---,xn+1]€Pn; xt^0}. These are obviously open sets. Next we define pi : Ui —» Rn by VXi Xi Xj Xi / Then, after a brief consideration, we see that w is a homeomorphism. Moreover, if i ^ j, an explicit calculation shows that <pj o tp'1 : <Pi{Ui CiUj) -> <pj{Ui n Uj) is a C°° map. Thus, we see that P71 is a C°° manifold. Here we mention complex manifolds very briefly. We write C for the set of complex numbers. C is a field under the usual four basic
22 1. MANIFOLDS operations of arithmetic. Geometrically, we can identify C with R2 by the correspondence C9z = i + iy^ (x,y) € R2 (this is called the Gaussian plane). Therefore we can consider C as a 2-dimensional C°° manifold. The product Cn of n copies of C is a 2n-dimensional C°° manifold. However, actually Cn has a deeper structure, that of an n-dimensional complex manifold. In the definition of complex manifolds, a holomorphic mapping defined over an open set of Cn plays the role of a C°° map in C°° manifolds. Roughly speaking, a holomorphic mapping is a differentiable function with respect to each complex variable. If in the definition of a C°° manifold we replace Rn and C°° maps by Cn and holomorphic mappings, we obtain the definition of a complex manifold. Since holomorphic mappings are, of course, C°° maps, all n-dimensional complex manifolds are automatically 2n-dimensional C°° manifolds. However, it is known that holomorphic mappings have much finer properties than C°° maps, and this gives rise to deeper geometric structures on complex manifolds. On the other hand, compared with the holomorphic case, we can use various constructions for C°° maps freely, and the same is true for C°° manifolds. Example 1.17. In the definition of real projective spaces, if we replace R by C, we obtain the definition of complex projective spaces. That is, we denote the set of all complex lines through the origin in Cn+1 by CPn and call this the n-dimensional complex projective space. CPn is an n-dimensional complex manifold and therefore a 2n-dimensional C°° manifold. (f) Lie groups. Lie groups are by themselves a big research field. We briefly summarize aspects of Lie groups that we will need for this book. Definition 1.18. If a group G is also a C°° manifold, and both the multiplication G x G 3 (g>h) h-> gh e G of the group and the map G 3 g h-» g~l e G of taking the inverse are of class C°°, then G is called a Lie group. Furthermore, if G is a complex manifold and the above two maps are holomorphic mappings, then G is called a complex Lie group. Example 1.19. The set GL(n\C) of regular complex square matrices of order n is a complex Lie group, called the general linear group over C. Example 1.20. The set 0(n) of orthogonal matrices of order n is a Lie group, called the orthogonal group of order n. 0(n) is
1.3. TANGENT VECTORS AND TANGENT SPACES 23 defined by the equation lX X = E {lX is the transpose of X and E is the identity matrix) in the set M(n\ R) of all real square matrices of order n. Then we can see that the condition of Example 1.13 is satisfied (we put the case n = 2 as Exercise 1.2), and 0(n) is a C°° manifold. It is easy to verify the other conditions. The subgroup SO(n) of all elements of 0{n) with determinant 1 is called the special orthogonal group of order n. SO(n) is a connected component of the identity of 0{n), and the quotient group 0(n)/SO(n) is a finite group of order 2. 0(n) and SO(n) are compact Lie groups. Example 1.21. The set U(n) of all unitary matrices of order n is a Lie group, called the unitary group of order n. U{n) is defined by the equation X*X = E in the set M(n;C) of all complex square matrices of order n. Here, we denote by X* the matrix whose (i,j) entry is the complex conjugate of the (j,i) entry of X. Also in this case, the condition of Example 1.13 is satisfied and U(n) is a compact Lie group. 1.3. Tangent vectors and tangent spaces (a) C°° functions and C°° mappings on manifolds. C°° functions on manifolds give an important hint for studying the structure of C°° manifolds. Definition 1.22. Let M be a C°° manifold and / : M -» R a real valued function on M. If, for all locaL coordinate systems [U,ip) in an atlas that defines the C°° structure of M, / o <p~l : </?(?/) —> R is a C°° function on the open set <p{U) of Rn, then / is said to be a C°° function on M. In the above definition the phrase "all local coordinate systems in an atlas" may be replaced by "all local coordinate systems in the maximal atlas". This is because once the above condition is satisfied for one atlas, it is satisfied for all equivalent atlases. Next we consider the local description of C°° functions. Let A7, </>) be an arbitrary local coordinate system. Then a point p on U is identified with a point ip{xi(p), ¦ • ¦ yxn(p)) on Rn. Hence the function / o ip~l defined on <p(U) can be described as fo<p-l(x) = F(xu--- ,xn) {xe<p(U)) using the coordinate functions x\, • • • ,xn. This is the description of / on U by the C°° function F.
24 1 MANIFOLDS If U is an arbitrary open set of M, then U is naturally a C°° manifold (see Example 1.14). Hence we can consider C°° functions on U. For example, if A7, <p) is an arbitrary coordinate system of M, each coordinate function x* : U —* R is obviously a C°° function. We denote by C°°{M) the set of all C°° functions on M. For two functions / and g in C0C(M), the sum / + <?, the product fg and the multiplication af (a 6 R) by a real number are also elements in C°°(M). Equipped with these operations, Cco(M) has the structure of what is called an algebra over R. As the term algebra used here will appear frequently, we define it now. Definition 1.23. Let K be a field (in this book K may be thought of as R). If a vector space A over K has a ring structure by a product operation AxA3(A,/i)hA^A and the condition a(A/z) = (aA)/x = X(afi) (a € K, A,/i e A) is satisfied, then A is called an algebra over K. Example 1.24. Any real valued polynomial f(xi, ¦ • ¦ ,xn) of n variables defines a function on Rn. Hence all those polynomials form a subalgebra Pn of C°°(Rn). By restriction to 5n_1, these are also considered to be C°° functions on Sn-1. By extending the notion of C°° functions on manifolds, we obtain that of C°° maps between manifolds. Definition 1.25. Let M, N be C°° manifolds. A continuous mapping / : M —> N is said to be a C°° map, if the composition tp o f o ip~l is of class C°° wherever it has meaning (that is, on <p(?/n/-1(V))) for an arbitrary local coordinate system (U,<p) of M and an arbitrary local coordinate system (V}i>) of N. Although the above definition is natural, it is easy to see that we may rephrase it as follows. A necessary and sufficient condition for a map / : M —* N to be a C°° map is that for an arbitrary C°° function h on N, h o f is a C°° function on M. Theoretically, this definition may be said to be more straightforward. From the definition of a C°° map, it is easy to check that the composition go f of two C°° maps / : M —¦ N and g : N —* P is also a C°° map. Definition 1.26. Let M,N be C°° manifolds. A one to one C°° map / : M —» N from M onto N is called a C°° diffeomorphism or simply a diffeomorphism if the inverse mapping of / is also of class
1.3. TANGENT VECTORS AND TANGENT SPACES 25 C°°. If a diffeomorphism from M to N exists, M and N are said to be mutually diffeomorphic. The classification problem of C°° manifolds by diffeomorphisms is a fundamental problem in the field called differential topology. Example 1.27 (Hopf map). We define a map h : Sz -> S2 from the 3-dimensional sphere to the 2-dimensional sphere as follows. Consider S3 as the unit sphere S3 = {(zuz2)eC2; |*i|2 + M2 = l} in C2 and S2 as 1-dimensional complex projective space CP1 (see Example 1.17 and Exercise 1.3). Then we set . h(z1,z2) = [zuz2], {zi,z2)eS3. It is easy to see that this is a C°° map. This map, called the Hopf map after its discoverer H. Hopf, is a very important map with a rich structure. (b) Practical construction of C°° functions on a manifold. On a C°° manifold, there are a huge number of C°° functions. By handling those functions quite freely, we can also change them so that they will have various properties. Here we will introduce a well known and fundamental technique. We denote by D(r) the open disk {ar = (n,--- ,xn)€Rn; x\ + • • • + x2 < r2} of radius r > 0 with center at the origin in Rn, and by D(r) its closure. We choose a C°° function b : Rn —¦ R such that for all points x e Rn we have 0 < 6(x) < 1 and A.3) W-f1' X€^A)' When n = 1 the graph of b(x) is shown in Figure 1.14. Since its shape looks like a bump, we use the notation b(x). Such a function is constructed, for instance, as follows. We define a function h : R —¦ R (e-1/*, x>0, ' ~ \0, x < 0. h{x) = < It is easy to see that this function is of class C°°. Then, letting
1. MANIFOLDS ,f~ -2 -I ~\ 1 2 b(x) Figure 1.14. Graph of b{x) h{4 - x\ xj) h{4 - x\ x2n) + h{x\ + ... + x*-iy we can show that the required conditions are satisfied. Now, using the function b(x), we can construct various kinds of C°° functions on a manifold. Lemma 1.28. Let M be a C°° manifold. Let U be an open neighborhood of a point p of M, and f : U —> R an arbitrary C°° function defined on U. Then there exist an open neighborhood V ofp such that V C U and a C°° function f defined on the whole of M satisfying the condition '«> = {o(,)' qYv PROOF. We take a coordinate system (W, <p) around p satisfying W c U and <p(p) = 0, tp(W) D DC). Recall that DC) is an open disk of radius 3 with center at the origin of Rn. Such a local coordinate system is easily constructed, if necessary, by composing a homothety with the origin of Rn as the center. Let b = bcxpy using the function b of A.3). Then b is a C°° function on W and is 0 outside of <p~l(DB)) c W. Then, defining it to be 0 in the complement of Wy we can consider b as a C°° function defined on the whole of M. Now, if we let V = y>_1(D(l)), then V is an open neighborhood ofp, and obviously V C U and the value of b on V is 1. Hence, letting [0, q & W, we can see that / satisfies the required property. ¦ Applying this lemma to an arbitrary local coordinate system (U\x\, — - ,in) of My we can say the following. While each coordinate function ?» has meaning originally only on U, if we cut down
1.3. TANGENT VECTORS AND TANGENT SPACES 27 the domain a little it can be extended to a C°° function defined on the whole of M. However, of course, the extended functions no longer play the role of coordinate functions on the whole of M. (c) Partition of unity. To study the structure of manifolds, it is necessary to construct functions on them with various properties, vector fields and differential forms that will be defined later, and Riemanninan metrics, etc. On the other hand, C°° manifolds are constructed by glueing open sets in Rn by diffeomorphisms one by one. Accordingly, to construct the above items on manifolds, it is necessary to glue the constructed things on each coordinate neighborhood. In this process, it is the partition of unity that plays an important role. Let X be a topological space (not necessarily a manifold). A family of subsets {Ua} of X is called a covering of X if the union U Ua is the whole of X. If all the Ua are open sets, we call {Ua} an open covering. A covering {t/a} is said to be locally finite if, for each point on X, there exists an open neighborhood U such that the number of a with U n UQ ^ 0 is finite. A covering {Vp} is called a refinement of {Ua} if for an arbitrary C, there exists an a such that v0 C ua. Recall that a topological space X is said to be compact if an arbitrary open covering has a finite refinement. Relaxing this condition, if an arbitrary open covering of X has a locally finite refinement, X is said to be paracompact. The following proposition shows that all manifolds are not only paracompact but also have better properties. Proposition 1.29. Let M be a topological manifold. Then for an arbitrary open covering of M there exists a locally finite open covering {Vi,i = 1,2, •••} that is a refinement of it and has countably many elements with all V* being compact. In particular, M is paracompact. If necessary, we can make it satisfy the following stronger condition: each (V*,^) is a coordinate neighborhood, ipi(Vi) = DC), and {^t_1(^(l))} is already an open covering of M. Proof. Since M satisfies the second countability axiom, there exists a base {Oi\i = 1,2,-••} of open sets with countably many elements. Moreover, since M is a manifold, it is of course a locally compact Hausdorff space. From these facts, we see that the collection of all Oi such that Oi is compact is a base of the open sets. So, from the start, let us assume that the O, are all compact.
28 1. MANIFOLDS We construct a sequence of open sets E\, E2, • • • that grows with i as follows. In the first place, let E\ = 0\. By induction we assume that Ei, • • • , Ek are defined, and have the form Ek = OiU02U---UOifc. Then, since Ek is compact, for sufficiently large i, we have ?fccOiU02U---uOi. Among those i such that ik < i, let ik+i be the smallest and let Ek+i=OiU02U---UOik+l. Then obviously, for arbitrary k each Ek is compact and Ek C Ek+i. Moreover [)k Ek = M (Figure 1.15) (if M is compact, this operation terminates in a finite number of steps). Now let {?/a}a€A be an arbitrary open covering of M. Fix an arbitrary i > 1. For each point p € Ei — Ei-i, choose ap such that p € Uapt and take a local coordinate system (VJ,, 0P) around p that satisfies ipp{p) — 0, tpp(Vp) = DC) and Vp C ?/Qp n (Ei+i - E{-2) (we set E_i = E0 = 0). Let W = i>~l(D(l)). Since Ej - Ei-i is compact, we can choose a finite number of points p in it so that the corresponding open sets Wp cover Ei - -E^-i. We apply this operation to all i, and line up the local coordinate systems (Vp,ipp) corresponding to the finite number of points p chosen at each stage to have {(Vii^i)}t=:i,2."-- Then, {Vi} Figure 1.15 is a refinement of {Ua} and a locally finite open covering of M. Each Vi is compact, and obviously also satisfies the stronger conditions of the last part of the proposition. ¦ For a continuous function / : X —» R on a topological space X, the closure of the set of all the points where the value of / is not 0, suppf = {x€X;f(x) /0},
1.3. TANGENT VECTORS AND TANGENT SPACES 29 is called the support of /. Definition 1.30. Let M be a C°° manifold. A family {/* ;t = 1,2, • • • } of at most countably many C°° functions on M is called a partition of unity on M if it satisfies the following two conditions: (i) For each t, fi(p) >0(p€ M) and {supp fj} is locally finite. (ii) At all points p on M, J2i fiip) ~ 1- Furthermore, if {supp /»} is a refinement of an open covering {Ua}, the partition of unity {fi\i = 1,2, • • • } is said to be subordinate to the open covering {Ua}. THEOREM 1.31 (Existence of partitions of unity). Let M be a C°° manifold and {UQ} an open covering of M. Then there exists a partition of unity {/t ; i = 1,2, • • • } that is subordinate to {Ua}- Proof. Let {VJ be a locally finite refinement of {Ua} satisfying the stronger condition of Proposition 1.29. In particular, (V*,-^) is a local coordinate system such that ipi{Vi) — DC). Now, for each i, using the function 6 of A.3), let -{r «,)-<;'*«• \\l Then b, is a C°° function defined on the whole of M (see the proof of Lemma 1.28). Moreover, since supp 6» C K, we can set / = X> Then / is a C°° function defined on the whole of M. Furthermore, if we let Wi = Vt_1('C)(l))> as {Wi} is already an open covering of M, for an arbitrary point q € M there exists j such that bj(q) — 1. Consequently, the value of / is never 0. Then, if we let for each i, then {/i, /2, • • • } is a partition of unity subordinate to the open covering {Ua}- ¦ (d) Tangent vectors. We will review the case of En (see § 1.1(d)). A tangent vector at a point x on Rn is a vector with starting point x, and the set TxRn of all of them forms an n-dimensional vector space. We can take
30 1. MANIFOLDS -—, • • • , -— as its standard basis, and an arbitrary tangent vector dxi dxn v G TzRn can be described uniquely as their linear combination d d , = ai__ + ... + an_ Among the roles played by tangent vectors, the following two are fundamental, namely, the role of the velocity vector along a curve and the role of the directional derivative that gives the direction of the partial derivative of functions. The second role for the above tangent vector v gives a map t; : C°°(Rn) - R defined by v(/)=ai^(x)+",+anl{:(:r) (/€C,°°(Rn))- That is, v(f) is the partial derivative of the function / at x in the direction of v. It is easy to see that this map v has the following two properties: (i) v(f + g) = v(f) + v{g), v(af) = «;(/), (ii) v(fg)=v(f)g(x) + f(x)v(g). The first property simply says that v is linear. The important one is the second property, and it shows directly that v is a differentiation. We now introduce the notion of tangent vectors to a general C°° manifold M. In this case, a vector with the starting point p on M has no meaning in general. While, as in Example 1.13, if M is in Rn we can consider a vector with starting point p and tangents to M, a general manifold is defined intrinsically and is not always embedded in Rn beforehand. Then, we define tangent vectors on the basis of the above two properties. Definition 1.32. Let M be a C°° manifold and p a point of M. If a map v : C°°(M) —> R satisfies the conditions (i) v(f + g) = v(f) + v{g), v(af) = av(f), (ii) v(fg) = v(f)g(p) + f(p)v(g), for arbitrary functions f,g G C°°{M) and o € R, then v is said to be a tangent vector to M at p. We denote by TPM the set of all the tangent vectors at the point p of M, and call this the tangent space at the point p of M. If
1.3. TANGENT VECTORS AND TANGENT SPACES 31 we define the sum v + v' of two tangent vectors v,v' ?TPM and the multiplication by a real number av as (v + t/)(/)=v(/) + t/(/), M(/) = av{f), these are also tangent vectors and thus TPM is a vector space. For the function 1 G C°°(M) that is identically 1, v(l) = 0 for arbitrary v € TPM. The reason is that if we let / = g = 1 in condition (ii) of tangent vectors as differentials, we have v(l) = 2v(l). Furthermore, for the constant function a € C°°(M) (o G R), by the condition (i), we have v(a) = av(l) = 0. The operation of tangent vectors on C°°(M) is local in the following sense. If two functions /, g € C°°(M) coincide on an open neighborhood of p, we have v(f) = v(g) for arbitrary v e TPM. We can see this as follows. By the proof of Lemma 1.28, there exists a C°° function b that is identically 1 in a neighborhood of p and 0 outside of this open neighborhood. Then (/ - g)b = 0. Therefore, 0 = v((f - g)b) = v(f - g)b(p) + (/ - g)(p)v(b) = v(f - g). Thus v(f) = v{g). Now, let (U, 4>) be a coordinate system around the point p, and x\, ¦ • • ,xn its coordinate functions. Then it is easy to see that the correspondence C°°(MKf~d-^fl(v(p)) defines a tangent vector at p. We denote this tangent vector by eTpM. If the point p is clear beforehand, we sometimes dxi THEOREM 1.33. Let M be an n-dimensional C°° manifold. Then the tangent space TpM at an arbitrary point p on M is an n- dimensional vector space. Moreover, if (U;x\, ¦ • • }xn) is a local coordinate system around p, then the tangent vectors \dxxJp 'KdXnJp form a, basis ofTpM. PROOF. First of all, ( -—) , • • • , ( -—J are linearly indepen- \OX\/p \OXn/p dent, because if we apply these tangent vectors on each coordinate Uxjp
1. MANIFOLDS function Xj (extended to be a C°° function on the whole of M, see Lemma 1.28), obviously we have Next, we show that TPM is generated by the. above n tangent vectors. We identify the coordinate neighborhood U of p and <p{U) C Rn by ip, and assume that y(p) = 0 and that <p{U) is a convex set in Rn. For an arbitrary function / € C°°(M), the restriction of / to U is, under the above identification, described as a C°° function F(x) defined on <p{U). Explicitly, F = fof~l. Now since, for an arbitrary x e <p(U), fl dF if we let <?i(x) = / -z—{tx)dt, this is a C°° function and we have Jo OXi n A.4) F(x) = F@) + ^xi9t(x). i=l Obviously, pj(O) = -—@). We will apply an arbitrary tangent vector OXi v e TPM to A.4). Paying attention to the conditions of a tangent vector, we have n v{f) = v{F)=Y,v(xi)gi@) n <9 Since / was arbitrary, it follows that v = Y^ v(xi)-—. The proof is . OX4 complete. ¦ Next, we mention another role of the tangent vector as the velocity vector of curves. A C°° map from an open interval to a C°° manifold is called a C°° curve on M. Let c : (a, 6) —¦ M be a C°° curve passing through the point p = c{to) {to € {a,b)) of M. Then
1.3. TANGENT VECTORS AND TANGENT SPACES 33 the velocity vector c(to) of c at the point p is defined as follows. For / € C°°(M), we put A.5) c((„)(/) = ^| . at \t=to It is easy to check that this satisfies the conditions of a tangent vector, and so we have c(io)(/) € TPM. We sometimes write -r-(io) or — at at 110 instead of c(to)- If we consider the velocity vector of various curves, we see that all the elements in TPM appear. In the case of Rn and general surfaces (Example 1.13) in Rn, the tangent space at a point on it has an explicit geometric meaning. In the case of a general manifold, since the definition of the tangent space is abstract, its image might be difficult to grasp. However, we may understand it as a first-order approximation of neighborhoods of each point on the manifold by a vector space of the same dimension. Finally, we will see how the expression of the tangent vector changes under a coordinate change. Proposition 1.34. Let (U;xir- ¦ ,xn) and (V;yi,--- ,yn) be two local coordinate systems around a point p on a C°° manifold M. Then, dxi ^dx^'dyj' As the proof is easy, we leave it to the reader. (e) The differential of maps. Let M and N be C°° manifolds and / : M — N a C°° map. Then for each point p of M, a linear map /,:TpM->T/(p)/V is defined as follows. This is called the differential of / at p. We sometimes write dfp for /¦. Let v e TPM be a. tangent vector to M at the point p; then the correspondence C°°(iV)9/iHt,(fto/) eR is a linear map; and moreover it is easy to see that it satisfies the property of differential at the point f(p) on TV. We denote this by f*{v) e Tf(p)N'. It is easy to see that /„ is a linear map. As for the differential of the composition g o / : M —> P of two C°° maps / : M —> N and g : N —* P, it is easy to see that
34 1. MANIFOLDS (9 ° /)¦ = 9* ° /¦ • TPM -+ Tgof{p)P. That is, the differential of the composition is the composition of the differentials of each map. We describe the differential of a map explicitly by local coordinates. Let / : M —> TV be a C°° map and let f(p) = q (p € M). Let (U\x\, ¦ • ¦ ,xm) be a local coordinate system around p and {V'i yi» • • • ,Ifn) a local coordinate system around q. Then it is easy to check that A-6) f(±\=Y?VL±. K } J'\dxJ ^dXi dVj Using A.6) and the formula for the differential of a composition, we obtain the following proposition. We leave the detailed proof to the reader as Exercise 1.8. Proposition 1.35. Let M and N be C°° manifolds and f : M -* N a C°° map. Then, for an arbitrary tangent vector v € TPM at the point p on M and an arbitrary function h 6 C°°(N) on N, v(W) = /.(tOfc. (f) Immersions and embeddings. For a C°° map / : M -* TV, its differential /¦ : TPM — T/(p)/V reflects the geometric properties of /. We give the following definition. Definition 1.36. Let / : M — N be a C°° map. (i) If at each point p € M the differential /» : TPM —> Tf(p)N of / is an injection, we call / an immersion. (ii) If / : M —> N is an immersion and also / is a homeomorphism from M onto the image f{M) of /, we call / an embedding. Note that we give f{M) the relative topology as a subset of N. (iii) If / : M -* N is a surjection and at each point p the differential /» : TPM —» Tf(p)N of / is a surjection, we call / a submersion. The difference between immersions and embeddings concerns whether M intersects itself in its image f{M) or whether it approaches itself repeatedly or not, and is a global matter. We show their conceptual figure in Figure 1.16. A typical example of a submersion is the Hopf map (Example 1.27). Embedding is closely related to the concept of submanifolds (Definition 1.15). Actually the following theorem holds.
1.3. TANGENT VECTORS AND TANGENT SPACES Figure 1.16. Immersion and Embedding THEOREM 1.37. Let f : M -* N be an embedding from a C°° manifold M to N. Then, f{M) is a submanifold of N and f gives a diffeomorphism from M onto f{M). Conversely, if M is a submanifold of N, the inclusion map i : M —» N is an embedding. PROOF. Let the dimensions of M and Nbem and n respectively. By assumption, for an arbitrary point q € f(M) on /(M), there exists a unique point p G M such that f(p) = q. If we choose a local coordinate system (V;y,\, ¦ ¦ ¦ ,yn) around q such that V is small enough, then U = f~l{V) is a coordinate neighborhood containing p. Let xi, • • • , xm be the coordinate functions defined on U. For the sake of simplicity, we will identify U and V with open sets in Rm and Rn respectively, and also let p and f(p) be the origins respectively. Then the restriction of / to U is described by n C°° functions fi as follows: Vi = /t(*i,--- ,xm) (t = l,--- ,n). Since / is an embedding, the rank of the Jacobi matrix of / at the point p is m. If m = n, by the inverse function theorem 1.2, / is a diffeomorphism in a neighborhood of p. Thus, the claim of the theorem follows easily. Let m < n. Then, if necessary changing the order of the variables y\, • • • , yn, we may assume that "(if H,=,.. 7^0. Now, if we define a function F = (F1, • • • , F„) : U x Rn~n (x,tu) = (xi,--- ,xm,u;i,--- ,wn-m) e U x Rn-m, by » Rn, for Fi{x,w) /t(x) + lUi_ (t = 1,- • • ,m), (i = m + 1, • • • , we see that the Jacobian of F does not vanish at the origin. Again by the inverse function theorem, F is a diffeomorphism in a neighborhood of the origin. Therefore, for a sufficiently small neighborhood V CV of q = /(p), if we consider F : V -» {7 x R"-m c Rn, (V^F)
36 1. MANIFOLDS is a local coordinate system around q. Then, while the coordinate functions on V are xi,--- ,xm,wi,--- ,tun_m, by the definition of F, obviously f{M) r\V = {q* e V']Wi{q') = • • • = wn.m{q') = 0}. This shows that f{M) is a submanifold of TV. Since the proof of the remaining claims of the theorem is easy, we will leave it to the reader. ¦ Next we prove that an arbitrary compact C°° manifold can be embedded into RN for a sufficiently large N. By this (although there is the assumption of being compact), it is shown that an abstractly defined arbitrary manifold is realized as a submanifold of RN. THEOREM 1.38. An arbitrary compact C°° manifold can be embedded into RN for a sufficiently large N Proof. Let M be a compact n-dimensional manifold. Let {Ui,(pi}i=i>...,r be an atlas of M with a finite number (r) of coordinate neighborhood systems satisfying the following conditions. The image <Pi{Ui) of each <pi : U —> Rn is an open disk DB) of radius 2 with center at the origin, and if we put V* = (p^1{D{l)I {Vi} is already a covering of M. The existence of such an atlas follows immediately from the compactness of M (see the proof of Proposition 1.29). Then, we construct a C°° map fi:M-*SnC Rn+1 from M to the n- dimensional sphere 5n to satisfy the following two conditions: (i) the restriction of / to V, is a diffeomorphism from Vi onto the southern hemisphere {x € 5n; xn+\ < 0} of 5n, and (ii) / maps the complement of Vi to the northern hemisphere. Intuitively, the existence of this map seems obvious. As the proof is not so difficult, we wish the reader to try it. Now, if we define a map / : M —» Rr(n+1) as /(p) = (/l(p),..-,/r(p)) (P€M), it is easy to check that / is an embedding. ¦ In fact, more generally, it is known that an arbitrary n-dimensional C°° manifold can be embedded in R2n+1 as a closed submanifold (Whitney's embedding theorem). 1.4. Vector fields (a) Vector fields. In §l.l(d), we defined vector fields on Rn. In this section, we consider vector fields over general manifolds.
1.4. VECTOR FIELDS 37 Let M be a C°° manifold. A vector field X on M is an assignment, to each point p € M, of a tangent vector Xp eTpM in such a way that Xp is of class C°° with respect to p. If we let A7; X\, • • • , xn) be a local coordinate system of M, for each point p € U, X is described as follows: A-7) *P = X>(P)^-. where a* is a function defined on U. This is called a local expression of X. Xp is said to be of class C°° with respect to p, if each coefficient a,i is a C°° function. Let yi, • • • , yn be another system of coordinate functions defined around p. Then by Proposition 1.34, we have *'-g(?*w?w)?- Therefore, the condition of each coefficient di being a C°° function does not depend on the choice of local coordinates. We shall denote the set of all vector fields on M by X{M). When two vector fields X, Y e X(M) are given, by putting (X + Y)p = Xp+Yp we can define their sum X+Y € X(M). Also, for an arbitrary real number a € R, by putting {aX)p = a(Xp) we can define the multiplication aX € X(M) of X by a. As is easily seen, with these two operations X(M) becomes a vector space over R. Moreover, when / € C°°{M) is an arbitrary C°° function on M, fX e X{M) is defined by {fX)p = f(p)Xp. Therefore, X{M) has a structure of module not only over R but also over C°°(M). Briefly, as for vector fields, we can do operations such as multiplication by function, addition and subtraction freely. Recall that one important role of the tangent vector is the directional differentiation of functions. Using this role, we can make the vector field X act on any function / € C°°(M). That is, if we put (Xf)(p) = xp(f) (j>eM), we obtain a function XfonM. If we give a local expression of X as A.7), then (Xf)(p)=J2ai(p)^-(p),
38 1. MANIFOLDS so that Xf is also a C°° function on M. Xf is called the derivative of / by the vector field X. In this way, we obtain a map X{M) x C°°{M) 3 (XJ) ^Xf? C°°(M). While this map is obviously linear with respect to X, as for / it is easy to see that it satisfies the following two conditions: (i) X{af + bg) = aXf + bXg (a, 6 G R, f,g € C°°{M))\ (ii) X(fg) = (Xf)g + f(Xg). Generally, we call a map C°°(M) —* C°°(M) satisfying these two properties a derivation of the algebra C°°(M) over K. The derivation of functions by vector field characterizes it completely. That is, the following proposition holds. Proposition 1.39. Let M be a C°° manifold and X,Y vector fields on M. If Xf = Yf for an arbitrary C°° function f on M, then X = Y. Proof. It is enough to show that Xp = Yp at an arbitrary point p on M. We choose a local coordinate system (U\x\, ...,i„) around p. Let *-§«•?• Y=pi be the local expressions of X, Y with respect to this local coordinate system. By Lemma 1.28, there exists a C°° function X{ defined on M and coinciding with each coordinate function X{ in a neighborhood of p. Then, since o^ip) = XpXi = YpXi — 6i(p), we obtain Xp = Yp, and the proof is finished. ¦ (b) The bracket of vector fields. Let X,Y € X{M) be two vector fields on a C°° manifold M. Then, both XtY act on C°°(M) as derivations. Consider a map A.8) C°°{M) 3 f -> X{Yf) - Y{Xf) € C°°(M). By an easy calculation, we can check that this map also has the two properties of the derivation. If we rewrite it as X(Yf) - Y{Xf) = (XY - YX)f, it indicates that XY - YX expresses a vector field on M. Actually, using the symbol [X, Y] instead of XY - YX, we consider the correspondence A.9) C°°(M) 3 f - [XtY]pf = Xp(Yf) - Yp(Xf) € R at each point p 6 M. From the fact that A.8) satisfies the properties of a derivation, we immediately see that the correspondence A.9)
1.4. VECTOR FIELDS 39 satisfies the condition (see Definition 1.32) of tangent vectors at the point p. That is, we can consider [X, Y]p as a tangent vector to M at p. If it is shown that [X, Y]p is of class C°° with respect to />, we can conclude that [X, Y) is a vector field on M. In order to check it, we give X, Y the local expressions x = Ta°, y = y>#-. f^ dxt ^ dxi By an easy calculation, we have tj=i J Prom this, we see that [X, Y] is a vector field on M, and simultaneously its local expression is given by The vector field [X, Y] defined in this way is called the bracket of X and Y. We will cite some properties of the bracket. PROPOSITION 1.40. The bracket of vector fields has the following properties. (i) [aX + bX\ Y] = a[X, Y) + b[X', Y) (a, b e R), and the same forY. (n) [Y,x] = -(x,n (iii) (Jacobi identity) ([A",y],Z] + [[K,Z],X] + [[Z,X],y] = 0. (iv) lfX,gY] = fg[X,Y] + f(Xg)Y-g{Yf)X (f,g e C°°(M)). If we calculate carefully, the proof is not so difficult. We leave it as Exercise 1.7. We call a vector space with operation [, ] satisfying the above conditions (i)-(iii) a Lie algebra. Accordingly, X(M), the set of all vector fields on M, becomes a Lie algebra over R with respect to the bracket operation defined above. In fact, we can see that X(M) is naturally isomorphic to the set of all derivations (in which the Lie algebra structure is also introduced) of C°°{M) as Lie algebras. The interested reader may consider the proof. (c) Integral curves of vector fields and one-parameter group of local transformations. Let M be a C°° manifold and X a vector field on M. A curve c : (a, 6) —» M in M is called the integral curve of X if the velocity
40 1. MANIFOLDS vector c{t) e TC(()M (t € (a, 6)) at each point coincides with the value of X at that point. Imagine that X expresses the velocity vector of a smooth flow of water over M. Then if we float a small grain at a point p on M, it drives down the flow and draws a curve on M. This kind of curve is the integral curve. If we change the point to float the grain, its curve will be different in general. However, no matter how we choose those two curves, they never cross each other (Figure 1.17). Figure 1.17. Integral curves We shall find equations that should be satisfied by the integral curve through an arbitrary point p on M. Let (U;xi,-- ¦ ,xn) be a local coordinate system around p. Then X is locally expressed as on U. Let c : (a, 6) -¦ M be the required integral curve, and for simplicity, choose a parameter such that c@) = p. If we describe the position of c(t) by the local coordinate system as c(t) = (xi{t), • • • , x„(i)), we have Hence the required equation is dx A.11) -~@ = a<(*i@.-",*n@) (; = !,•••,n). The condition of passing through the point p at t = 0 is expressed as the initial condition; Xi@) = Xi(p). Note that if we choose another point for p, only the initial condition changes; the equation itself is the same. By the way, A.11) is a system of ordinary differential equations of first order, and, as is well known, for this type of differential equations the following theorem of existence and uniqueness of the solution holds.
1.4. VECTOR FIELDS 41 Theorem 1.41 (Existence and uniqueness of the solution of ODEs). For the ordinary differential equations A.11), the following hold. (i) (Existence) For an arbitrary initial condition Xi@) = x»(p) (p € U), if we choose € > 0 small enough, A.11) has a solution of class C°° defined on — c < t < e. (ii) (Uniqueness) // two solutions of A.11) have the same value at a point t = to on their domains, they coincide all over the intersection of their domains. (iii) (Differentiability of the solution with respect to the initial condition) For an arbitrary point p € U, if we choose a neighborhood V (C U) of p small enough, there exists an e > 0 such that for all q € V a solution satisfying the initial condition Xi(to) = %i(q) is defined on to - e < t < to + e. Furthermore, if we consider the family of these solutions as a function of t and X\, • • • ,x„, it is of class C°°. By Theorem 1.41 (i), we see that there exists an integral curve through an arbitrary point p when t = 0 anyway. We shall consider extending the (domain of the) integral curve as long as possible. Since, by Theorem 1.41 (ii), if two integral curves pass through the same point at the same time, then they are connected as a single integral curve; the way of extension is unique. If it is going out of the coordinate neighborhood U, we can continue the discussion by choosing a new coordinate neighborhood. The reason is that, even if we change the coordinate neighborhood, the form of the equation does not change. In this way, for each point p e M there exists an integral curve c(p) that passes through the point when t = 0, and it cannot be extended any more. We call this integral curve the maximal integral curve through p. By Theorem 1.41 (i), c(p) is a curve of class C°°. Let (ap, bp) be the domain of c(p). Of course, — oo < ap < 0 and 0 < bp < +oo. For an arbitrary point q on c(p), c(q) is different from c(p) only by a translation of the parameter, and is the same curve. That is, if it starts from p and arrives at q at the time s, we have A.12) c(q)(t)=c{p)(t + s) (<? = c(p)(s)). We can see this fact, if we fix s and calculate the velocity vector of c(p)(* + s) and apply Theorem 1.41(h) (see Figure 1.18). In this way, M is covered completely by all the maximal integral curves that are pairwise disjoint. Though it is still called an integral curve, the integral curve through a point where the value of the vector
42 1. MANIFOLDS 0 = c(p)(s) Figure 1.18 field is the zero vector (such a point is called a singular point of the vector field) consists of a single point. Let W = {(t,p) e R x M\av < t < bp} and define a map $ : W — M by $(t,p) = c(p){t) {{typ) e W). That is, $(?,p) is the point at which we arrive, at the time t, after starting from the point p and going through the integral curve. By Theorem 1.14(i), (iii), we see that W is an open set of R x M containing 0 x M, and the map $ is of class C°°. For an arbitrary i, let Mt = {p e M\ ap < t < bp}. Then Mt is an open set (empty in some cases) of M. We define the C°° map tpt '¦ Mt —¦ M by <Pt(p) = ${t,p) (p G M). Of course, Mo = M and <^0 = idw- PROPOSITION 1.42. For an arbitrary t, <pt '• Mt —* M is a diffeo- morphism onto an open submanifold of M. Furthermore for arbitrary t, s E 1R and p G M, the equation A.13) <Pt°<Ps(p) = <Pt+s(p) holds as long as both sides are defined (that is, if p e Ms, <ps{v) € Mt, p € Mt+a). PROOF. We prove the latter part. By the definition ift+s(p) = c(p)(t 4- s), and on the other hand, if we write y?5(p) = c(p)(s) = q, the left hand side of A.13) is c{q){t). Then by A.12), the equation A.13) holds. The former part can be proved as follows. First, by the definition of Mt, we see that M_t = y?t(Mt). Next, by the above facts we have (pto<p-t = <P-t ° <Pt = id, so that </?t is a diffeomorphism from Mt onto M-t and ip-t serves as its inverse. ¦ Definition 1.43. Let X be a vector field on a C°° manifold M. The set of all diffeomorphisms {y?(; t € R} constructed above is called the one parameter group of local transformations generated by X. In the above notation, the reason why we say "group of local transformations" is that yt is not always defined on the whole of M.
14. VECTOR FIELDS 43 Definition 1.44. A vector field X on a C°° manifold M is called a complete vector field if an arbitrary element <pt of the associated one parameter group of local transformations {</?*} is defined on the whole of M (that is, if Mt — M or equivalently W = R x M). In other words, X is complete if the integral curve through an arbitrary point on M is defined on the whole of R. Now we shall denote the group of all diffeomorphisms of M by DiffM. We define the multiplication by the composition of diffeomorphisms. This group is an infinite-dimensional group called the diffeomorphism group of M, and it is a very important group whose global structure remains mostly unknown. If a vector field X on M is complete, by Proposition 1.42, the correspondence R 3 t >-> <pt G DiffM is a group homomorphism. That is, the vector field X generates a one-parameter commutative subgroup in DiffM. In this case, we call {</><; t e R} a one parameter group of transformations generated by X. We sometimes write Exp tX for it. The following theorem gives a sufficient condition for X to be complete. THEOREM 1.45. Let M be a compact C°° manifold. Then an arbitrary vector field X on M is complete. Proof. Consider W = {{t,p) e R x M;ap < t < bp}. As we have seen above, W is an open set containing 0 x M. Since M is compact by the assumption, we can choose c > 0 small enough so that (-e, e) x M C W. That is, the integral curve c{p) going through p when t = 0 is defined for -e < t < e. In this case, we can see that the domain of c(p) can be actually extended to - §e < i < |e for any p. This is because if we start from p and go along c(p) till time |t| = ^e (let q be that point), then we can keep going along c(q) (= c(p)) for \t\ < e. But then, by the discussion in A.12), we can conclude that we started from p and went along c(p) till the time |f| < |e. Repeating this discussion, we can go along c(p) for an arbitrary time, and the theorem is proved. ¦ Example 1.46. As a simple example of a noncomplete vector field, we have M = R2 - {0}, X = —. In this case, the integral ox curve through the point A,0) € M, for example, is defined only for -1 < t < oo.
44 1. MANIFOLDS (d) Transformations of vector fields by difFeomorphism. Let M,N be C°° manifolds and /:M-»iVa difFeomorphism from M to N. Then, for an arbitrary vector field X on M, a vector field f+X on N is defined by {f.X)q = f.{Xf-Hq)) (q<EN). Or equivalently we can write f*(Xp) = (f*X)f(p) (p € M). Then, for an arbitrary function h e C°°(N) on N, we have A.14) (f.X)h = X(hof)of-\ This holds because, for a point q € N, we have ((/.x)/i)(<?) = {f.x)qh = /•(x/-i(„)/i, and on the other hand, by Proposition 1.35, we have f.{Xf-Hq))h = Xf-iM(h o /) = (X(h o /) o /-%). 1.5. Fundamental facts concerning manifolds (a) Manifolds with boundary. Up to now, we have defined a manifold as a figure that is locally homeomorphic to an open set in JRn. Therefore, whatever point we choose on a manifold, its neighborhood looks the same. However, when we construct a surface B-dimensional manifold) by glueing sheets of paper, for instance, it is not always a manifold at each step in the construction. To begin with, a slip of paper has corners and boundaries, and disks and cylinders also have boundaries. Similarly, in the study of manifolds, manifolds with corners and boundaries often play important roles. We now introduce manifolds with boundaries. First, we define "the upper half space" Hn of Rn, by Hn = {x = (xi,--- ,in) eRn; xn > 0}. Its subset dUn = {x € Hn; xn = 0} is called the boundary of Hn. It can naturally be identified with Rn_1. Here Hn\dHn is called the interior of Hn. The case n = 2 is illustrated in Figure 1.19. A continuous map </? : U —> V between two open sets U and V in Mn is called a C°° map from U to V if it can be extended to a C°° map from an open set U' of Rn containing U to an open set V of Rn containing V. We call ip : U —> V a diffeomorphism if it is a homeomorphism and both y? and ip~l are C°° maps. In this case, it
1.5. FUNDAMENTAL FACTS CONCERNING MANIFOLDS 45 <?H2 Figure 1.19. The upper half plane H2 is easy to see that the restriction of ip to U n dMn is a diffeomorphism from UndWn to VnMn. Now, everywhere in the definition of C°° manifolds (Definitions 1.4 and 1.7), let us replace Rn by Hn. That is, for a HausdorfF space M satisfying the second countability axiom, let an open covering {Ua}a€A of M and a homeomorphism tpa : Ua —» Va from UQ onto an open set Va of Hn be given, and for arbitrary a, 0, assume that h<x = V/3 ° V^1 : VaiUcc H U0) -»<p0(Ua C\ Up) is a C°° map. Let S = {{Ua,<Pa)}aeA- We denote by dM the set of all the points p in Ua that are mapped by (pa to dMn, where a moves arbitrarily. If dM ^ 0, we say that S defines a structure of a C°° manifold with boundary on M. In this case, we call dM the boundary of M. We see that dM has an (n — l)-dimensional C°° manifold structure in a natural way. Local coordinate systems, atlases, maximal atlases, etc. are defined in the same way as in the case without boundary. Disks, cylinders and the Mobius strip (Fig. 1.20), which will appear in the next subsection, etc. are all 2-dimensional manifolds with boundary. Figure 1.20. Mobius strip and cylinder Example 1.47. The n-dimensional disk Dn = {x e Rn; x\ + H x2 < 1} is a manifold with boundary, and dDn = 5n_1.
46 1. MANIFOLDS A compact C°° manifold without boundary is called a closed manifold. Closed manifolds are the most important objects in the study of manifolds. (b) Orientation of a manifold. We now compare a Mobius strip, which is a famous surface where we cannot distinguish the two faces, with a cylinder, where we can distinguish them. How can we formulate the difference in the properties of these two surfaces mathematically? Since both are 2-dimensional manifolds, we can construct them by glueing some small pieces that are homeomorphic to open sets of K2 one by one. Specifying the two faces of the pieces beforehand, we consider the construction of surfaces using only identifications that preserve the right side and the reverse side respectively. Though cylinders, spheres, tori, etc. can be constructed in this way, the Mobius strip cannot. However it is only for convenience that we mention the two faces. For example, when a cylinder is constructed from several pieces, even if there is one that is inside out, we need only reverse it to settle the case. The essential point is whether, when a surface is constructed by glueing some pieces, we can distinguish the two faces of each piece so that they can be well connected as a whole. A surface on which this can be done is called an orientable surface. Non-orientable surfaces include, besides the Mobius strip, the projective plane RP2 and the Klein bottle. While here we used the word orientation, we shall present it somewhat differently from the distinction of two faces. We can define orientations at each point on a surface, and there are exactly two. We shall present them by two kinds of arrows, clockwise and counterclockwise. We call them opposite orientations. When an orientation is specified at a point, the "same" orientation is specified at an arbitrary point in a neighborhood of the point. This is called the coherent orientation (Figure 1.21). We specify an orientation at a point on a surface, and choose the coherent orientation at each point on a curve starting from the point. If the curve goes back to the starting point, the original orientation at that point may or may not coincide with the orientation propagated along the curve. Now a surface is orientable if the orientation propagated along any curve always comes back to the starting orientation. In this case we can assign an orientation to all points on the surface in such a way that near points have mutually coherent orientations.
1.5. FUNDAMENTAL FACTS CONCERNING MANIFOLDS Figure 1.21. Orientation on a surface We shall generalize the above facts to an n-dimensional manifold Mn. In the same way as in the case of a surface, we wish to define exactly two orientations at each point p € M. While for general n we cannot use an arrow, its role is played by an ordered basis of the tangent space TPM at p. Let tii, • • • ,un and t>i, • • • ,vn be two ordered bases of TPM. Then they can be transformed into each other by a regular linear transformation T of TVM. We say that two ordered bases are equivalent if the determinant of T is positive. It is easy to see that this is an equivalence relation. Then we call an equivalence class of ordered bases of TVM an orientation at p. As this definition is somewhat abstract, we shall examine the case of R3. If we use the ordinary coordinates x,yyz, we can take 7T-1 7T-. tt as a basis of the tangent space TVR3 at each point p e ox ay oz K3. If we consider this as an ordered basis, we call the resulting orientation the right-hand system. The reason is this: with respect to the usual coordinate axes, if we point the thumb and the index finger of the right hand in the positive directions of the x axis and the y axis respectively, then the middle finger is pointing in the positive direction of the z axis. Against this, for example, both —, —, — oz ay ox and - —, —, — determine another orientation (called the left-hand ox oy oz system). If we observe the reflection of ourselves in a mirror, we see that the right hand and the left hand are reflected to the opposite hands. Mathematically speaking, this means that an arbitrary planar symmetry is a transformation that reverses orientation.
I. MANIFOLDS If an orientation is specified at a point p on M, the coherent orientation is determined at an arbitrary point near p. Therefore it is natural to define the orientation as follows. Definition 1.48. If we can assign an orientation to each point on a manifold M in such a way that the orientations at any two sufficiently near points on M are coherent, we say that M is orientable. If such an orientation is specified, we call it the orientation of M and call M an oriented manifold. We see that there are exactly two orientations on a connected orientable manifold. We call them opposite orientations. We sometimes denote by —M the manifold with the opposite orientation for an oriented manifold M. In the case of a differentiable manifold, there is an easy method to judge the orientability by the glueing maps. We give Rn the natural orientation obtained from the usual coordinates - that is, the orientation obtained by the ordered basis ——, • • • , ——. We shall oxi oxn call this the positive orientation. Let U, V be open sets of Rn, and let <p : U -+ V be a difTeomorphism. How does <p transform the orientation? By the differential of <p <p. : TpRn - r„(p)Rn at a point p e U, the above canonical basis in TpRn is mapped to the basis d d <p'dxl'"','Pmdxn of T^(p)Rn. The matrix which transforms the canonical basis of Ty(P)Rn to the above basis is just the Jacobian matrix of <p at p. Therefore, the positive orientation at p is transformed by </?* to the positive or negative orientation at y?(p) according to the sign of the Jacobian of <p at p. From this consideration and Definition 1.48, it is easy to see that the following proposition holds. PROPOSITION 1.49. A C°° manifold M is orientable if and only if there exists an atlas <S = {{Ua,<pa)}a€A of M such that the Jacobians of all coordinate changes fpa = tpp o ip~ * are positive at all the points on<pa(Uaf)Up). Usually, on a C°° manifold, the orientation is given by an atlas S satisfying the above condition. In that case, we call a local coordinate
1.5. FUNDAMENTAL FACTS CONCERNING MANIFOLDS 49 system (U, <p) of this manifold a positive local coordinate system if it belongs to «S or if adding it to S preserves the same condition. Let M, N be connected oriented C°° manifolds and / : M —* N a diffeomorphism. The orientation given at a point p of M is mapped by the differential of / to a certain orientation at a point f(p) of N that may coincide with the assigned orientation of N or may not. Since M is connected, the situation is determined independently of the choice of p. If the former holds, we call / an orientation preserving diffeomorphism. Next we shall consider an n-dimensional C°° manifold with boundary. The boundary dM of M is an (n- 1)-dimensional C°° manifold. The orientability of M is defined in almost the same way as in the cases without boundary. Let M be orientable and specify an orientation. Then an orientation is induced to dM in the following way. Let p € dM and let xi,¦ ¦ • ,xn be the positive local coordinates defined in a neighborhood of p. For any point on dM, xn = 0. Then, xi, • ¦ • ,xn-i serve as the local coordinates of dM in a neighborhood of p. We define the induced orientation of dM (at p) to be the one determined by the ordered basis dx\ <9x2' ' dxn-i oiTp(dM). By this definition it is relatively easy to see that an orientation is determined on dM independently of the choice of the point p. We encourage our readers to try it by themselves (Exercise 1.9). Although the sign (-l)n may look artificial, it is for compatibility with the definition (see §3.1) of the boundary of an oriented simplex, and it also has the advantage of putting Stokes' theorem (Theorem 3.6) in an elegant form. More geometrically, an equivalent definition is obtained in the following way. Let v be the "outward" normal vector at the point p 6 dM. With respect to the above local coordinates Xi, • • • ,xn, we have v = --—. Then we say that an ordered dxn basis vi, ¦ • ¦ ,fn-i of Tp(dM) is the induced orientation of dM when V)^i? • • • ,vn-\ coincides with the orientation of M at p. (c) Group actions. On a set or figure X, let a kind of structure be given. For example, let X be a C°° manifold, and as the structure we can consider the Riemannian metric, which will be defined later in §4.1, or a complex structure, etc. Then if G denotes the set of all the transformations
50 1. MANIFOLDS of X that preserve the structure, generally G becomes a group. This will be called the automorphism group of X with respect to this structure. Conversely, let a group G act on X from the left in the following sense. That is, a map f:GxX->X is given, and if we denote f{g,x) (p,/i € G) simply by gx, the following two conditions are satisfied: (i) ex = x (e is the identity of G), (ii) {gh)x = g{hx) {g,heG). Then for an arbitrary element g € G, if we define f9 : X —» X by fg{x) = gx, we see that this is a one to one correspondence from X onto X. Here, the following question naturally arises: what is the structure of X that is invariant under this action of G? For x € X, if we put Gx = {g € G\gx = x}, Gx is a subgroup of G. This is called the stabilizer at x. The action such that, for all x, Gx consists of only the identity element is called a free action. Also we call Gx = {gx\ g e G} the orbit through x. If we define two points x,y on X to be equivalent provided they lie on the same orbit, X is endowed with an equivalence relation. We write X/G for the quotient space of X by this equivalence relation, and call it the orbit space or quotient space of this action. Although we described the left action above, the right action X x G —¦ X is essentially the same. In the case where X is a C°° manifold M, we consider differen- tiable group actions. The largest one among all such groups is the group DifTM of all diffeomorphisms of M, which we already defined in § 1.4(c). Corresponding to diverse geometric structures on M, various subgroups of DiffM appear. A specially important one is the differentiate action of a Lie group G on M, that is, an action such that /:GxM—»M is a C°° map. In this case, we usually give the orbit space the quotient topology. As a special case of Lie groups, there is the action of a discrete group. Definition 1.50. Let a group r act differentiablly on a G°° manifold M. If for an arbitrary compact set K C M there are only a finite number of 7 € T such that 7/f D K ^ 0, we say that this action is properly discontinuous. As a trivial example of a properly discontinuous action, there is the action of a finite group.
1.5. FUNDAMENTAL FACTS CONCERNING MANIFOLDS 51 (d) Fundamental groups and covering manifolds. Let M be a C°° manifold. One method of constructing a new manifold from M is to take covering manifolds. Definition 1.51. Let M, N be connected C°° manifolds, and 7r : N —» M a C°° map. If, for an arbitrary point p e M, there exists an open neighborhood U of p such that each connected component of 7r-1(t/) is mapped difFeomorphically onto U by it, we call n a covering map. Also we say that TV is a covering manifold of M. Obviously, the dimension of a covering manifold is the same as the dimension of the original manifold. By virtue of the general theory for covering maps in a broad sense, which is developed in the context of general topological spaces, the homomorphism 7T* : ITiN —» 1X\M induced by n : N —> M between the fundamental groups is an injection, and consequently niN can be considered as a subgroup of 7TiM. Conversely, an arbitrary subgroup of it\M determines the corresponding covering space over M, and it is easy to see that this has a natural C°° manifold structure. Thus, the set of all conjugacy classes of subgroups of -k\M and the set of all isomorphism classes of covering manifolds of M can be naturally identified by the above correspondence. Among the covering manifolds of a connected C°° manifold, the most important one is the manifold M called the universal covering manifold. While the universal covering manifold is characterized as being the simply connected one among all covering manifolds, it can be practically constructed as follows. We fix a base point po of M. Two C°° curves c» : [0,1] -* M (i = 0,1) on M with terminal point po and the same starting point are said to be homotopic (with fixed end points), if they are connected by a family of C°° curves cs : [0,1) —¦ M (s € [0,1]) on M with the same starting and terminal points. We denote by M the set of all homotopy classes, with end points fixed, of C°° curves on M whose terminal points are pq. The projection 7r : M —>• M is defined by putting 7r([c]) = c@). Here, [c] denotes the homotopy class containing the C°° curve c. Now, let S — {(J7a, <pa)}a€A be an atlas of M such that each coordinate neighborhood UQ is contractible. Let pQ ? UQ be a point on a coordinate neighborhood Ua and let ca be a curve connecting pa and the base
52 1. MANIFOLDS point po € M. Then for an arbitrary point p on UQ, we choose a curve Cp connecting p and pQ in Uay and put Ua(ca) = {[cP-ca]\peU0t} (see Figure 1.22). Here, Cp ¦ ca denotes the path obtained by connecting two curves cp)ca in this order. Then, while Ua is a subset of M, the restriction of the projection ix to Ua(ca) obviously induces a one to one map n : Ua(ca) —> UQ. Figure 1.22. Construction of universal covering manifold Then, if we put «S = {(^(cJ.^ott); ae4, ca € P(p<>,Po)}> we can check that this is an atlas of M, and moreover we see that ¦n : M —> M is a covering map. Here P(pa,po) denotes the set of all (homotopy classes of) paths connecting pa to po- Now, the universal covering manifold M of M is simply connected and the fundamental group if\M of M acts on M naturally (from the right) as follows. That is, M xitiM 3 ((c],7) *-* [c-7] € M ([c] € M, 7 € ttjM). This action is easily seen to be a properly discontinuous and free action, and its quotient space is naturally identified with M. Therefore, more generally, an arbitrary subgroup T of ttiM acts on M properly discontinuously and freely. We shall prove the following general proposition. PROPOSITION 1.52. Let a group T act differentiably on a connected C°° manifold M, and suppose that it is a properly discontinuous and free action. Then a natural C°° manifold structure is induced on the quotient space M/T, and the projection it : M —* M/T is a covering map of class C°°.
1.5. FUNDAMENTAL FACTS CONCERNING MANIFOLDS 53 Proof. First, we shall show that the quotient space is a Haus- dorff space. Let p,q be two points on M lying on distinct orbits. We choose compact neighborhoods F and G of p and q such that F n G = 0 and F and G do not intersect the orbits of q and p respectively. Then, since K = F U G is compact, by the assumption there are only a finite number of 7 6 T such that K D 7/f ^ 0. Hence, putting F' = F\\J1K} G' = G\{JiK, we see that these are neighborhoods of p and q respectively. The projections of F' and G' to the quotient space are disjoint neighborhoods of the images of p and q by n. Next, by the assumption, we see that for an arbitrary point p € M, we can choose a sufficiently small open neighborhood U of it so that all 7C/ G € T) are pairwise disjoint. Therefore, each yU is mapped homeomorphically onto the projection image of U in the quotient space. It follows easily that a natural C°° manifold structure is induced on the quotient space and the projection -n : M —> M/T is a covering map. ¦ In this way, we see that the properly discontinuous and free action on a manifold and the concept of covering manifold are essentially the same. Example 1.53. Consider the unit sphere S2n+1 = {(z0, ¦¦-,*„)€ Cn+1; \z0\2 + ¦¦• + \zn\2 = 1} in Cn+1. Let p be an arbitrary natural number, and let q\, ¦ • • ,qn be n integers prime to p. We define an action of the cyclic group Zp of order p on Sn+1 by ^0,---^n) = (^0,C<?l^l,---,C<7^n), where t is its generator and C, = expBni/p). It is easy to see that this action is (obviously properly discontinuous and) free. We call the quotient manifold 52n+1 /Zp = L(p; qi, ¦ ¦ • ,qn) a lens space. EXAMPLE 1.54. By addition of vectors or, what is the same thing, by translation, an abelian group Rn acts on itself. By restricting this action to a subgroup Zn = {{mi,-- ,mn);mi € Zn} of Rn, the free abelian group Zn of rank n acts on Rn. It is easy to see that this action is properly discontinuous and free. Moreover,
54 1. MANIFOLDS the quotient manifold Rn/Zn can be naturally identified with the n- dimensional torus Tn. Since Rn is simply connected, the universal covering manifold of T1 is Rn and T^r71 S Zn. The 2-dimensional torus T2 is the genus 1 orientable closed surface Ei that we mentioned in "Outline and Goal of the Theory". Therefore the universal covering manifold of Ei is R2. It is known that for g > 2, the universal covering manifold of E5 is also R2. In this case, although 7TiEp is not an abelian group any more, it acts on R2 properly discontinuously and freely. Summary 1.1 A Hausdorff space satisfying the second countability axiom and locally homeomorphic to an open set of Rn is called an n-dimensional topological manifold. 1.2 If a homeomorphism </? : U —¦ Rn from an open set U of a topological manifold into Rn is given, we call the pair A7, <p) a local coordinate system and U a coordinate neighborhood. 1.3 A family of coordinate neighborhood systems of a topological manifold M is called an atlas if its coordinate neighborhoods cover all of M. 1.4 A topological manifold with an atlas whose coordinate changes are all of class C°° is called a C°° manifold. 1.5 If a group structure is defined on a C°° manifold and the operations of multiplication and taking the inverse are both of class C°°, this is called a Lie group. 1.6 A tangent vector at a point on an n-dimensional C°° manifold gives "a direction to partially differentiate" functions defined around that point. The set of all tangent vectors at each point forms an n-dimensional vector space. We call this the tangent space at that point. 1.7 A C°° map from a C°° manifold to a C°° manifold is called an immersion if its differential is an injection at any point. 1.8 A C°° map that is an immersion and a homeomorphism onto its image is called an embedding. 1.9 If at each point of a C°° manifold a tangent vector which moves smoothly depending on the point is given, it is called a vector field. 1.10 A vector field acts on the commutative algebra of all C°° functions as a derivation.
EXERCISES 55 1.11 An orientation on a manifold is an assignment of an ordered basis to the tangent space of each point so that they are coherent along all the paths. Exercises 1.1 For a natural number ra, define a map fm : C —¦ C from the complex plane C to C by fm(z) = zm (z € C). Let z = x + iy (x, y e R), and consider fm as a function of x and y. Compute the Jacobian matrix of fm. 1.2 Prove that the set of all the orthogonal matrices of order 2, denoted 0B) (the orthogonal group of order 2), becomes a C°° manifold in a natural way (Example 1.20). 1.3 Show that 1-dimensional complex projective space CP1 is dif- feomorphic to S2. 1.4 Show that 50C) is diffeomorphic to the 3-dimensional real projective space RP3. 1.5 Let M, TV be C°° manifolds and / : M — N a C°° map. Then, show that Tf = {(p, /(p)); p € M} is naturally a submanifold of the product manifold M x N. We call Tf the graph of /. 1.6 Show that a linear map L : Rm -> Rn from Rm to Rn is of class C°°. Moreover, if the tangent space at an arbitrary point on Rm is naturally identified with Rm and the same is done for Rn, prove that the differential L* of L at an arbitrary point on Rm coincides with L itself. 1.7 Prove Proposition 1.40. 1.8 Prove Proposition 1.35. 1.9 Let M be an orientable manifold with boundary. Prove that its boundary dM is also orientable. 1.10 Prove that the real projective space RPn is orientable if n is odd and non-orientable if n is even.
CHAPTER 2 Differential Forms In this chapter, we shall define the leading character of this book, the differential forms on differentiate manifolds. Differential forms have two main roles. One is that they describe various system of partial differential equations on manifolds, and, ever since the pioneering work by Pfaff in the 18th and 19th centuries, they have played an important role in analysis. The other is that they are used to express various geometric structures on manifolds. By applying appropriate operations on those differential forms, various kind of differential forms are induced, and by integrating them on manifolds, certain geometric "invariants" are obtained. These invariants are quantities that reflect the global structure of manifolds, and are very important- in fact, indispensable - in the study of manifolds. The above two roles of differential forms are deeply related to each other, rather than independent. However, in this book, keeping mainly the second role in mind, we shall introduce differential forms. That is, we consider differential forms to be something "which should be integrated on manifolds". 2.1. Definition of differential forms (a) Differential forms on Rn. We start with differential forms on Rn, for the sake of simplicity. Recall that if an associative product is defined on a vector space A over the real number field R so that a ring structure is given and for arbitrary oGR and A, /i € A the condition a(Xfi) — (a\)n = X(afj.) is satisfied, then A is called an algebra over R (Definition 1.23). An algebra generated by dxi, ¦ • ¦ ,dxn over R with unity 1, that satisfies the equation B.1) cte, A dxj = -dxj A dxi 57
58 2. DIFFERENTIAL FORMS for arbitrary i,j, is denoted by A*. Here A is a symbol that stands for the product of this algebra. We call A* the exterior algebra generated by dx\, • • • ,dxn. By B.1), we see that dxi A dxj = 0 for arbitrary i. By taking the degree of dxi to be 1, for each monomial of A* the degree is defined. For example, the degree of dx\ /\dx2?\dx$ is 3. If we denote by A? the set of all linear combinations of monomials of degree k, the direct sum decomposition A; = 0An = A^eAie---eA^ Jt=0 holds. It is easy to see that as a basis of A? we can take B.2) dxix A • • • A dxik, 1 < ii < • • • < ik < n, and hence dim A? = (?). Also if k > n, then A* = 0 and dim A* = 2n. A linear combination U = >J /»i -tfc(xl> * * " yxn)dXil A • • • A dXjfc tl< <ik of each element of B.2) with C°° functions on Rn as coefficients is called a degree k differential form on Rn, or simply a fc-form. The above description is sometimes simply denoted by >J // (x)dxil A • • ¦ A dxik. i We usually express differential froms by Greek letters. We denote the set of all /e-forms on Rn by Ak{Rn). More precisely, Ak{Rn) = {u;: Rn -» A?; C^map} or Ak{Rn) = C°°{Rn)®A*. Collecting differential forms of each degree, we can consider the algebra of all differential forms on Rn, n A*{Rn) = Q)Ak{Rn). fc=0 In particular, .4°(Rn) = C°°(Rn). That is, differential forms of degree 0 are simply C°° functions. The product u>At) G .4fc+i(Rn) of a fc-form u € .4*(Rn) and an /-form t? e Al{Rn) is defined by u) A 77 = 2^, fl9J dxix A • • • A dxik A dxjt A • • • A dxjt 1,J
2.1. DEFINITION OF DIFFERENTIAL FORMS 59 if they are expressed as u) = Y^ fi(x)dxi1 A • • • A dxik, 77 = \J9j{x)dXjl A • • • A dxj,. / J We call this the exterior product of u; and 77. In the above description, if we replace Rn by an open set U in Rn, we can consider the algebra A*{U) of all differential forms on U. Example 2.1. Put U = R2 - {0}. Then, -5-—i>dx + -5——jdy x2 + y2 x2 + y2 is a 1-form on U: However it is not a 1-form on R2, because it is not defined at the origin. The exterior differentiation, which is an important operation applied to differential forms, is a linear map d:^(Rn)-^^fc+1(Rn), defined as follows. That is, for u> = f(x\,-•¦ ,xn) dxix A• • • Adxik, let B.3) duj = J2^-(x)dx3,Adxi} A--- Adxik. For a function / e .4°(Rn) on Rn, its exterior differentiation df € ^x(Rn) is df = Y^ -—dxi and is equal to so-called total differential. Z—' OX{ For practice, let u> be the 1-form in Example 2.1; if we calculate its exterior differentiation <ko by definition, we have V — x y — x ^ = t 2 ¦ i\i dyAdx + dxAdy = 0. [x2 + y2J {x2 + y2J Lemma 2.2. // we repeatedly operate the exterior differentiation twice, it is identically 0. That is, do d = 0. Proof. If we operate d again on dw in B.3), we have n n d2 f d(dw) = Y,Y, ^rir dxtAdxjAdXi, A • • • Adxik. Then the facts that the order 2 partial differentiation with respect to Xj and xi does not depend on the order and dxi A dxj = -dxj A dxi immediately imply d{duj) = 0. ¦
60 2. DIFFERENTIAL FORMS A differential form u> such that du> — 0 is called a closed form, and a differential form 77 that can be written 77 = du for some u is called an exact form. The above Lemma 2.2 claims that exact forms are always closed forms. Conversely, there arises a natural question whether closed forms of degree k are always exact, and we will find later that in the case of Kn, this is true for k > 0 (§3.3, Poincare lemma (Corollary 3.14)). However, in the case of general C°° manifolds, a closed form is not always exact, and the "gap" will reflect the global structure of manifolds. This is the content of the theory of de Rham cohomology, which is the theme of Chapter 3. Since the proof of the following proposition is easy, we leave it to the reader (Exercise 2.1). Proposition 2.3. Foru e .4*(Rn) and 77 e .4'(Rn), we have (i) t)/\uj = (-l)klu A 77, (ii) d{uj A 77) = ch A 77 + {-l)ku> A drj. Now let U, U' be two open sets in Rn and v? : U —> U' a diffeo- morphism. Then a homomorphism <p*:A*{U')—*A'{U) from the algebra A*{U') of all differential forms on U' to the algebra A*(U) of all differential forms on U is defined as follows. For an arbitrary function / € A°(U'), let ^(/) = /o^e A°{U) and let (p*(dxi) = d(<p*(xi)). We extend this to differential forms of general degree in such a way that V?*(wA77) =?>*((*;) A ^G7) for an exterior product a; A 77 of two differential forms. Practically, we proceed as follows. Let the coordinates of U' be xi, ¦ • • ,xn and the coordinates of U be y\, • • • , yn (to distinguish these from the coordinates of U'). Then each Xi is written as a function x» = X{(y\, ¦ • ¦ ,yn) of 2/1, • • • ,Vn- Then we have </?*(dxi) = > —— dj/,-, and from this we see that B.4) </(<irtlA...AdxtJ= J2 nf!i|,!",!<*|d^A---Ady>-
2.1. DEFINITION OF DIFFERENTIAL FORMS 61 D(Xi ¦ ¦ ¦ ,Xi ) Here ^, lI> T\ denotes the Jacobian of xv,, • • • , xvt with re- spect to j/jj, • • • ,y^^. Then we see (verification is Exercise 2.2) that do<p* = tp* od, and by the consideration of v?-1, we can verify that <p* is in fact an isomorphism. Henceforth, </?*(u;) will sometimes be denoted simply by <p*v. (b) Differential forms on a general manifold. Let M be an n-dimensional C°° manifold and {(?/<*. <Pa)} an atlas of it. In brief, a degree k differential form on M is a family {u>a} of /c-forms u>a on each coordinate neighborhood UQ (which can be considered as an open set of Rn) such that for arbitrary a,/3 with UaC\Up ^ 0, ua and up are transformed to each other in the sense of B.4) by the coordinate change. We denote the set of all /c-forms on M by Ak(M), and we put A*(M) = ®Ak[M). k=0 As we saw in the previous subsection (a), the homomorphism (f* : A*{U') —> A*(U) between algebras of all differential forms induced by a coordinate change preserves the exterior products and commutes with the operation of exterior differentiation. Prom this, we see that the exterior products and the exterior differentiation d : Ak(M) —> Ak+l(M) are defined also on A*(M), and d o d = 0. Furthermore, Proposition 2.3 holds for differential forms on M. Although this definition is right, the formula B.4) is fairly complicated, and from the standpoint of studying the whole M it may not give a good insight. Therefore, we shall define these differential forms independently of the local coordinates. We need to prepare some abstract facts for it. It is not appropriate to ask which of these two definitions is better, and the important thing is that we learn from them what differential forms are after all. (c) The exterior algebra. We shall start by giving the relationship between the exterior algebra A* generated by dx\, ¦ • ¦ , dxn and the tangent space ToRn of Kn at the origin. ToR" is an n-dimensional vector space with a basis d d "^—»•••) -z—• On the other hand, each dxi can be considered as an ox i dxn
62 2. DIFFERENTIAL FORMS element of the dual space Tq Rn = {a : T0Rn -+ R ; a a linear map} of TbRn. This is because x{ can be considered as a C°° function Xi : Rn -+ R and the differential dx{ : T0Rn -> T0R = R of this function at the origin is linear. Then obviously B.5) "Kj-H- From another point of view, since -— is a unit tangent vector in the OXj direction of Xj, we can consider that B.5) reflects the fact that if we integrate the constant function 1 with respect to Xi from 0 to 1 along the Xj-axis, the value is 6ij. Thus A^ is identified with ToRn: Ai = T0*Rn. In general, an arbitrary element in A J is described as a linear combination of the elements of the form u> = cn\ A • • • A ajt {cti G A\), while such an u> defines a map B.6) w : T0Rn x • • • x TpR^ — R k as follows. That is, for Xi € T0Rn (i = 1, • • • , k), we put B.7) u{Xu~- ,Xfc) = idet(ai(Xj)). Here, (ai{Xj)j denotes a matrix whose (i,j)-entry is oti(Xj). Using the properties of determinant, it is easy to see that the above value is uniquely determined, independently of the expression of u>. For example, if we write u> = -a^ A c*i A a^ A • • • A ctk, the value is the same. The geometric meaning of this value is roughly as follows. For example, dx\ A ^2(^1,^2) is the (signed) area of the orthogonal projection of the triangle spanned by two tangent vectors X\, X2 in 7bRn onto the (xi,^-direction, and in general, B.7) is considered to present "the (signed) volume in the direction of (a\}- • • ,0^)" of the /c-dimensional simplex (a generalization of triangle, see §3.1) spanned by ^i, • • • , Xk. If we recall these facts when we define the integration of differential forms on manifolds later in Chapter 3, it may help our understanding. For a general element ueAj, the map B.6) is also defined by extending the above definition linearly.
2.1. DEFINITION OF DIFFERENTIAL FORMS 63 We see that the map u> of B.6) has the following two properties. Since the proof can be given easily by using the properties of determinant, we leave it to the reader. (i) u) is multilinear. That is, for an arbitrary Xj, it satisfies the linearity condition uj{Xu--- ,Xi-UaXi + bX'i,Xi+u--- ,Xk) = au(Xu--- ,XU-- ,Xk) + bu){Xu--- ,X;,--. ,Xk). (ii) u) is alternating. That is, for arbitrary i < j, if we interchange Xi and Xj, its sign changes. Therefore for an arbitrary permutation a € 6n of n letters, <*>(*<7(i)»'" >*a(n)) =sgncrw(Xi,--- ,Xn). Here sgn a denotes the sign of a. We call the map T0Rn x • • • x T0Rn (n-fold direct product) -> R satisfying the above two conditions an alternating form of degree k on 7oRn. As a result, by the correspondence B.6), a map B.8) A? = all alternating forms of degree k on T0Rn is defined, and this turns out to be a one to one correspondence. Here, the right-hand side of B.8) does not contain the coordinates Xj of Rn and is presented purely in terms of linear algebra. With this in mind as a clue to go on, we shall give a definition of differential forms on general manifolds which is independent of the coordinates. We shall describe it without worrying about some repetition. Let V be a vector space over R. Since we need only the case of tangent space ToM at a point p on a C°° manifold M, it may be read as V = TPM. The dual space V* of V is a vector space defined as V* = {a : V —> R; a a linear map }. Definition 2.4. Let V be a vector space over R. An algebra with unit 1 generated by the elements of V over R satisfying the relation B.9) X AY = -Y AX for arbitrary X, Y € V is denoted by A*V and called an exterior algebra of V or a Grassmann algebra. Here A stands for the product of this algebra. By condition B.9), X A X = 0 for an arbitrary X € V. Conversely, it is easy to see that B.9) follows from this condition. The
64 2. DIFFERENTIAL FORMS previous A* is nothing but A*ToRn. In the same way as in the case of A*, if dim V = n, we have a direct sum decomposition A*V = 0AfcV. fc=0 Here AfcV is the subspace of A* V consisting of all elements of degree k. Let ei, • • • , en be a basis of V. Then we can take B.10) eix A • • • A eik, 1 < »i < • • - < iik < n as a basis of AkV} and therefore dimAfcV = (?). Also, A°V = R and AlV can be naturally identified with V. While we defined the exterior algebra of V, the exterior algebra A* V* of V* is also defined similarly. It is this case that we use later. Next we shall define alternating forms on V. Definition 2.5. Let V be a vector space over R. A multilinear map w : V x • • ¦ x V —> R k from fc-fold direct product of V to R that is alternating, namely u{Xo{l) ¦ •¦*<,(*>) = sgn ou{Xu- ¦ • ,Xk) (Xi € V) for an arbitrary permutation a of k letters, is called an alternating form of degree k on V. The set of all alternating forms of degree k on V is denoted by Ak(V). Ak(V) is a vector space with respect to the natural sum and the multiplication of alternating forms by real numbers. We shall consider all alternating forms A*(V) = ($Ak(V) fc=0 with different degrees on V. Here we define A°(V) = R, and it is easy to see that Ak(V) = 0 for k > dim V, by the alternating condition. A degree preserving linear map l: A*V* —^A'{V) from the exterior algebra A* V* of the dual space V* of V to the vector space A*{V) of all alternating forms on V is defined as follows. It is enough to define tfc : AkV* —+ Ak{V)
2.1. DEFINITION OF DIFFERENTIAL FORMS 65 for each k. For an element of the form u> = ai A- • ¦ Aajt G AkV* (a, G V*), we set tk(u){Xu--.,Xk) = ±det(ai{Xi)) and extend it linearly for general elements. It is easy to see that tfc is well defined independently of the expression of cj, by using the properties of determinant in the same way as before. PROPOSITION 2.6. The map i : A*V —> A*{V) is an isomorphism. That is, the exterior algebra A*V* of V* and the vector space A*{V) of all alternating forms on V can be identified by i. Using this, a product is defined on A* (V) which is described as follows. If for u> G AkV* ,n € A*V*, we consider their exterior product uj An as an element of Ak+l{V) by the identification l, we have B.11) w A77 {Xu--- ,Xku) = (k + nj S SSn ° ^Xo{\).' • ' >Xo(k)) ^(^<r(fc+l).-" ^a(k+l)) (Xt G V). Here a runs over the set &k+i of all permutations of k + I letters 1,2,--- ,k + l. Proof. First, we show that t* is an isomorphism. Let e\, ¦ • • , en be a basis of V and ai, • • • , an its dual basis of V. They satisfy a»(Cj) = 6ij. Then by B.10) we can take a^j A • • • A Qjk, 1 < i\ < ¦ • ¦ < ik < n as a basis of AkV*. We can check that the images of elements of this basis by t are linearly independent as elements of Ak(V) by applying them to (ej!,--- .e>j€ V x---xV, ji < ¦•¦ < jfc. Next, let u) G Ak{V) be an arbitrary element. Then if we set u;(et,, • • • ,eik) — a^...^ and, using these constants, define ? = fc! ^2 an ik<*ii A'-- Aan e AkV, we see that t;t(eD) = u. Therefore, tfc is a surjection and hence an isomorphism.
66 2. DIFFERENTIAL FORMS Next, we prove the latter half of the claim. It is enough to prove it for the elements a;, 77 of the form uj — o.ix A • • • A (Xik, 77 = ctjj A • • • A (Xjl by the linearity of i^. Furthermore, we may assume that Ji > • • • , 31 are all distinct, because otherwise we have u> A 77 we rearrange these numbers in order of size so that mi < • If we let the permutation of rearrangement be r, we have uj A 77 = sgn TCtmx A • Therefore, tfc+i(wA»7)(eTOl,--- ,emjt+l) On the other hand, if we calculate X^sgncr tfc(w)(cma(l),.-- ,emc(M)t!()))(em,(H1I'-' >erna(k+l)), we see that it is sgn r. In this way, we see that the claim is true for (em,, • • • ,emfc+l).,But since for every other element of the form (eni»• • • , Cnfc+j) the value is 0, the proof finishes. ¦ The above isomorphism 1 : A*V* = A*{V) is not the unique natural one. Actually, if we let t'k = k\tk, we obtain another isomorphism l' : A*V* = Am{V), and this defines another product on A*(V) (however, the difference between the two products is only up to scalars and is not essential). This is equivalent to considering the volume of the parallelotope spanned by each vector instead of the volume of the fc-dimensional simplex defined by the origin and the end point of each vector in the description following B.7). While these two methods have their own merits, we use t in this book because there are some inconveniences with 1' when we describe the general theory of characteristic classes in Chapter 6. However, since 1' is defined over Z, it has the advantage of eliminating fractional constants in various formulae. For example, the coefficient in the formula of exterior k + 1 differentiation (Theorem 2.9) is not necessary if we use l'. (d) Various definitions of differential forms. While we have already defined differential forms on general C°° manifolds in subsection (b), in this subsection we shall give a more intrinsic definition without using local coordinates. = 0. Then ¦ Aamk + l- (k + l)\ sgn r.
dxA 2.1. DEFINITION OF DIFFERENTIAL FORMS 67 The dual space T*M of the tangent space TPM at a point p on M is called the cotangent space at p. By the description in the previous subsection, we can consider its exterior algebra A'T*M. Definition 2.7. Let M be a C°° manifold. We say that u> is a fc-form on M if it assigns ujp € A*T*M to each point p € M and uip is of class C°° with respect to p. Let U be an arbitrary coordinate neighborhood, and x\, • • • ,xn coordinate functions defined on U. Then, for any point p G U, \dxx)p \dxn)p become a basis of the tangent space TVM. We shall find the dual basis for the dual space T*M. Each xt can be regarded as a C°° function x{ : U -> R. Consider the differential (dxi)p : TPM -* TXi(p)R of this map at p. Since TIt(p)R can be naturally identified with R, we can consider {dxi)p as an element in T*M. Then obviously, (see B.5)). Therefore, (dx^p,--- }(dxn)p become the dual basis of T*M. It follows from this fact that u;p in the above Definition 2.7 is presented as B.12) up= J2 ft, ¦xMdxilA---Adxik. ii<-<ik u)p is said to be of class C°° if each coefficient f^-..^ (p) is of class C°° as a function of p. The expression B.12) is called the local expression of the A;-form u> on M. Thus, Definition 2.7 and the definition in subsection (b) are related. If we use the terminology of vector bundles which will appear in Chapter 5, we can interpret the above as follows. If we set t-m = \Jt;m, p it is easy to see that this is a vector bundle over M. We call this the cotangent bundle of M. Similarly, if we set Afcr*M = (J Afcr;M,
68 2. DIFFERENTIAL FORMS this is also a vector bundle over M. Note that A.lT*M = T*M. In these terms, k-forms on M are nothing but sections of AkT*M of class C°°. That is, Ak{M) = all sections of AkT*M of class C°°. Finally, we mention another view of differential forms. Let u be a fc-form on M. Then the value u>p of u at each point p determines an alternating form TPM x • • • x TPM —> R of degree k. Putting all p together, u> induces a multi-linear and alternating map B.13) a; : X(M) x • • • x X(M)—> C°°(M). Here X(M) denotes the set of all vector fields on M and C°°{M) denotes the algebra of all C°° functions on M. It is important here that X(M) is not only a vector space over R but also a module over C^iM). That is, for / G C°°{M) and X e X(M), fX is also a vector field on M. Then the meaning of B.13) being multilinear is that it is also linear with respect to the multiplication of vector fields by functions. More precisely, u(Xu--- JXt+gX'i1--- ,Xk) =M^i,"- ,*»,¦•• ,xk) + guj(Xu--- .x;,--- }Xk) for arbitrary Xi e X(M) and f,g € C°°(M). Conversely, we see that any map B.13) with these two properties (that is, multilinear as a C°°(M) module and alternating) defines a differential form. Namely, the following theorem holds. Theorem 2.8. Let M beaC°° manifold. Then the set Ak{M) of all k-forms on M can be naturally identified with that of all multilinear and alternating maps, as C°°(M) modules, from k-fold direct product ofX{M) toC^iM). Proof. Suppose that a map Zj : X(M) x • • • x X{M) -> C°°(M) with the above conditions is given. First of all, we shall see that for arbitrary vector fields Xi € X, the value uj(X\,--- ,Xk){p) at a point p is determined depending only on the values Xi(p) of each vector field Xi at p. For that, by linearity, it is enough to show that if Xi(p) = 0 for some i, then the above value is 0. For the sake of simplicity, assume that i = 1, and let (U;xi,--- ,xn) be a local coordinate system around p. Then we can write X\ = V^ /{ -— on Z—' OXi U with fi(p) = 0. We choose an open neighborhood V of p such that V C U, and a C°° function h € C°°(M) such that it is identically 1
2.2. VARIOUS OPERATIONS ON DIFFERENTIAL FORMS 69 on V and 0 outside of [/(see Lemma 1.28). Let Yi = h -—. Then we axi have Yi € ?(M), and if we set fi = h fit then we have U € C°°{M). Now it is easy to see that *i=?/in + (l-/i2)Xi- Therefore, we have ?(Xi,--- ,Xfc)(p) = SA(p)S(yn^2.---,^fc)(p) + (i-%J)S(xll--.>A:fc)(p) = ol and the claim is proved. Now we define a fc-form u> as follows. At each point p e M, if tangent vectors X\, • ¦ • , Xk € TPM are given, we choose vector fields Xi over M such that X{(p) = Xx. If we let uip{Xu • • ¦ ,Xk) = uj(Xi, • • • , Xfc)(p), then, as we saw above, this is determined independently of the choice of Xi. Since it is easy to see that u>p is of class C°° with respect to p, u; is the required differential form. ¦ 2.2. Various operations on differential forms Let M be an n-dimensional C°° manifold. We denote all k-forms on M by Ak(M) and consider their direct sum X(M) = ® Ak(M) Jfc=0 with respect to k, that is, the set of all differential forms on M. In this section, we shall define various operations on A*{M). (a) Exterior product. The exterior product u>Atj € Ak+l(M) of a fc-form uj € Ak(M) and an /-form rj G Al(M) on M is defined as follows. Since at each point p € M we have up G AkT*M, r\p e AlT*M, their product wp A r)p e Ak+lT*M is defined. Then, we put (w A t?)p =wpA 77p. By definition, the exterior product is obviously associative. That is, if r ? Am(M), we have (w A r/) A r = u> A G7 A r). Therefore
70 2. DIFFERENTIAL FORMS we do not need the parentheses. If they are locally expressed as u) = / dxir A • • • A dxik ,77 = g dxjl A • • • A dxjt, we have u) A 77 = fg dx^ A • • • A dxik A dxjl A • • • A dxj,. The exterior product induces a bilinear map Ak{M) x .4Z(M) 9 (W,T7) ~ w A 77 e Ak+l(M) and it has the following properties, (i) 77 Aw = (-l)fc/WA77. (ii) For arbitrary vector fields Xi, • • • , Xk+i € ?(M), B.14) u Ar)(Xi,--- ,Xfc+0 = (fe ¦ ni S sgno'u;(^a(l),--- >*»(*)) ^(^(fc+l).--' •^(fc+o)- Property (i) is obvious from the description above, and (ii) follows from B.11). (b) Exterior differentiation. For a fc-form u> G Ak{M) on M, its exterior differentiation dw € Ak+1(M) is the operation defined by du) = > -— dxi A dxi, A • • • A dxi. ; ^ OXj J j J here u> is locally expressed asw = / dx^ A • • • A ekc^. In view of the fact that for the isomorphism (p* : A*{U') —» «4*(?/) induced by an arbitrary diffeomorphism (p : U -+ U' between two open sets U, U' of Rn, the equation doip* = ip* od holds (see the description following B.4)), we see that the above d does not depend on the local expression. Therefore, the operation of taking the exterior differentiation defines a degree 1 (that is, increasing the degree by 1) linear map d:Ak{M)—+Ak+l{M), and from Lemma 2.2 and Proposition 2.3, we see that it has the following properties. (i) dod = 0. (ii) For u) € Ak{M), d{u A 77) = dw A 77 + (-1)* u A d-q. Next, we shall characterize the exterior differentiation without using the local expression. Namely, we have the following theorem.
2.2. VARIOUS OPERATIONS ON DIFFERENTIAL FORMS 71 Theorem 2.9. Let M be a C°° manifold and u> e Ak(M) an arbitrary k-form on M. Then for arbitrary vector fields X\, • • • , Xk+i € X{M), we have du){Xi,--- ,Xk+i) +]?(-i)i+M[*i.*J-].*i,"- ,?,¦¦¦ .*;,¦•¦ .A-fc+,)}. /fere ?/ie symbol Xi means Xi is omitted. In particular, the often-used case of k — 1 is dw{X,y) = ^PM^) - Yu>{X) - o>{[XyY))} (w G ^(M)). Proof. If we consider the right-hand side of the formula to be proved, as a map from the (k + l)-fold direct product of 3C(M) to C°°(M), we see that it satisfies the conditions of degree k + 1 alternating form as a map between modules over C°°(M). Since it is easy to verify this fact by using Proposition 1.40 (iv), we leave it to the reader. Therefore, by Theorem 2.8, we see that the right-hand side is a (k + l)-form on M. If two differential forms coincide in some neighborhood of an arbitrary point, they coincide on the whole. Then, consider a local coordinate system (?/; x\, ¦ ¦ ¦ .xn) around an arbitrary point p € M. Let the local expression of uj with respect to this local coordinate system be u; = Y^ /u • tfc dxiv A • • • A dxilc. Then, we have B.15) dw = ^ dfi^.^dx^ A--- /\dxik. h<-<ik From the linearity of differential forms with respect to the functions on M, it is enough to consider only vector fields Xi such that Xi — -— (t = 1, ¦¦¦ ,k -r 1) in a neighborhood of p. Then [XitXj] = 0 near p. Moreover, by the alternating property of differential forms, we may assume that ji < • • • < jk+\- Then, if we apply B.15) to
72 2. DIFFERENTIAL FORMS {Xi,-- ¦ ,Xk+i), we have On the other hand, when we calculate the right hand side of the formula using [Xi, Xj) = 0, we obtain the same value. This finishes the proof. ¦ We can consider Theorem 2.9 as a definition of the exterior differentiation that is independent of the local coordinates. (c) Pullback by a map. We shall study the relationship between differential forms and C°° maps. Let be a C°° map from a C°° manifold M to N. Consider the differential /, : TPM -* Tf(p)N of / at each point p € M. /. induces its dual map /* : Tj{p)N -* T'M, that is, the map defined by f(a){X) = a(/.(X)) for a 6 T}{p)N, X € TPM. Furthermore, /* defines a linear map /* : A.kTJ^N -* AkT*M for an arbitrary k, and they induce an algebra homomorphism For a differential form u € Ak(N) on N, f*u> € Ak{M) is called the pullback by /. Explicitly, for Xu • • • , Xk € TPM, fu>(Xu--- tXk)=u{fmXw- ,f.Xk). Proposition 2.10. Let M, N be C°° manifolds. Let f : M -> N be a C°° map and f* : A*{N) —> A*{M) the map induced by f. Then f* is linear and has the following properties. (i) /•(c;Aij)=/*u;A/*i7 (w € Ak(N), tj 6 AL(N)). (ii) d(f*u) = r{dw) (u, € Ak{M)). Since the proof can be given easily by using the previous results, we leave it to the reader. (d) Interior product and Lie derivative. Let M be a C°° manifold and X € 3C(M) a vector field on M. Then a linear map i{X):Ak{M)—>Ak-\M)
2.2. VARIOUS OPERATIONS ON DIFFERENTIAL FORMS 73 is defined by (i(X)u)(Xu--- ,Xk-l) = ku>{X,Xu--- ,**.,) for u> e Ak(M), Xu ¦ • ¦ , Xk-i e X{M). Note that if k = 0, we define i(X) = 0. We call i{X)u) the interior product of u by X. By definition, x(X) is obviously linear with respect to functions. That is, i(X)(fu) = fi(X)u>. Using Proposition 2.6, we see that i(X) is an anti-derivation of degree -1, that is, i(*)(wAT7) B'16) =i{X)ujAr} + {-l)kujAi(X)r) {u> e Ak{M), -q e Al{M)). Next, we shall define a linear operator Lx :Ak{M)^Ak{M), called the Lie derivative, also concerning the vector field X € X(M). This is defined by B.17) (Lxuj)(Xu.-,Xk) k = Xo;(X1,---,Xfc)-^u;(X1,---,[X)Xi])-..,X,). It is easy to see that the right-hand side of this formula satisfies the condition of Theorem 2.8, so that Lxu is definitely a differential form. Obviously Lx is linear. This definition B.17) is extremely algebraic. Although we may say that the formula is neat and beautiful, it is not clear what it means geometrically. We shall give a definition that makes the meaning clearer in the next subsection. Similar things can be said also for the exterior product and the exterior differentiation. We first introduced both exterior product and exterior differentiation with geometric definitions in terms of local expressions. However, leaving them aside, we can use the formula B.14) for exterior product and Theorem 2.9 for exterior differentiation as algebraic definitions. As for the Lie derivative, we use B.17) as its definition for the moment, and proceed. (e) The Cartan formula and properties of Lie derivatives. The following theorem represents the relationship between two operators concerning a vector field X, namely, the interior product i{X) and the Lie derivative Lx, and is sometimes called the Cartan formula.
74 2. DIFFERENTIAL FORMS Theorem 2.11 (Cartan formula), (i) Lxi(Y)-i(Y)Lx = i([X>Y}). (ii) Lx =i{X)d + di(X). Proof. First we prove (i). It is obvious for k — 0, so let u> be an arbitrary /c-from with k > 0. Then, for any Xi, • • • , Afc-i € 3C(M), (Lx^M^h-.^-i) = X((t(KH(X1)...,Xfc_1)) fc-1 -^(yM^-.ix,^],-,^.,) B.18) »=i = k{x(uJ(Y,Xu.--,Xk.l)) Jt-i On the other hand, (i(Y)Lxu>)(Xl,--.,Xk-1) = kLx^{Y,Xu---,Xk-l) B.19) =fc{x(o;(K,X1,--.,Xfc_1))-a;([X,r],X1,-..)Xfc-1) t=i Subtracting B,19) from B.18), we have Lxi{Y)u> - i{Y)Lxu = i([X,Y])uj, and the proof is finished. Next we shall prove (ii). When k = 0, since Lxf = Xf for a function / and on the other hand i{X)f = 0 and i(X)df = df(X) = Xf, (ii) holds. Thus, let k > 0, and let w be a /c-form and
2.2. VARIOUS OPERATIONS ON DIFFERENTIAL FORMS 75 Xi, • • • ,Xk vector fields. Then, we have B.20) (i(X)dw)(Xu--- ,Xk) = (fc+l)cM*,*i. ••¦,**) = x{u>{Xi,- ¦ •, **)) + ^(-lyx.Mx, xlt • • •, xt, ¦ ¦ •, xk)) it + ^(-i)M[^^].^i,--,^.-".-X'fc) 3=1 «<; and on the other hand we have B.21) (di(X)w)(Xlr--,Xk) ^(-ir+^Mx,*!, ••-,?, •••,**)) t=i +X>ir+M*> [*<>*,], *!>••• ¦*»•¦• »**•¦••,Xfc)- Summing up B.20) and B.21), we have (i(X)d + di(X))u>(Xly--- ,Xk) k = xmx1,-.. ,xfc)) + E(-i)M[^.^]>Xi,--- ,*,,.¦• fA-fc) = (Lxu/)(Xi, ••¦,**), and (ii) is proved. ¦ Using the Cartan formula (Theorem 2.11), we can prove some properties of the Lie derivative Lx- Proposition 2.12. (i) Lx(wAjj) = LxwAt] + wALx'J (w € Ak{M),Tj e Al{M)). (ii) Lxdu; = dLx<*> (u> e .4fc(M)). (Hi) L*Ly - LyLx = L,Xlyj (*> ^ € 2(M)).
76 2. DIFFERENTIAL FORMS PROOF. Since the proofs of (i) and (ii) are easily done by using the Cartan formula, we put them as Exercise 2.3. We shall prove (iii). We use induction on k. First, if k = 0, since L[x,y]/ = [X, Y]f = {LxLy - LyLx)f for a function /, it certainly holds. Next assume that it is true up to k (> 0), and we shall prove the case of k + 1. Let u> be an arbitrary (k + l)-form. Then since, for an arbitrary vector field Z, i(Z)u) is a fc-form, by the assumption of induction we have B.22) Llx,Y)i{Z)v = {LxLy - LYLx)i{Z)u. On the other hand, by the Cartan formula (i), we have B.23) L[x,Y)i(Z) = i(Z)L[x,Y] + t([[X, Y], Z\)t and, again using (i), B.24) LxLYi(Z) = Lx{i{Z)LY + i{%Z))) = i{Z)LxLY + i([X, Z))LY + i({Y, Z))LX + i([X, [Y, Z})), and similarly, B.25) LYLxi{Z) = i(Z)LYLx + t([Y, Z\)LX + i([X, Z})LY + t((Y, [X, 2]]). Subtracting B.25) from B.24), we have LxLYi(Z) - LYLxi(Z) {- } =i(Z)(LxLY~LYLx)+i(\X,[YyZ}))-i([Y>[X,Z}}). Also subtracting B.26) from B.23), we have B.27) {L[x,Y) - LxLy + LyLx)z(Z) = i(Z)(L[X,Y) ~ LxLy + LyLx). Here we used the Jacobi identity [[X, Y], Z] + [[Y, Z], X] + \\Zy X], Y) = 0. If we substitute B.27) in B.22), we have i{Z)(L[XtY) ~ LXLY + LYLx)u = 0. Here, since Z was an arbitrary vector field, we obtain (L[X,Y] - ^xLy + LyLx)u = 0, and the proof is finished. ¦
2.2. VARIOUS OPERATIONS ON DIFFERENTIAL FORMS 77 (f) Lie derivative and one-parameter group of local transformations. Here, as we promised in the previous subsection, we shall give a more geometric definition of the Lie derivative. Suppose a vector field X is given on a C°° manifold M. We can consider X as an assignment of a direction Xp € TPM at each point p on M. Therefore, for instance, if a C°° function / E C°°{M) on M is given, we can "differentiate / in the direction of X". This is nothing but Xf. Now what would happen if a differential form is given on M instead of a function? Since a function is a special case (the case of degree 0) of differential forms, it will be natural to try to "differentiate" also a general differential form u> in the direction of X. Actually, such a natural operation is defined, and furthermore it operates not only on differential forms but also on so-called tensors, which is a notion including vector fields on a manifold. We call this a (general) Lie derivative. The geometric definition of a Lie derivative is given using the one-parameter group {<pt} of local transformations on M generated by X (see § 1.4(c)) rather than the vector field X itself. At first, we shall study the relationship between the differential Xf of a function/ € C°°(M) by X and the one-parameter group of local transformations. The result is B.28) W)(P) = lim^LQM^iM (peM) (here iplf stands for / oipt). This follows because, by the notation of §1.4 (c), tpt(p) = c(p)(t) and c(p)@) = Xp, we have lim (v»f/)(p)-/(rt = lim MHW = x/_ t-o t t-0 t p Though (ft is not always defined on the whole of M, for each point p € M, (ft is defined in a neighborhood of p for sufficiently small t, and there is no problem in the above calculation. Next we shall see that the bracket [X, Y] of vector fields can be considered as the Lie derivative of Y by X (the symbol LxY is used). That is, B.29) [X, Y) = lim ^-t)*^~Y Here, the equation B.29) means that the values of both sides are equal at each point p on M, and then the limit of the right-hand side is taken with respect to the usual topology of TPM as a vector space.
78 2. DIFFERENTIAL FORMS We shall prove B.29). By Proposition 1.39, it is enough to show that the operations of both sides on an arbitrary C°° function / € C°°(M) on M are equal. We shall calculate the operation of the right-hand side on /. Since, by A.14), ({<P-t).Y)f = Y(f O y,_t) O ypt = tf (y (/ O <p_t)), we have Um (jPzlhLzlf = iim y?(W°y-t)) ~ <PtiYf) + Vt(Yf) ~ Yf = Y(-Xf) + X(Yf) = [X,Y]f. Here we have used B.28), the fact that the functions which appear in the calculation are all of class C°° so that we can change the order of differentiation, and also the fact that {v-t} is the one-parameter group of local transformations generated by -X. Thus B.29) is proved. As for the Lie derivative of differential forms, the following proposition holds. PROPOSITION 2.13. Let X be a vector field on a C°° manifold M, and {(pt} the one-parameter group of local transformations generated by X. Then for an arbitrary k-form u € Ak(M), we have <p*tu - u Lyu = hm —= . t—o t PROOF. First, we shall show that if <p : M —¦ M is an arbitrary diffeomorphism, we have B.30) {<p*u))p(Xi,--- ,Xk) = <p*(u){(p+Xi,--- ,<p*Xk)) for vector fields X\, • • • ,Xk on M. By the definition of pullback of differential forms, we have {<p*u))p(Xi,'-> ,Xk) = ^(pj^.Xi,--- ,<p.Xk) for an arbitrary point p € M. B.30) immediately follows from this. If we calculate the right-hand side on X\y • • ¦ , Xk using B.30), we
2.2. VARIOUS OPERATIONS ON DIFFERENTIAL FORMS 79 obtain lim(y^)(*i>--- ,xk)-u>{xu-'- ,xk) *-o t = ]lm(ptM((Pt)*Xi1--- y{<pt)*Xk))-u>{Xu--- ,Xk) t-+o t = lirn y>?(^((y>t)*-X~i. — ,(<Pt)*xk))-<p*t{u{xu--- ,xk)) ?—0 t <pI{lj(Xi%--- ,Xk)) -u;(Xu--- }Xk) + am . t-*o t Let A be the first term and B the second term in this last formula. Then by B.28) we have B.31) B = X(j{Xi,--- ,Xk). On the other hand, we have = lim </?? ( ¦w{{<Pt)*Xi,--- y{ipt)*Xk) -uj(Xi,--- ,Xk)^ t + (u{kPt)*Xw- ,(y?f)>Xfc) t w(Xit{<pt)*X2,--- ,(y>t)*ATfc)\ + te^l } uj(X1,X2,(<pt).X3,---,(<Pt)*Xk)\ m fuj{Xi,X2,-'- ,Xk-i,((pt).Xk) -u>{Xi,--- ,Xk)\ k = Y,"(Xu---A-X,Xi}r--,Xk). t=i Therefore we have k A + B = Xuj(Xly-- ,**)-]? uf{Xw- ,[X,Xi\,---Xk) = (Lxuj)(Xu--.,Xk) and the proof is completed. I
2. DIFFERENTIAL FORMS 2.3. Frobenius theorem (a) Frobenius theorem— Representation by vector fields. Let M be a C°° manifold. If a vector field X is given on M, an integral curve through each point p 6 M is determined. Integral curves with the most extended domain (that is, the maximal integral curves) do not intersect each other and completely cover the whole of M (§ 1.4(c)). Then in the case where two vector fields are given on M, how does it go? Is it possible to construct, so to speak, "integral surfaces" by integrating them? How about the cases with 3, 4 or more vector fields? When we consider these questions, the following definition naturally arises. Definition 2.14. Let M be a C°° manifold. An r-dimensional distribution V on M is an assignment of an r-dimensional subspace Vp of TPM at each point p of M such that Vv is of class C°° with respect to p. Here Vv is said to be of class C°° with respect to p if there exist vector fields X\, • • • , Xr, defined in a neighborhood of each point p (of course, of class C°°) such that X\, ¦ • • , Xr become a basis of Vq at all points q in the neighborhood. A submanifold N of M is called an integral manifold of V if TpN = Vp for an arbitrary point p on N. If an integral manifold exists through each point of M, V is said to be completely integrable. As is easily seen by using the integral curves of a vector field, a 1-dimensional distribution is always completely integrable. To study higher dimensional distributions, we define a term. A vector field X on M is said to belong to V if Xp ? Vp for any point p e M. Proposition 2.15. Let V be a distribution on a ^°° manifold M. I/Vis completely integrable, then for any two vector fields X, Y belonging to V, the bracket [X, Y] also belongs to V. PROOF. It is enough to show that [X, Y)p e Vp at an arbitrary point p on M. By the assumption, there exists an integral manifold N through p. Let the dimensions of M and V be n and r respectively. Choose a local coordinate system (?/; ii, ¦ • • ,xn) around p such that p corresponds to the origin and the submanifold N is given by xr+i = • • = xn = 0. Then for an arbitrary point q on TV, Vq is the span of -—, • • • , 7t—. If the local expressions of X, Y with respect to this OXi OXr
2.3. FROBENIUS THEOREM 81 local coordinate system are then, since X, Y belong to ?>, we have B.32) di{xi,--- ,zr,0,--- ,0) = 6i(x1)--- ,xr,0, ••• ,0) =0 (i > r). It follows immediately that B.33) |^@) = ||@)=:0 (i<r,j>r). On the other hand, if the local expression of the bracket [X, Y] is [X,Y] = Yjcid~' then by A-1°) we have j j db3 daj\ -?Kt->€) and by B.32) and B.33) we see that Cj@) = 0 for arbitrary j > r. Therefore we have [X, Y]p € Vp, and the proof is finished. ¦ Based on Proposition 2.15, we make the following definition. Definition 2.16. A distribution V on a C°° manifold is said to be involutive, if the bracket [X, Y] of two arbitrary vector fields X, Y belonging to V also belongs to V. It is easy to see that if a distribution V on a C°° manifold M is involutive, then for an arbitrary open submanifold U, the restriction V\u of V to U is also involutive. Theorem 2.17 (Frobenius theorem). A necessary and sufficient condition for a distribution V on a C°° manifold to be completely integrable is that T> is involutive. This Frobenius theorem is very important for theory as well as applications. The proof of this theorem is given in subsection (c), after we study commutative vector fields in the next subsection.
82 2. DIFFERENTIAL FORMS (b) Commutative vector fields. Let M be a C°° manifold. Two vector fields X,Y on M are called mutually commutative vector fields if their bracket [X, Y) is 0. For example, on R2, — and — are commutative but — and ox oy ox x— are not commutative. Mutually commutative vector fields have oy geometrically good properties. Here is an example of it. Let {<?*} and {tpt} be the one parameter groups of local transformations generated by X and Y respectively. Then the following proposition holds. PROPOSITION 2.18. The following three conditions for two vector fields X,Y on a C°° manifold M are equivalent. (i) X and Y are commutative. That is, [X, Y] = 0. (ii) Y is invariant by <pt. That is, for an arbitrary t, (<Pt)*Y = Y holds whenever it is defined. (iii) ipt and rpt are mutually commutative. That is, for arbitrary t, s, the equality <pt o t/>s = t/>9 o <pt holds whenever they are defined. Proof. First we shall show that (ii) follows from (i). Considering t as a variable, if we differentiate a family of vector fields (ipt)*Y on M at t = to, we have ?((w).n| .u (<*.*).y-(«*.).*• dt \t-t0 t—o t .. . . (pt).y - y = hm(vO* -t = (v>to)*[-*,n = o. Here the last equality follows from B.29). This shows that {ft)*Y does not depend on t, and hence (<ft).Y = (<po)»Y = Y. Next we prove that (iii) follows from (ii). In general, given a (local) diffeomorphism <p of M, the vector field (p+Y of Y transformed by ip is defined (§ 1.4(d)), and it is easy to see that the integral curve of (p+Y through a point p on M has the form y?oc, where c is the integral curve of Y through tp~l(p). Hence, we see that the one parameter group of local transformations generated by tp*Y is {(p o rpt o y?-1}. If we apply the above fact to (tpt)*Y for each t, the one parameter group of local transformations generated by the vector field (<pt)* Y is given by <Pt°ips0(P7l w*tn 5 as a parameter. Therefore, if we assume (ii), we have <Pt°ips° fT1 = ^s* and hence tpt°ips = i>s°<Pt-
2.3. FROBENIUS THEOREM 83 Finally, we prove that (i) follows from (iii). By assumption we have <pt o ips o <p~l = tps. For each point p on M, since rps{p) is the integral curve of Y through p, clearly we have 4:Mp)\ =Yp. as ls=o On the other hand, when t is fixed, since <pt°i>s° ft^^iP) ls tne integral curve of the vector field {<ft)*Y through p, we have as U=o Therefore, if we assume the condition (iii), we obtain (<pt),Y = Y. Then by B.29), we have [x.yi-ibn-fr-'^-^o, 1 J t^O t and the proof is finished. ¦ (c) Proof of the Frobenius theorem. We now prove the Frobenius Theorem 2.17, which claims that a distribution on a C°° manifold is completely integrable if and only if it is involutive. The only if part was already shown in Proposition 2.15. Here we show that if a distribution is involutive, it is completely integrable. PROOF. Let V be an r-dimensional distribution on an n-dimen- sional C°° manifold M, and assume that it is involutive. Then it is enough to construct an integral manifold through an arbitrary point p on M. If we choose a sufficiently small coordinate neighborhood U of p, we can take vector fields Y\, • • • , Yr that are linearly independent at each point on U and such that each Y{ belongs to V. Let the local expression of Yi with respect to coordinate functions xi, • • • ,xn defined over U be Then, since the V* (i = 1,- • • ,r) are linearly independent, by changing the order of the Xi if necessary, we may assume that <kt(ayfo)) ^0 (qeU).
84 2. DIFFERENTIAL FORMS Define functions bij on U by (M<?)) = (M<7))_1 (9 €17) and let r Xx = ^jYj (i = l,.--,r). 3 = 1 Then it has the form Here the c^- are functions on 17. X\, • • • , Xr are, clearly, linearly independent and form a basis of V at an arbitrary point on U. Since V is involutive by the assumption, there exist functions /jt on U so that we can write [XuXA-j^fkXt. fc=l On the other hand, [Xi, XA is a linear combination of - , • • , —— OXr+l oxn by B.34). Therefore, we have fk = 0 (k = 1, • • • , r), namely [Xi,Xj] = 0. This shows that the vector fields X\, ¦ • ¦ ,Xr are mutually commutative. Now, let {ip\} be the one parameter group of local transformations of U generated by Xi. By Proposition 2.18, we see that the <p\ commute. That is, for arbitrary t, s, H>\ ° "Pi = v{ ° f] (*»J = 1'--' >r)> whenever they are defined. Then let V be a sufficiently small open neighborhood of the origin of Rr, and define a map <p: V —* U C M by ?>(*1,--- >*r) =<fltl O-'-O^Cp). Obviously, y? is a C°° map, and if we consider the differential of <p at the origin of Rr, we have Since X\,--- ,Xr are linearly independent, v?» : ToRr —> TPM is an injection. Therefore, we may assume that y? : V —> M is an
2 3. FROBENIUS THEOREM 85 embedding, by taking a smaller V if necessary. Then, the image N = Imtp of if becomes a submanifold of M. We now prove that N is the integral manifold of P. It is obvious that TPN = T>p. It is enough to show that TqN = Vq also, at an arbitrary point q on N. By the definition of <p, we can write q = <p(tu- ¦¦ , tr) = (pltl o • • • o <prtr(p) for some (<i, • • • , tr) e V. Since the </?J commute, we can rewrite this as B.35) q = <p\. o^c-o c^-i o $+1 o ¦ ¦ ¦ o <prtr(p) for arbitrary i = 1, • • • ,r (this is the most essential part of the whole discussion). In B.35), if we fix all tj but U, and vary U by a small amount, a curve on N through q is defined and is an orbit of <p\ through q\ that is, none other than an integral curve of X{. Therefore, we see that the velocity vector of this curve at q is Xi(q), and Xi(q) €TqN. Since this fact is valid for all i, we have TqN = Vq as a conclusion. That is, N is an integral manifold of Z>. Now, the existence of the integral manifold of V through an arbitrary point on M is shown, and the proof of Probenius theorem is finished. ¦ Although the proof is completed, we show that a stronger fact can be deduced by refining the last part of the above discussion. Assume first that the point p corresponds to the origin of a local coordinate system xi,-- ,xn, and let W = {q e U;xi(q) = • • • = xr{q) = 0}. Then W is an (n - r)-dimensional submanifold of U intersecting N transversally at p. We shall define a map (p : V x W —» M by ?(*i,--- ,tr,q) = ?>t\ ° '• ¦ °<PtM) (<?€ W)- Then we see that </? is an embedding in the same way as above and, for an arbitrary q € W, <p(V x q) c U is exactly the integral manifold through q (see Figure 2.1). In the case of an involutive distribution of general dimension, we can also obtain the notion of a maximal integral manifold by considering the union of all connected integral manifolds through a point, just in the same way as in the case of integral curves. Furthermore we can prove that a natural C°° manifold structure is induced on any maximal integral manifold and the inclusion of it in the original manifold is a one to one immersion. (We omit the details, because it is not very difficult. The interested reader is encouraged to try
86 2. DIFFERENTIAL FORMS FIGURE 2.1. Integral manifold of an involutive distribution it.) However this map is not an embedding in general, so a maximal integral manifold is not always a submanifold (in the definition of this book). (Some authors define an embedding and a submanifold to be a one to one immersion and its image; those definitions are weaker than the ones adopted in this book.) (d) The Frobenius theorem Representation by differential forms. Here we describe the Frobenius theorem in terms of differential forms instead of vector fields and distributions. These two representations of the Frobenius theorem are equivalent, and both are important. Suppose that an r-dimensional distribution V is given on an Tridimensional C°° manifold M. That is, at each point p € M, an r-dimensional subspace Vp of TPM is assigned. In order to give a presentation of V in terms of differential forms, we proceed as follows. For an arbitrary k > 1, let Ik{T>) = {ueAk{M); u{Xu--- ,XJfc) = 0forXieP} and let /(P) = 0/fc(P). fc=i I(V) is the set of all differential forms on M which "vanish on V". Lemma 2.19. (i) I(V) is an ideal of A*{M). That is, 1(D) is a linear subspace ofA*(M), and we have 9 A u> e I(V) for arbitrary 9 e A*{M) andu> e I{V).
2.3. FROBENIUS THEOREM 87 (ii) Locally 1(D) is generated by s = n — r linearly independent 1-forms. That is, for an arbitrary point p e M, there exist an open neighborhood U of p and 1 -forms u>\, • • • , ujs that are linearly independent at each point on U such that an arbitrary u> € 1(D) can be written as u) = y_] Ot Aoji. t=i Here the $i are differential forms onU. In this case we clearly have Vq = {Xe TqM- u>x(X) = • • • = u>s{X) = 0} for an arbitrary point q € U. Proof. Obviously (i) follows from the property B.14) of the exterior product. We shall prove (ii). If we choose a sufficiently small open neighborhood U of p, there exist vector fields Xs+i, • • • ,Xn that are linearly independent at each point on U, such that V is spanned by the Xi (i = s + 1, • • • , n) on U. We add vector fields X\, • • • , Xs on U so that X\, • • • , Xn become a basis of the tangent space at each point. Let u\y • • • ,u>n be 1-forms dual to X\, ¦ ¦ • , Xn. That is, they satisfy Ui(Xj) = 6{j. Then an arbitrary fc-form u> on U can be described uniquely as a linear combination of /c-forms of the form cJi, A • • • A ujlk (ii < •¦• < ifc), with functions as its coefficient. Now such an u belongs to Ik{T>) if and only if all coefficients in the above description corresponding to i\, • • • , ik which are all different from 1, • • • , s, become 0. In other words, this is nothing other than that w belongs to the ideal generated by a>i, • • • , u>s. This is what we needed to prove. ¦ Proposition 2.20. Let V be a distribution on a C°° manifold M, and 1(D) the ideal of A*(M) consisting of all differential forms that vanish on V. Then a necessary and sufficient condition for V to be involutive is that 1(D) becomes a differential ideal in the sense that it is closed with respect to the operation of exterior differentiation: dI(D) C 1(D). PROOF. First we shall show that the condition is necessary. Let w e Ik(D) be an arbitrary element. Then, for any vector fields Xi, • • • > ^Jt+i belonging to ?\ since V is involutive, [Xi, Xj) also belongs to V. Then by Theorem 2.9, we have dw(A"i,¦ • • ,Xk+i) = 0.
88 2. DIFFERENTIAL FORMS Therefore du € Ik+l(T>), and we see that 1(D) is closed with respect to the exterior differentiation; that is, it becomes a differential ideal. Next we assume that I{V) is a differential ideal, and we show that for arbitrary vector fields X, Y belonging to V, [X, Y] also belongs to V. For this, it is enough to show that u>([X, Y)) = 0 for an arbitrary element u e Il{V). By assumption we have du{X, Y) — 0, and again by Theorem 2.9 we have MX,Y) = \{X(w(Y)) - Y{u(X)) -u,([X,F])}. Since w(X) = v(Y) = 0 here, we obtain uj([X, Y]) = 0, and the proof is finished. ¦ We shall write the above result more practically. An arbitrary r-dimensional distribution on an n-dimensional manifold M is represented locally by equations B.36) u>x = • • • = u>s = 0, where u>i, • • • ,u>s are s = n — r 1-forms that are linearly independent on a neighborhood U of a given point p e M. That is, we have Vq = {X e TqM;u)i(X) = • • • = ujg(X) = 0} at an arbitrary point q € U. B.36) is sometimes called a system of Pfafflan equations. Then we can interpret Proposition 2.20 another way. Namely, a necessary and sufficient condition for V to be involutive (on U) is that there exist 1-forms u>ij on U such that 5 B.37) du>i = ^,^3 Awi (i = 1,-*- .«)- This condition B.37) is called the integrability condition. From the above facts, we can formulate the Frobenius theorem in terms of differential forms as follows. Theorem 2.21 (Frobenius theorem). A necessary and sufficient condition for a distribution V on a C°° manifold M to be completely integrable is that if we represent T> as Vq = {X e TqM-^x{X) = ¦-.= us{X) = 0} by linearly independent l-forms u>\,- ¦ ¦ ,a>s on an open neighborhood of each point on M, they satisfy the integrability condition B.37).
2.4. A FEW FACTS 2.4. A few facts (a) Differential forms with values in a vector space. Let M be a C°° manifold. Then a fc-form w 6 Ak{M) on M assigns to each point peMan element u>p in hkT*Mt that is, an alternating multilinear map B.38) ujp : TPM x • • • x TVM —» R which varies differentially with respect to p. Generalizing this, for a vector space V, we can define a V-valued fc-form on M by replacing R in B.38) by V. We denote the vector space of all V-valued fc-forms on M by Ak{M;V). When we choose a basis vi, • • • ,vr of V, an arbitrary element u> € Ak{M; V) is expressed as ijj = \_] tc'it;* i=l in terms of usual fc-forms u\, • • • , ur. The exterior differentiation of u>, dw € ^4fc+1 (M; V), is defined in a natural way as du> = Y~] du;^. For a V-valued fc-form u> € Ak(M; V) on M, and an /-form 77 e Al{M; W) with values in another vector space W', their exterior product wAtN >-+'(M; V ® W) is defined by wAr?^,-- ,Xfc+j) = (fc.nj ]C sgnau;(X<T(l)»--- >XG(k))^77(Xa(k+1))-- ,X<r(k+1)) (see B.14)). If we express 77 as 77 = /JfyWj with respect to a basis ; w\, • • • , u>s of W, then we have u> A 77 = yj ^t A ^j vt ® wj ¦ Then, by m Proposition 2.3 and its generalization, it is easy to see that B.39) d{u> A 77) = du A 77 + (- l)fcw A cty. Next assume that a bilinear map V x V —* V is given on V and V is a Lie algebra with respect to this. The most important case is the one where V is the Lie algebra g of a Lie group G; this case will appear in the next subsection. Then for u> ? Ak{M;V) and 77 e Al{M\ V)} their product {^,77] e Ak+l(M; V) is defined by the composite Ak{M; V) x Al{M; V) —¦ Ak+l(M\ V <g> V) —¦ ^+'(M; V).
90 2. DIFFERENTIAL FORMS Here the first map is the above defined exterior product and the second map is induced by the bracket V <g> V —> V of the Lie algebra V. If we express u> = NJ^t^i? V — z_]'njvj w*th respect to a basis * 3 v\ i • • • , vr of V, we have [cj, -q] = Y^ tJi A rjj [vi, Vj]. Therefore we have »,> h,u;] = (-l)w+1K»?]. Practically, for example, for o> G A1{M; V) we have B.40) ku;](Xf y) = ^{MX),u;(F)] - [W(y),"P0]} = MX),^)] (X, y € X(M)). Also, as for the exterior differentiation, we see from B.39) that B.41) d[w,rj\ = \dw,rj\ + (-l)*^,^]. Moreover, by using the Jacobi identity of the Lie algebra V, it is easy to see that B.42) [[u>,u>],u;]=0, for an arbitrary u> e A1{M; V). (b) The Maurer-Cartan form of a Lie group. Let G be a Lie group. Readers not familiar with the general theory of Lie groups may assume it to be a matrix group such as G = GL(n;R), GL(n;C), 0(n), C/(n), etc. ft is only these cases that are used in this book. The tangent space TeG at the identity e of G is called the Lie algebra of G and usually denoted by the corresponding German letter 0. For an element g of G, we shall denote the left action of g by Lg : G —» G. That is, it is a map defined by Lg(h) = gh (h € G). An arbitrary element X € 0 can be considered to be a vector field on G by putting X{g) = (Lg)*X. Here (Lg). : TeG — TgG is the differential of Lg at e. The vector field X obtained in this way is left-invariant, that is, (Lg)*X = X for an arbitrary g € G. Since it is obvious that all left-invariant vector fields on G are obtained in this way, 0 can be considered as the set of all left-invariant vector fields on G. For X, Y e 0, their bracket [X,Y] belongs to 0, since it is also left-invariant. Equipped with this product, 0 becomes a Lie algebra.
2.4. A FEW FACTS 91 Example 2.22. In the case of G = GL(n;R), its Lie algebra gl(n;R) can be identified with the set M(n;R) of all real square matrices of order n in a natural way. Practically, for any X € M(n;R), exptX (t € R) is a C°° curve through the identity of GL(n;R), and so we can associate its velocity vector. Then we see that the bracket is given by \X}Y) = XY - YX. It is known that we have, similarly, gl(n;C) = M(n;C) (the set of all complex square matrices of order n), o(n) = {X e 0l(n;R); X + lX = 0} (the set of all alternating matrices of order n), u(n) = {Xe gl(n;C);X + <X = 0} (the set of all skew Her- mitian matrices of order n). Let G be an m-dimensional Lie group and B\, • • • , Bm a basis of g. Then, since the bracket [Bi,Bj] is also left-invariant, we can describe it uniquely as B-43) (Bi.B,-] = ?<$?*. k The constants c^ are collectively called the structure constant of the Lie algebra g with respect to the above basis. Example 2.23. As a basis of g((n;R), we can take all matrices XXj {i,j = l,--- ,n) such that its (i,jf)-entry is 1 and the others are all 0. The structure constant for this basis is immediately obtained from X) X* = Sjk X\. A differential form w on G is called a left-invariant differential form, if L*g u> = u> for an arbitrary g G G. It is obvious that the left-invariant differential forms are determined only by the value at the identity e. Now we can consider an arbitrary element u> in the dual space g* of the Lie algebra g as a left-invariant 1-form on G. Practically, it is enough to set u>(X) = u>{{L~l).X) for X € TgG. Moreover, it might be obvious that they exhaust all the left-invariant 1-forms on G. That is, we have 0* = the set of all left-invariant 1-forms on G. A left-invariant 1-form on G is called a Maurer-Cartan form. Given u € g*, since u>(X)>lj(Y) are constant functions on G for arbitrary XyY e g, we have Y{w(X)) = X{u>{Y)) = 0. Therefore, by
92 2. DIFFERENTIAL FORMS Theorem 2.9 we obtain B.44) MX,Y) = -±u>([X,Y]). Now let u)\, • • • , ujm be the dual basis (to the above basis of g) of g*. Then by comparing B.43) and B.44), we obtain the following Maurer-Cartan equation: B.45) <^i = -^ c)k ujj A u)k. 1 3,k The above facts can be described more simply by using g-valued differential forms as follows. Let u € A1(G;q) be a g-valued 1-form on G such that w(A) = A for A e g. Using the above basis, it is described as This w is also called a Maurer-Cartan form. Then the Maurer-Cartan equation has the following form: B.46) dLj = --[w,w]. Here, [u)}u>] denotes, as in B.40), a g-valued 2-form defined by [u,uj)(X>Y) = [«>(X)MY)\ for arbitrary vector fields X, Y. Example 2.24. We shall find the Maurer-Cartan equation of GL(n; R). We denote by u) the dual basis of the basis {X) } of Ql(n\ R) given in Example 2.23. Then, by the structure constants determined there, we immediately see that the required equation is k Summary 2.1 A /c-form on a C°° manifold is an assignment at each point of an alternating form from the A:-fold direct product of its tangent space to R in such a way that it varies in a C°° manner with respect to the points. 2.2 In other words, a /c-form on a C°° manifold M is a multilinear and alternating map from the fc-fold direct product of the set X{M) of all vector fields on M to C°°{M) as C°°(M)-modules.
EXERCISES 93 2.3 The exterior differentiation d is a linear map of degree 1 operating on all differential forms, and it satisfies d o d = 0. 2.4 A differential form that vanishes under the exterior differentiation is called a closed form, and a differential form that can be expressed as an exterior differential of some other differential form is called an exact form. 2.5 A vector field on a C°° manifold operates on differential forms by the interior product as well as the Lie derivative. The interior product is an anti-derivation of degree -1, and the Lie derivative is a derivation of degree 1. 2.6 A distribution on a C°° manifold is an assignment at each point of a subspace of its tangent space with a fixed dimension in such a way that it varies in a C°° manner with respect to the points. 2.7 A necessary and sufficient condition for a distribution on a manifold to be completely integrable is that it be involutive, that is, the bracket of any two vector fields belonging to V always belongs to V. This is called the Frobenius theorem. Exercises 2.1 Verify directly that, for u; G Ak{Rn), rj e Al{Rn), A) T)Auj = {-l)kl UJAT), B) d(u A 77) = dw A 77 4- (-\)ku A drj. 2.2 Let U, U' be open sets of Rn, and let tp : U —> U' be a diffeo- morphism. Then, verify directly that d(tp*u>) = <p*(duj) for an arbitrary <*/ € Ak{U'). 2.3 Prove Proposition 2.12(i),(ii) using the Cartan formula (Theorem 2.11). 2.4 Define a 2-form on R2n by u = dx\ A dx2 + dx^ A dx^ 4- • • • + dx2n-\ A dx2n- This is called the standard symplectic form on R2n. Then compute u>n. 2.5 Let Nbea closed submanifold of a C°° manifold M, and let i : N —¦ M be the inclusion map. Then show that the map ¦i* : A*{M) —» A*(N) induced by i is a surjection. 2.6 Let / : M —* N be a submersion (see Definition 1.36). Then show that the map /* : A*(N) —» A*{M) induced by / is an injection.
94 2. DIFFERENTIAL FORMS 2.7 Define an (n - l)-form u) on the space Rn - {0}, obtained from Rn by removing the origin, by u> = X^(-1)t-1 xi ^1 A ¦ • ¦ A do:* A • • • A (ix„ \\x\\ i=1 Then prove that dw — 0. 2.8 Let X be a vector field on a C°° manifold M, and let {<?><} be the one-parameter group of local transformations generated by X. A necessary and sufficient condition for a differential form u> € A*(M) on M to be invariant under {<pt} (that is, (p*tu — u for all t) is that Lx^> = 0. Prove this fact. 2.9 Let (r,0) be the polar coordinates defined on R2 outside of the origin. Then describe the 1-forms dr, d9 in terms of the ordinary coordinates x,y. 2.10 Find the Maurer-Cartan equation of the 3-dimensional Lie group SU{2) = {Ae UB)- detA = 1}.
CHAPTER 3 The de Rham Theorem We have a quantity called the homology group which can express global structure of figures. It was created by Poincare, and at present, about one hundred years later, it is a completely established theory. Roughly speaking, the homology group measures the essential'number of "holes", called cycles, in a given figure of each dimension. It is an important problem in the study of manifolds to determine their homology groups. The de Rham theorem guarantees that the homology groups of difFerentiable manifolds can be "detected" by differential forms. More precisely, by integrating closed forms on cycles, we can investigate their non-triviality and the relationships among them. Furthermore, we can also describe homology within the framework of differential forms. The description of the homology (or its dual, co- homology) of manifolds in terms of differential forms opened a way to study the deeper structure of manifolds by analytical methods, and we cannot overestimate its influence on the later development of the theory of manifolds. The de Rham theorem was proved by de Rham, as its name indicates. Among differential forms, the history of degree one differential forms is old, and, for instance in Abelian integrals, the integral values on cycles (namely periods) were already considered at the beginning of the nineteenth century. However it may be said that the global study of general differential forms of higher degrees was begun in E. Cartan's extensive research, which started in the 1920's. E. Cartan conjectured the de Rham theorem in a 1928 paper, and it was not long before de Rham gave a proof. As cohomology theory developed, more sophisticated proofs have been published, and among them the proof by Weil is elegant and well compatible with the sheaf cohomology theory developed thereafter. In this book, while we use Weil's method, we shall give a proof whose geometric meaning will be easy to grasp, at the cost of slightly decreasing its theoretical beauty. 95
96 3 THE DE RHAM THEOREM 3.1. Homology of manifolds (a) Homology of simplicial complexes. Here we briefly summarize homology theory, which is fundamental in studying the global structure of figures, and we also fix notations. Let X be a topological space. Roughly speaking, the /-dimensional homology group Hi(X) of X measures the essential quantity of structures called "/-dimensional cycles" in X. There are several methods of defining the homology, depending on the shape of X. The homology theory of simplicial complexes is the method that is historically the oldest, and intuitively easy to understand. In this method, we decompose X into points, line-segments, triangles, • • •, and in general, fundamental items called /-simplices (such a decomposition is called a triangulation), and define homology using their combinatorial structure. If we generalize simplices to cells for fundamental items (thus admitting n-polygons rather than simply triangles), then we obtain the homology theory of cell complexes. On the other hand, if we use singular simplices, which are, so to speak, the ultimate generalization of the usual simplices, then we obtain the singular homology theory. Though each of these homology theories has its own merits as well as demerits, the important thing is that they are all equivalent for reasonable spaces (including differentiable manifolds). Although it is possible to define simplicial complexes in an abstract fashion, here we consider them in RN for a sufficiently large N. I + 1 points t>o, vi- • ¦ • . v/ in RN are said to be in general position if the vectors v\ - vq,V2 — vq, ¦ ¦ • , vi — vo are linearly independent. For a set a = {vo,fi- • • • ,vi] of / + 1 points in general position, the smallest convex set |a| = {a0VQ H h aivr, a» > 0, a0 H h ai = 1} including those points is called an /-simplex. It is also written \cr\ = l^o^i •••vi\- Each Vi is called a vertex of the simplex, and I is called its dimension. For I = 0,1,2 and 3, an /-simplex is a point, a line segment, a triangle and a tetrahedron respectively. For an arbitrary non-empty subset r of the set a of vertices of an /-simplex |<r|, \r\ is also a simplex, and such a simplex is called a face of |<j|. Definition 3.1. A set K of simplices in RN is called a Euclidean simplicial complex if it satisfies the following conditions, (i) If \a\ G K, then an arbitrary face of \a\ belongs to K.
3.1. HOMOLOGY OF MANIFOLDS 97 (ii) If two simplices |<r|, \r\ € K intersect, then their intersection |<j| H \t\ ^ 0 is a common face of \<j\ and \r\. (iii) For an arbitrary point x on an arbitrary simplex \a\ in K, we can choose an open neighborhood U of x such that there are only a finite number of simplices in K that intersect U. For a simplicial complex K, the union of all simplices belonging to K is denoted by \K\. A subset of RN obtained in this way is called a polyhedron. For a topological space X, if we can choose an appropriate simplicial complex K and a homeomorphism t : \K\ —> X, it is called a triangulation of X. If we extract only combinatorial structure from the above Euclidean simplicial complex, we obtain the notion of abstract simplicial complex. That is, given a set V of elements called vertices, a subset K of the power set 2v (the set of all subsets of V) is called an abstract simplicial complex if the following two conditions are satisfied: (i) for all v € V, {v} G K, and 0#K, (ii) if a e K then for all t c o such that t ^= 0, t e K. A Euclidean simplicial complex is obviously an abstract simplicial complex. Conversely, for instance if V is a finite set, we can prove that any abstract simplicial complex with V as the set of vertices can be realized as a Euclidean simplicial complex. Hereafter, Euclidean simplicial complexes or abstract simplicial complexes are simply called simplicial complexes. For a simplicial complex K, its homology group H+{K) is defined as follows. To begin with, we consider an ordering of the vertices fo>Vi, • • • ,vi of each /-simplex |a| = |vot;i • • • v/|. Two orderings are said to be equivalent if they can be transformed into each other by even permutations. An equivalence class of orderings of vertices is called an orientation of that simplex. If / > 1, there are exactly two orientations in each Z-simplex; we call them opposite orientations. A simplex with a specified orientation is called an oriented simplex, and we write it as (a). If the vertices are ordered as v^i;^, • • • , Vit, the corresponding oriented simplex is.denoted by {viovil ¦¦¦vil). We now assign an orientation to each /-simplex \ai\ of a simplicial complex K, and write it as (<7t). We denote the free abelian group generated by {&{) by Ci(K), and call any element of this group an /-chain of K. We denote |<7i| with the opposite orientation by -(cr,). A homomorphism d'.Ci{K)-^Ci^{K)
98 3. THE DE RHAM THEOREM called the boundary operator is defined by setting i d(v0vi ¦¦¦vl) - Y2 {-iy(v0---Vi---vi) i=0 on each oriented /-simplex and extending it linearly. Here Vi means that V{ is omitted. The important thing here is that if we apply the boundary operators twice, the value is always 0; that is, d o d = 0. This may be phrased as "a boundary has no boundary", and this is the starting point where Poincare created the notion of homology. By virtue of this fact, if we let Zi{K) = {ceCi{K)\ dc = 0}, Bl(K) = {dc;ceCl+l(K)}, then Bi{K) C ZL{K). We denote the quotient group Zt(K) /Bi{K) by Hi{K) and call it the /-dimensional homology group of K. Any element of Z\{K) is called an /-dimensional cycle, and any element of Bi{K) is called a boundary cycle of K. The homology class represented by a cycle z 6 Zi(K) is usually denoted by \z\ € Hi(K). Also the two cycles z, z' € Zi(K) are called homologous if they represent the same homology class, in other words, if there exists a chain c € C\+\{K) such that z' — z = dc. The above fact can be stated simply as follows: the homology group Hm(K) of K is the homology group of the chain complex C.(K) = {a(K),d}. The above homology group H,(K) refers to the integral homology group, and when we want to emphasize this fact it is denoted by H*(K\Z). The homology group with coefficients in a general abelian group A is defined as the homology group of the chain complex C+(K) <8> A and is denoted by H*(K; A). If L is a subcomplex of K, the relative homology group Ht(K,L;A) is defined as the homology group of the chain complex C.(K) <g> A/C.(L) <8> A. Next we give a brief description of the cohomology. In a single word, cohomology is the dual of homology. If a chain complex C* = {Ci,d} is given, its dual cochain complex C* = {Cl,6} (although it is often written Hom(C»,Z), here we simply denote it by C*) is defined as follows. First, let C' = Hom(C/,Z) be the set of all homomorphisms from Ci to Z; the dual boundary operator is defined by setting Sf(c) = f{dc), c € C\+\ for / € Cl =
3.1. HOMOLOGY OF MANIFOLDS 99 Hom(C{, Z). It is easy to see that 6 o S = 0. If we let Zl(C*) = {feCl;6f = 0), Bl(C*) = {6f]feC1-1}, we see that Bl(C*) C Zl(C*). Then we denote the quotient group Zl{C*)/Bl(C*) by Hl{C*) and call it the /-dimensional cohomology group of C*. (In the case of C. = C.{K) as above, Hl{C*{K)) is called the /-dimensional cohomology group of the simplicial complex K.) Any element of Zl(C*) and Bl(C*) are called an /-dimensional cocycle and coboundary of C* respectively. The cohomology class represented by a cocycle / € Zl(C*) is usually denoted by [/] 6 Hl{C*). Also two cocycles /, /' € Z[{C*) are called cohomologous if they represent the same cohomology class - in other words, if there exists a cochain g € Cl~l such that f — f = 6g. Between the homology group H» (C.) of a chain complex C+ and the cohomology group H*(C*) of its dual cochain complex C*, there exists a bilinear map Hi{C,)®Hl{C*)-*Zy called the Kronecker product, which is defined by ([z], [/]) h-> f(z). The Kronecker product shows that cohomology classes have the function of detecting homology classes. (b) Singular homology. We now recall the singular homology, which is defined for general topological spaces. We consider Ak = {rr = (xi,--- ,xfc) € Rk\ x4 > 0, xx + • ¦ • + xk < 1} as a representative of all fc-simplices. This is called a standard k- simplex. (Although in the previous subsection we used subscript / Figure 3.1. A3
100 3. THE DE RHAM THEOREM to avoid confusion with the symbol K of simplicial complex, hereafter we use k to match the subscript of differential forms.) For a topological space X, an arbitrary continuous map a : Afc -> X is called a singular fc-simplex of X. The free abelian group generated by all singular fc-simplices of X is denoted by Sk(X), and an element of it is called a singular A:-chain. For i = 0,1, • • • ,k, we define continuous maps e* : Afc_1 —> A* by fc-i C.1) €0(X!,--- ,Ifc_i) = (l - ^Xi.Xi,--- .Xfc-ij, t=i C.2) e*(*i»--- ,Xk-i) = {x\y--- ,xt_i,0)xi,--- ,Xfc_i) (i = l,--- ,fc). Using these, the boundary operator 0:Sfc(X)-Sfc_i(X) is defined by da — /"](—1)* ° ° e*- Here a^so we see tnat ^ ° ^ = 0, t=0 and hence Sm{X) = {Sk{X),d} is a chain complex. This is called the singular chain complex of X. Its homology group is denoted by H*(X) and called the singular homology group of X. In the case where X is a polyhedron \K\ of a simplicial complex K, it is known that the simplicial homology group H* (K) of K and the singular homology group H»(\K\) of X = \K\ are naturally isomorphic. In particular, the simplicial homology group is topologically invariant. That is, homology groups of polyhedra are determined independently of the choice of triangulations. Definitions of the singular homology group H*(X; A) with coefficients in a general abelian group A and the relative homology group H* {X, Y; A) for a subspace Y of X can be given similarly as in the case of simplicial complexes. Also, the cochain complex Hom(S*(X), Z) is usually denoted by S*(X) and its cohomology group H*(S*(X)) is called the singular cohomology group of X. (c) C°° triangulation of C°° manifolds. Generally in the study of given figures, it is often convenient to consider triangulations of them. However, it is not an easy problem to decide whether a given figure is triangulable or not, and in fact the problem of existence and uniqueness (in the combinatorial sense)
3.1. HOMOLOGY OF MANIFOLDS 101 of triangulations of topological manifolds was one of major themes in the 1960's, in the field called topology. In the case of differentiable manifolds, existence was known early (in the 1930's). Definition 3.2. Let M be an n-dimensional C°° manifold. A triangulation t : \K\ —> M of M by an n-dimensional simplicial complex K is called a C°° triangulation, if, for an arbitrary n-simplex |a| of K, the restriction t\\g\ o( t to \o\ is a. C°° embedding. Here t\\a\ is a C°° embedding if it can be extended to a C°° embedding from an open neighborhood U of |a| in the n-dimensional subspace spanned by \cr\ (in RN where \K\ is contained) to M. THEOREM 3.3 (Cairns, J.H.C.Whitehead). Any C°° manifold has a C°° triangulation, and any triangulation of the boundary of a C°° manifold can be extended to a triangulation of the whole manifold. We cannot give the proof of this theorem here, since it is technically rather complicated. However, the relationship between quantities combinatorially denned using triangulations of C°° manifolds and those defined by making essential use of differentiable structures such as integrating differential forms has been one of the main themes of the geometry of manifolds until now. Recent progress seems to indicate that the importance of the combinatorial viewpoint using triangulations will increase in the future. Using the above theorem, we can prove an important fact concerning the homology of compact C°° manifolds. Let M be an n-dimensional closed C°° manifold that is connected and oriented. Let t : \K\ —¦ M be a C°° triangulation. Now we shall see that on each n-simplex |cr* | of K, an orientation is induced from the orientation of M. Let the vertices of |<7j| be vo, ¦ ¦ • , vn and denote the point t(vo) on M by po- For i — 1, • • • , n, let the unit vector from vQ toward Vi be Ui. We can consider u* a tangent vector to |o-t| at vq. Then, if we let Wi = t»(ul), w\, • • • ,wn form an ordered basis of TPoM. Here, if necessary, by exchanging vo and v\ for instance, we can make the orientation determined by this ordered basis coincide with the orientation (at po) of M. We now give |<7j| an orientation determined by the ordering vq, vi, • • • , vn of vertices and denote the oriented n-simplex obtained in this way by (<7t) (Figure 3.2). Now, we define an n-dimensional chain Co € Cn{K) of K by co = X>>-
102 3. THE DE RHAM THEOREM Figure 3.2 Then we can see that cq is a cycle (that is, 8cq — 0), in the following manner. 6cq = 2J dfai) 1S a linear combination of appropriately oriented (n - l)-dimensional simplices of K. Now let |r| be an arbitrary (n — l)-simplex of K. By the definition of C°° triangulation, we see that there exist exactly two n-simplices of K with \r\ as their face. Let them be \at\ and |<7j|. In the boundaries d{<Ji), d{aj) of the oriented n-simplices (&{), (oj), a term of |r| with an orientation appears. As we see in Figure 3.3, where the case of n = 2 is illustrated, it is easy to check that these orientations are in fact opposite. Therefore, dc0 = 0. Figure 3.3 Next, let c= yja,(Gt) (a* € Z) be an arbitrary n-dimensional cycle of K. Then, by a similar argument as above showing dc = Yjaid(<7j) = 0, we see that each coefficient a» has to be independent of t, so it is a constant. That is, it has the form c = aco- On the other hand, since K is an n-dimensional simplicial complex, its
3.1. HOMOLOGY OF MANIFOLDS 103 n-dimensional boundary is trivial. In this way, we see that the n- dimensional homology group Hn{K; Z) of K is an infinite cyclic group generated by the homology class represented by Cq. H^{K\Z) and H*(M\Z) can be naturally identified, and the homology class represented by c0 is denoted by [M] e H*(M;Z). This is because the topological invariance of the homology group and the definition of cq imply that [M] is determined independently of the choice of C°° triangulations. We call [M] the fundamental class of M. If we reverse the orientation of M, the fundamental class changes its sign and we have [-M] = —\M). We summarize the above facts as follows. THEOREM 3.4. Let M be an n-dimensional closed C°° manifold, and assume that it is connected and orientable. Then tfn(M;Z)^Z. Furthermore, the fundamental class \M] which is. determined if we specify an orientation on M is a generator of this group. This theorem says that an arbitrary orientable closed manifold is itself a cycle. Although it is surely a simple fact, this is an origin of the notion of homology, and what it means is indispensable. Actually, in the 1950's, Thorn proved that a certain multiple of any homology class of an arbitrary figure can be realized as the "image" of the fundamental class of an oriented closed manifold. (d) C°° singular chain complexes of C°° manifolds. Let M be a C°° manifold. Since M is also a topological space, its singular chain complex S*(M) = {Sk{M),d} is defined. However, this is inconvenient for the consideration of integration of differential forms on cycles. Thus the necessity of C°° singular chains comes out, and that will be introduced below. A C°° map a : Ak — M from the standard fc-simplex Ak to M is called a C°° singular k- simplex of M. Here a- is a C°° map if it can be extended to a C°° map from an open neighborhood of Ak in Rk to M. The free abelian group generated by all C°° singular fc-simplices of M is denoted by S^(M), and an element of it is called a C°° singular fc-chain of M. For an arbitrary c 6 S^{M), we see that dc € SgL^M). Therefore, Sr(M) = {S?(M),d}
104 3 THE DE RHAM THEOREM is a subcomplex of the singular chain complex 5»(M) of M. We call Sf{M) the C°° singular chain complex of M. Also we denote its dual complex HomEf>(M),R) by S^{M) and call this the C°° singular cochain complex of M with coefficients in R. An important fact here is that the inclusion map S^°(M) C S*(M) induces a natural isomorphism ^E~(M))Si/,E.(M)). Although we do not prove this fact here, we point out that an argument similar to that in the proof of the de Rham theorem (§3.4) can be utilized. In this way, in the case of C°° manifolds, the homology group can be discussed using only C°° singular chains. 3.2. Integral of differential forms and the Stokes theorem (a) Integral of n-forms on n-dimensional manifolds. We wrote, at the beginning of Chapter 2, that differential forms should be integrated on manifolds. We shall clarify that statement in this section. For a C°° function f(x) on Rn with compact support, its Riemann integral C.3) / f(x)dXl---dxn= limn ?/(*,) k,| is defined. Here the Oj are n-dimensional small cubes which altogether cover supp/, Xj is a point on Oj, and \<jj\ is the volume of &j. Usually in such a formula, the function f(x) to be integrated is the main part and dx\ ¦ • -dxn sitting behind it just means the n-dimensional Riemann integral, or in the computation of it by iterated integral, it also symbolizes the n-fold multiple integral v'-oo J -c f(x) dx\ ¦ ¦ dxn However, here we change the viewpoint slightly and consider them as a whole (putting in the symbols A also), and set C.4) iA) = f(x) dxi A--- Adxn. Then this is nothing but an n-form on Rn. We may consider that the integral / u> of u> on Rn is defined by C.3). However, two problems arise. One is that if we change the order of dx\ and dx2, for example, in the formula of a;, the sign changes, while in C.3) the integral does
32. INTEGRAL OF DIFFERENTIAL FORMS; STOKES THEOREM 105 not change. Another one is that as long as we write / us, its value must be a quantity independent of the choice of coordinates of Rn. Is this true? Let us verify these points. Let yi,-•• ,yn be another coordinate system of Rn. Strictly speaking, we are given a C°° diffeomorphism <p : Rn —¦ Rn from Rn with coordinates y\, • • • >yn to Rn with the former coordinates X\, ¦ ¦ ¦ ,xn. However, we identify two Rn,s by </? to simplify the notation. Then, the former coordinate Xj can be expressed as a C°° function of t/i, • • • , yn. That is, we can write Xi =Xi{yu-- ,yn) (i = l,--- ,rt). We put these together and write x = x{y). The Jacobian matrix of dxi this coordinate change is denoted simply by ( -^- ). Then, if we express u> with respect to new coordinates yi,--- , yn, by B.4) we have C.5) w = /(x(y)) det ( -^ )dyx A • • • A dyn. On the other hand, by the formula of variable change, we have C.6) / f(x)dxl--dxn= f f(x(y))\det{p^)\dyi--dyn. ./Rn jRn Oyj dx Here the term I det ( 7—^- ) I appears, because if we transform a small dyj area by the coordinate change v?, the volume changes approximately by that ratio. Comparing the four formulae C.3),C.4),C.5) and C.6), we see that, if we consider only those coordinate changes the determinant of whose Jacobian matrices (Jacobian) are positive, then the above two problems are solved simultaneously. If we recall here the definition of the orientation of manifolds (§1.5(b)), we reach the following conclusion: If we specify an orientation on Rn, the integral of u> on Rn is determined uniquely and independently of the choice of coordinates. Also, in the above argument, we assumed that the coordinate change v? is defined on the entire Rn, but in fact we do not need to do so and it is enough to define it on an open set U containing supp/. We leave the verification of this fact to the reader as an exercise. Now the preparation is completed, we define the integral of differential forms on a manifold M. In general, for a differential form u>
106 3. THE DE RHAM THEOREM on M, the set suppu; = {p€M;wp/ 0} is called the support of u>. That is, suppu; is the smallest closed set such that u; is 0 outside of it. Now let M be an oriented n-dimensional C°° manifold, and u an n-form on M with compact support. Then the integral of w on M, /"• Jm is defined as follows. By Theorem 1.31, we can choose a locally finite open covering {Ui} of M consisting of coordinate neighborhoods, and a partition of unity {/*} subordinate to it. Then we set Jm t JUi Here, the meaning of the right-hand side is as follows. The support of fiU is contained in a coordinate neighborhood U^ Therefore, if we choose positive coordinate functions xi, • • • ,xn on ?/», the integral / A" is determined. As we observed at the beginning of this subsection, this is independent of the choice of positive coordinate functions. Also, since the support of u is compact and the open covering {f/i} is locally finite, the above integral is 0 except for a finite number of i, and the value of the total sum over i is settled. Proposition 3.5. The definition of the above integral fMto is independent of the choice of the open covering {Ui} consisting of coordinate neighborhoods and the partition of unity {/»} subordinate to it. PROOF. Let {V^} be another locally finite open covering of M consisting of coordinate neighborhoods, and {$,} a partition of unity subordinate to it. Then since Y^9j = 1, by the linearity of the i integral, we have / & w = J2 I fi9iw- JUi ^ J Ui
3.2 INTEGRAL OF DIFFERENTIAL FORMS; STOKES THEOREM 107 On the other hand, the support of figjUJ is contained in ?/»n Vj. Then we have / fi9j"= fi 9j v- JUi JVj Hence Y, f fi" = Y, I fi9iUJ = Yl I /* fc " = H / 9);UJ> i JVi ijJu* ijJv> jJvi and the proof is finished. ¦ The above definition of / u> can be applied to the case of mani- Jm folds with boundaries without much change. Also it is easy to see the linearity of the integral / auj + bv = a w + b n (a,b€R, u>,rj e AT Jm Jm Jm (M)). (b) The Stokes theorem (in the case of manifolds). The Stokes theorem is a fundamental formula concerning the integral of differential forms, and will serve as a base of the de Rham theorem. First we describe the case of manifolds. Theorem 3.6 (Stokes theorem). Let M be an oriented n- dimensional C°° manifold, andu an (n-l)-form on M with compact support. Then du> = u>. Jm JdM Here the right-hand side is the integral of oj on the boundary dM of M, and we assume that dM is equipped with an orientation induced from that of M. Proof. If we choose a locally finite open covering {Ui} of M consisting of coordinate neighborhoods and a partition of unity {fi} subordinate to it, then we have u) = Y^ fi u>. Since the Stokes theorem is obviously linear with respect tow, it is enough to prove the theorem for each fiU>. By the way, the support of fcco is contained in a coordinate neighborhood Ui. Therefore we may carry out the proof under the assumption that M = Rn or Hn. In this case we can write u) = V^aj(x) dx\ A • • • A dxi-i A dxi+i A • • • Adxn.
108 3. THE DE RHAM THEOREM Hence we have dw = (JT, (-1I ^ )dxi A • • • A dxn. In the case of M = Rn, by Pubini's theorem, we have Since di(x) has compact support, we have r J —c = aj(xi,--- ,Xi_i,oo,Xi+i,-• • ,x„) — Oi(xi,--- ,Xj_i, —oo,Xi4-i, • • • ,a = 0. Therefore, we have /. eta = 0. Since Rn has no boundary, the Stokes theorem holds for Rn. Next, let M = W1. In this case, it is enough to change the domain of integration of x„ from 0 to oo in the above consideration, and it is easily seen that only the term including an(x) remains, as follows: / du> = (-l)n / an (xi,- • • ,xn_i,0) dx\ ¦ • -dxn_i. On the other hand, if we restrict u> to c?Hn, obviously only the term including an(x) remains. Recalling the orientation induced on dWn, we have / duj = (-l)n / an (xi,-- • ,x„_i,0) dxi • --dxn-i and the proof of this case is also complete. ¦ Here we wish the reader to tackle Exercises 3.2 and 3.8. The next corollary follows immediately from Theorem 3.6. We state it here because of its importance.
32. INTEGRAL OF DIFFERENTIAL FORMS; STOKES THEOREM 109 Corollary 3.7. Let M be an oriented n-dimensional C°° manifold without boundary. Then for an arbitrary (n — \)-form u> on M with compact support, we have J Jm da> = 0. (c) Integral of differential forms on chains, and the Stokes theorem. Let M be a C°° manifold and 5f(M) = {S?{M),d} the C°° singular chain complex of M (see §3.1(d)). Recall that a C°° singular fc-simplex a : A* —> M of M is a C°° map from (an open neighborhood of) the standard simplex A* in Rk to M. Therefore, for a fc-form u> € Ak(M) on M, the pullback o*w of u; by a is defined. Now we define the integral of cj on a by I uj = I a*u>. Jo M* For a general chain c ? S?°(M), we extend this definition linearly. That is, if it is expressed as c = ^at<7i, we set h'^Lr With regard to this integral on chains, the Stokes theorem takes the following form. THEOREM 3.8 (Stokes theorem on chains). ForaC00 singular k- chain c G S?°(M) of a C°° manifold M and a (k - \)-form u> on M, the equality faw= f Jc Jdc holds. PROOF. By the linearity of the integral, it is enough to prove the case where c is a single singular ^-simplex a. Furthermore, since a*u is a (k - l)-form on A*, we can write k a*u) = 2_. ai{%) dx\ A • • • A dxi-i A dxi+i A • • • A dxk- t=i Again by the linearity of the integral, it is enough to prove the case of a*u = a(x) dx\ A • • • A dXj-\ A dXj+\ A • • • A dxk-
110 3. THE DE RHAM THEOREM Then we have o*du> = (-lV"-1-—dx\ A ••• Adxk, OXj and also Jt da = ]?^(-lI croet, i=0 so that the formula to be proved is C.7) k f = y(-iy I e* {a(x) dx\ A • • • Adxj-i /\dxj+\ A • • • A dxk). Now by the definition C.1), C.2) of €,, we can express e* dxi easily. If we substitute this, only the terms with i = 0, j remain and we see that the right hand side of C.7) is C.8) r k~l (-ly-1 / a(l -Vxi,xi,-- ^k-Adxi-dxk-i +(-l)J / a(xXl-- ,Xj_i,0,a:j,--- ,xk-i)dxi---dxk-i. Here, if we define a diffeomorphism ip : Rfc_1 -*Rk l of Rfc 1 by fc-i <f{Xi,--- ,Xk-l) = [x2,-- ,*j-l,l -^Xi,Xj,-- ,Xjt-iJ, i=i (p transforms Afc_1 onto itself, and since the determinant of its Ja- cobian matrix is (-1)-7-1, its absolute value is 1. Therefore, if we compute the integral of the first term of C.8) by the variable transformation by tpt we obtain . fc-i {-iy~l / a(xu--- tXj-itl-'STxiyXj,--- ,xk-i)dxi '-dxh-i-
3.3. THE DE RHAM THEOREM As a result, the right-hand side of C.7) is C.9) t k~l (-I)' / {a\xi>'" »xi-i»1 - X^Xi'xJ'"" »xfe-i) -a(xi,- • • >Xj-i,0,Xj, • • • ,Xk-i)?dxi • • dxk-i- On the other hand, the integral of the left hand side of C.7) can be expressed as /. —- dxi ¦ • dxk J{A')k~l Wo -— dxjjdxi ¦¦•dxj-idxj+i •••dxfc = / ^a(xi,--- ,Xj-i,l - y^i^Xj+u--- >Xk) —a(xi, • • -,?j_i,0,Xj+i, • • • ,Xk) fdx\ • • • dxj-\dXj+\ • • • dxk- Here (A')fc_1 is the standard (k — l)-simplex in (k — l)-dimensional space obtained by omitting the x^-direction from R*. Here again, if we apply an appropriate change of variables via which (A')fc_1 and Afc_1 are identified, we see that the above integral is equal to f k~l \ \a(xu-- ,Xj_i,l -Tii,!;,"- ,ar*-i) C.10) •/**-» L V fef ' -a(xi,--- .Xj-i.O.Xj,--- ,xjt_i)|dxi---dxfc_i. If we compare C.7), C.9), and C,10), we see that the proof is completed. ¦ 3.3. The de Rham theorem (a) de Rham cohomology. Let M be an n-dimensional C°° manifold. Recall that if we denote the set of all fc-forms on M by Ak{M), a linear map d:Ak{M)—+ Ak+l{M) called the exterior differential operator is defined. A fc-form u G Ak(M) is called a closed form if <Lj = 0, and an exact form if there exists a (k — l)-form 77 such that u> = drj. Since d o d = 0, any
112 3. THE DE RHAM THEOREM exact form is a closed form. Let us denote the set of all closed k-forms on M by Zk(M) and the set of all exact fc-forms by Bk{M). That is, Zk{M) = Ker(d : Ak{M) - Ak+l{M)), Bk{M) = lm(d : Ak'\M) -* Ak{M)). Here Ker and Im stand for the kernel and the image of a linear map. Both Zk(M) and Bk(M) are linear subspaces of Ak{M). Definition 3.9. Let M be an n-dimensional C°° manifold. The quotient space H^R(M) — Zk{M)/Bk{M) of the space of all degree k closed forms Zk{M) on M by the space of all degree k exact forms Bk(M) is called the fc-dimensional de Rham cohomology group of M. For a closed fc-form u> € Ak(M)y we denote the class it represents in the de Rham cohomology group by [u>] € H^^M), and call this the de Rham cohomology class represented by u>. Also we call the direct sum n fc=0 the de Rham cohomology group of M. In other words, the de Rham cohomology group of M is the cohomology of the cochain complex 0 - A°(M) 4 A\M) 4 A2(M) -i ¦ • • -i An(M) - 0. Thus if we set A*{M) = (&nk=0Ak{M), we have H*DR{M) = H*(A*(M);d). This cochain complex is called the de Rham complex. On the other hand, A*(M) is equipped with a product structure defined by the exterior product. This product structure induces that on H*DR{M) in the following way. If x 6 #?,fl(M), y G HlDR(M) are represented by closed forms u 6 Zk(M), rj e Zl(M) respectively, then we set Here, since d(u/ A 77) = du A 77 + (-l)fcu> A dry = 0, u> A 77 is certainly a closed form. Also, if we let u/ = u; + d?, 7/ = 77 + dr, then w'Arj' = (o; + d?) A G7 + dr) = u A 77 + <2((-l)fcu> Ar + ?A77 + ?A dr).
3.3. THE DE RHAM THEOREM 113 Hence the product xy is determined independently of the choice of closed forms representing x, y. Also, obviously, yx = {-l)klxy. H*DR{M) equipped with the product structure is called the de Rham cohomology algebra. Let M, N be C°° manifolds and / : M -> N a C°° map. Then, the pullback /* : A*{N) —> A*{M) of differential forms by / induces a homomorphism /* : HpR(N) —> H*DR{M) of de Rham cohomology algebras. In practice, if x e HpR(N) is represented by a degree k closed form u> on N as x = [u>], then /*cj is a degree fc closed form on M, and we set /*(x) = [/*cj]. It is easy to see that this is well defined independently of the choice of u>, by Proposition 2.10. Concerning the product, we have f*{xy) = f*(x)f*(y) for x,y € H^R(N). Also if we let g : N —» P be a C°° map, for the composition p o / : M —> P, we have {go f)* = f* o g*. (b) The de Rham theorem. Since the de Rham cohomology of a C°° manifold is defined using differential forms, it would seem to depend essentially on the differ- entiable structure of M. However, in reality, it is determined only by the properties of M as a topological space. It is the de Rahm theorem that expresses this fact concretely. Let M be a C°° manifold. Then, two cochain complexes are defined for M, namely the de Rham complex {A*{M)f d} and the C°° singular cochain complex {S^M), 8}. It is the integral of differential forms on chains which gives a relationship between these two. We define a map C.11) I:Ak{M)-^S^{M) as follows. First, for u> € Ak{M) and each C°° singular fc-simplex a : Ak -¦ M (e Sg°(M)), let I(u)){cr) — / <7*u>, and for an arbitrary singular /c-chain c € S^{M) define /(w)(c) by extending this linearly. Let us denote the collection of maps C.11) for each degree k by the same symbol / : A*(M) —» 5^(M).
114 3. THE DE RHAM THEOREM Lemma 3.10. The map I : A*{M) -* S^(M) is a cochain map. That is, the diagram Ak{M) —i-» Ak+l{M) 'I I' is commutative. PROOF. Let u) e Ak{M) be an arbitrary fc-form and c e S^^M) an arbitrary singular (A: + l)-chain. Then, by the Stokes Theorem 3.8 on chains, we have I{aw)(c)= [dw= [ u = I{u){dc). Jc Jdc Therefore, we have I od = 6o I. ¦ By Lemma 3.10, the map / : A*{M) —> 5^(M) induces a homo- morphism / : H*DR{M) -¦ H*{S^(M)). Theorem 3.11 (de Rham theorem). Let M be a C°° manifold. Then the cochain map I: A*{M) —* 5^0(M) induces an isomorphism I:HhR{M)&H*{Sroo(M)). As we remarked in §3.1(d), the natural inclusion 5J°(Af) C 5* (M) induces an isomorphism H.(S?(M)) S H*(S.{M)) = H.(M; Z), and therefore, H*(S^H(M)) is naturally isomorphic to the singular coho- mology group H*(M\R) of M with coefficients in R. If we combine this fact and the above de Rham theorem, we obtain a natural isomorphism H*DR(M)*H*(M;R). Especially, we see that the de Rham cohomology is topologically invariant; that is, the de Rham cohomologies of two homeomorphic C°° manifolds are isomorphic in a natural way. The above Theorem 3.11 takes a general form that holds for an arbitrary C°° manifold. However, it might be rather difficult to understand because it involves the integral on singular chains and there exist a huge number of singular chains. Therefore in the case where M is equipped with a C°° triangulation t : \K\ —> M as in §3.1(c), we shall state the de Rham theorem in an easier formulation.
3.3. THE DE RHAM THEOREM 115 Let (a) = (v0 ¦ ¦ ¦ vi) be an arbitrary oriented /-simplex of K, and uj E Al(M) an arbitrary Worm on M. Then, the integral of uj on (<t), ho) is defined as follows. Assume that the polyhedron \K\\s realized in RN for a sufficiently large JV, and let L be the /-dimensional subspace spanned by \a\ in RN. L is diffeomorphic to R', and an orientation is given on it which is induced by the orientation of (a). By the definition of C°° triangulation, t||a| : \a\ —* M can be extended to a C°° map from an open neighborhood U of \a\ in L to M. Hence we can consider t*u as an /-form on U. Therefore, the integral / t*u J\a\ is defined, and, using this, we set / cj = / t*u>. J(c) J\o\ If we define a map I:A*(M)—>C*{K;R) by /(a;)«a» = / u/, J(a) we see that / is a cochain map in the same way as Lemma 3.10. Then the following theorem holds. THEOREM 3.12 ( de Rham theorem for triangulated manifolds ). Let M be a C°° manifold and suppose that a C°° triangulation t : \K\ —> M is given. Then, the cochain map I : A*{M) —> C*(K\R) induces an isomorphism I:HbR(M)*Sr{KiR). For triangulated manifolds, Theorem 3.11 (the de Rham theorem) can be deduced from Theorem 3.12 as follows. We give a total order on the set of vertices of the simplicial complex K, and for an arbitrary /-simplex a = {vo, • • • , vi} of K we consider only the order of vertices such that vo < ••• < v/. Then a chain map C+(K) —» S?°(M) is defined, and it is known that this is a chain homotopy equivalence. This chain map induces a cochain map S^(M) —» C*(K;R), which gives an isomorphism H*(SZo(M)) = H*(K\R) in cohomology. Also,
116 3. THE DE RHAM THEOREM by the definition of the map /, it is easy to verify that the following diagram is commutative: HhRW -+-* H^S^M)) II I Hdr(M) —7- JH*;R) Therefore Theorem 3.11 follows from Theorem 3.12. Here, we shall explain the geometric meaning of the de Rham theorem. In general, the dimension of Hk{M\R) is denoted by 0k and called the fc-dimensional Betti number of M. This number expresses the essential number of fc-dimensional cycles in M, and is an important quantity that reflects the global structure of M. In some cases, the de Rham theorem not only determines this Betti number but also describes the way cycles are distributed in M. In order to "detect" a certain ^-dimensional cycle z in M, it is enough to construct a closed fc-form wonM such that jz u) ^ 0. An important fact here is that even if we replace 2 by an arbitrary cycle z' homologous to it, or replace u by an arbitrary closed form u/ cohomologous to it, the value of the integral never changes. In some cases, we can guess the place z occupies from the form of appropriately chosen u. Now the de Rham theorem claims the following. Let z\, • • • , zr (r = /3k) be linearly independent cycles which generate Hk{M',R)» and let oi, • • • , aT be arbitrary real numbers. Then there exists a closed form u> such that fz u) = a» for any i which is unique except for the ambiguity of adding exact forms. If we state it like this, we can understand how fundamental the de Rham theorem is. (c) Poincare lemma. The proof of the de Rham theorem 3.11 will be given in the next section. In this subsection we study the case of Rn as a preparation for it. Also we encourage the reader to solve Exercises 3.5 and 3.6 which will help understanding of the general case. Proposition 3.13. Let M be a C°° manifold. Let-n : M xR-* M be the projection to the first factor, and i:M-+MxRo map defined by i(p) = (p,0) (pG M). Then the map 7T- : HhR(M) —> H*DR(M x R) induced by rr is an isomorphism, and i* : H*DR(M x R) —> H*DR{M) is its inverse.
3.3. THE DE RHAM THEOREM 117 PROOF. Since obviously Trot = id^, we have i* on* = id. Therefore, it is enough to prove that it* oi* = id on H*DR{M x R). For this, it is enough to construct a linear map (such a map is called a cochain homotopy) $ : Ak{M x R) —> Ak~\M x R) connecting the identity map id and it* o i* in such a way that id - it* o i* — (d$ + $d) holds on Ak(M x R). The reason is that d$ + $d maps closed forms to exact forms and therefore becomes the 0 map on cohomology. Let uj € Ak(M x R) be an arbitrary Ar-form. For a local coordinate system A7; X\, • • • , xn) of M and a coordinate t of R, we can write uj as u) = 22 aii-ik(x>t) dxn A • • • Adxik ii< <tfc + Y^ ^ji-jk-i (x,t)dt A dXjx A • • • A dXjk_x, ji<-<jk-i where the latter term contains dt while the former does not. Now, paying attention to the second term, we define *" = Yl ( / bi* Jk-Ax^)dt)dxh A-.-Ada^.,. Then, for an arbitrary u> e Ak{M x R) it is enough to show that C.12) d{$u) + ${dw) = uj - it* o i*u). By linearity, it is enough to verify the following two cases. (i) uj = a(x, t)dxiv A • • • A dxik. (ii) uj = b{x, t)dt A dxJt A • • • A dxjk_l. In the case of (i), since $uj = 0 and rt q ${<kj) - / — dt dxix A • • • A dxik Jo vt = (a(x, 0 - a(x, 0))dxil A • • • A dxik — uj — tt* o i*u,
3. THE DE RHAM THEOREM C.12) holds. Next, in the case where u> is of the form (ii), since i*w = 0, we have (id - tt* oi*)uj — u>. On the other hand, d{$uj) - d( ( / 6(x,t)dt\ dxh A • ¦ • Adxjk_x\ dt) dxm A dxjx A • • • A dxjk dx. *(dw) = $(- ]P -— dt A dxm A dxh A • • • A dxjk_x\ = -]?(/ ¦^-dt^dxmAdxjlA---Adxjk_l. Therefore d($uj) + ${dw) — u>, and in this case C.12) also holds. Although in the above computation the local expression of u; is used, it is easy to see that $u; is determined independently of the choice of the local coordinate. Hence, $o> is a differential form defined on the whole of M, and C.12) holds for it. This completes the proof. ¦ By induction on n, we obtain the following corollary. Corollary 3.14 (Poincare lemma). The de Rham cohomology of Rn is trivial. That is, Hk(Rn) = Hk(one point) =r> * = °' In other words, if u> € ,4fc(Rn) (k > 0) is an arbitrary closed form, there exists a (k - I)-form n such that u> = dn. Corollary 3.15. Let M,N be C°° manifolds. If two C°° maps from M to N are homotopic, then the homomorphisms HqR(N) —> H*DR{M) induced by them are the same. PROOF. Let f,g : M —> N be homotopic C°° maps. Then there exists a C°° map F : M xR —> N such that -C ™- S.';"' If we take io,ii : M —¦ M x R to be i0(p) = (p>0), i\{p) = (p, 1), obviously we have / = Foi0) g = F o ix. Now by the proof of Proposition 3.13, we see that Zq = i\ = (n*). Therefore, /* = {Foi0y =i*oF* =t*oF* = (Foi1)* =g\
3.4. PROOF OF THE DE RHAM THEOREM 119 and the proof is finished. ¦ Two C°° manifolds M, N are said to have the same homotopy type, if there exist C°° maps / : M -> N and g : N -* M such that 9 ° /) f ° 9 are nomotopic to the identity maps of M, N respectively. Also a manifold with the same homotopy type as one point is said to be contractible. Corollary 3.16 (Homotopy invariance of de Rham cohomol- ogy). The de Rham cohomologies ofC°° manifolds with the same homotopy type are isomorphic. In particular, the de Rham cohomology of a contractible manifold is trivial. 3.4. Proof of the de Rham theorem (a) Cech cohomology. To prepare for the proof of de Rham theorem, we shall introduce Cech cohomology. Let X be a topological space. For an open covering U = {Ua}aeA of X, the Cech cohomology group H*{X\U) of X with respect to li is defined as follows. First we define a simplicial complex N(U) called the nerve of the open covering U. We take the set A of subscripts of the open covering U as the set of vertices of N(U), and k + 1 distinct elements ao, • • • ,ctk in A are assumed to span a /c-simplex if Uao D • • • fl Uak is not empty. That is, we put JV(W) = {{a0,•••,<*}; uaon---nuak*0}. It may be almost trivial that N(U) becomes an (abstract) simplicial complex. Then, the Cech cohomology group is defined as H*{X;U) = H*{N{U)). Here we may take an arbitrary Abelian group as the coefficients of cohomology. However hereafter we use only the real number field R. Example 3.17. Let K be a. (Euclidean) simplicial complex and V the set of vertices of K. For each simplex o € K, the set obtained from \cr\ by removing its boundary is denoted by (a), and this is called the open simplex of a. Also, for each vertex v € V, we put 0(v)= |J (a) v?o€K and call this the open star of v. That is, 0(v) is the union of all open simplices (a) of those simplices a of K which have v as a vertex. Now if we put U = {0(v)\v e K}, it is obviously an open covering of
120 3. THE DE RHAM THEOREM \K\. A brief consideration shows that we have a natural identification N{U) = K. This is because, given any distinct 1+1 vertices vq, • • • , vi of K, it is easy to see that a necessary and sufficient condition for them to span a simplex of K is O(u0)nO(ui)n-.-nO(vi) 7^0 (see Figure 3.4). Therefore, in this case the Cech cohomology group H*(\K\\U) of \K\ with respect to U can be identified with the ordinary cohomology group H*(K) of K. Figure 3.4. Intersection of open stars In order to define the Cech cohomology group directly without using the notion of nerve, we can proceed as follows. That is, to any ordered set c*o, - • • , a* of fc + 1 distinct elements of A, we assign a real number c(oto, • • • ,a/c) such that for an arbitrary permutation c(a<0,--- ,aik) = sgn i c(a0, • • • ,<**), and we call it a /c-cochain of X with respect to U. The set Ck{X;U) of all fc-cochains naturally becomes a vector space. The coboundary operator S:Ck(X-U)~>Ck+l{X\U) is defined by putting fc+i <5c(a0,--- ,<*fc+i) =X^-1)lc(a°>"*' »<*«»••• '^+1) i=0 for c € ^(X;^/). Then it is easy to see that 5 o 5 = 0, and therefore C*(X;?/) = {Ck(X;U)}6} becomes a cochain complex. The cohomology of this cochain complex is nothing but the Cech cohomology group H*{X\U).
3.4. PROOF OF THE DE RHAM THEOREM 121 (b) Comparison of de Rham cohomology and Cech coho- mology. In this subsection, as the first step of the proof of the de Rham theorem, we prove Theorem 3.19 below by the method of Weil. In general, an open covering U = {Ua} of a topological space X is called a contractible open covering, if the intersections of a finite number of open sets belonging to U are all contractible. PROPOSITION 3.18. An arbitrary C°° manifold M has a contractible open covering. PROOF. We shall give two different methods of constructing contractible open coverings. First let t : |jFsT| —> M be a C°° triangulation of M, and identify \K\ with M via t. Then it is easy to see that the set {0(v);v € V} of all open stars of each vertex of K is a contractible open covering of M. In the second method, elementary Riemannian geometry is used. If a Riemannian metric (see §4.1(a)) is given on M, we can take a geodesically convex open neighborhood around each point. The open covering consisting of such open neighborhoods is contractible, because the intersection of geodesically convex open sets obviously has the same property and therefore is contractible. ¦ THEOREM 3.19. Let M be a C°° manifold. Then, for an arbitrary contractible open covering U of M, there exists a natural isomorphism HhR(M)^H"(M;U). For the proof of this theorem, we prepare a few matters. First of all, to make a connection between the de Rham cohomology and the Cech cohomology, we consider the following. Let U = {Ua\cx € A}, and denote by Ak'l{U) the set of all assignments of an element w(a0, • • •,*k) e Al(Uao n • • • n /yQ,) to any ordered set Qo, • • • , ctk of k + 1 distinct elements of A, such that, for an arbitrary permutation ct{0, ¦ • ¦ ,aik of Qo, • • • , Qfc, o/(a<0,--- ,aifc) =sgniu{a0,--- ,ak).
122 3. THE DE RHAM THEOREM This has a natural structure of a vector space. We define two boundary operators 6 : Ak'l{U) —» Ak+U{U), d:Ak\U)—>Ak*+l{U) by setting fc+i (<5w)(o:o,--- ,afc+i) = 5^(-l)Mtto,--- ,3i,--- ,ak+i), t=0 (du;)(Q0)---,afc)=d(a;(ao,---,afc))G^+1(^aon---nGaJ, for u> G •4fc,i(lY). Here u;(ao,--- ><*i,-•• ,a:jfc+i) means precisely the Z-form restricted to UQo n • • • n UQk+l. By a brief consideration, we see that 8o6 = 0, dod = 0> Sod = do6. Now let us consider the following commutative diagram: C.13) 4 Al(M) — 4 4 A\M) — 4 ^°(M) — 4 4 '—+ A°'l(U) —5—* ^-'(W) 4 4 4 4 ^— .A0-1^) —^-» ^1-1(W) 4 4 ^ «4°'°(W) —-*—» .4X'0(W) 4 4 C°(W) —*—» C!(W) J 5 6 a 6 <5 s 6 4 ,4fc''(W)--. 4 4 Ak-x(U)--- 4 >tfc"°(W)--- 4 Ck(U)- Here r : ^(M) —* A0,l{U) is the map determined by restricting any /-form u) € ^(M) that is defined on the whole of M to each open set
3.4 PROOF OF THE DE RHAM THEOREM 123 Ua. Also Ck(U) stands for Ck{M\U), and t : Ck{U) - Ak>°{U) is the natural inclusion map. In the above diagram, the left column is the de Rham complex and the bottom row is the Cech complex. If we remove these two, it becomes a commutative diagram such that Ak,l{U) (kj > 0) are sitting side by side in the first quadrant, and each row and column is a complex with respect to the boundary operator 6 or d. Such a complex is called a double complex. PROPOSITION 3.20. For arbitrary k,l>0, C.14) 0 —- Al(M) -% A°'l(U) -!>...-!. Ak'l(U) -^ • • • , C.15) 0 —> Ck{U) -U Akfi(U) -** • • • -±> Ak\U) -^- • • • are exact sequences. That is, the kernel of an arbitrary homomor- phism in each sequence coincides exactly with the image of the preceding homomorphism. PROOF. We shall see that the first sequence is exact. First, it is obvious that r : Al{M) —* A0,l(U) is an injection. Next, assume that cu € A0,l(U) satisfies 6(u>) = 0. This means that for arbitrary ce,{3 e A, the restrictions of u)(a) and u>(@) to Uq D Up coincide. Therefore a; is a differential form defined on the whole of M. That is, w is in the image of the map r. For general k > 0, we define $:Ak>l{U)—+Ak~u{U) as follows. We choose a partition of unity {fa} subordinate to the open covering U = {UQ}, and for w € Ak'l{U) we set D>w)(a0, • • • ,arfc_i) = ^faw{a,a0,• • • ,ajk-i). a Here, although fau(a,ao,- • • ,ctk-\) is a differential form on Ua f) Ua0 H • • • n C/afc-i > ^ putting 0 outside of UQ, we can consider it as an element of Al(Uao D • • • D C/afc_J. Then for an arbitrary element u> € Ak'l{U), we see that C.16) 8($u)) + $Fu>) = u>.
124 3. THE DE RHAM THEOREM In fact, since k <5($cj)(a0,--- ,afc) =^2 (-1I ($w)(a0l--- ,5;,- • ,afc) t=0 z=0 a and $Eu>)(a0,--- ,«fc) = ]T]/a (<M(or,a0,--- >a*) a = (X^/a Mao,--- ,ojk) a fc + ^(-ir+1/aW(a1a0l-,5),..,4 a t=0 C.16) holds. Now if w G .A*''(W) satisfies <5u; = 0, C.16) implies that w = E($u>). Hence the first sequence is exact. Next we shall show that the second sequence is also exact. First, it is obvious that the map i : Ck{U) —» Ak'°(U) is an injection. Next, assume that u € Ak'l{U) satisfies du) = 0. This means that the function u>(giq, • ¦ ¦ , a*) on Uao A • • • n Uait becomes 0 if we take the exterior differentiation. On the other hand, since U is a contractible open covering, in particular Uao D • • • n Uait are all connected. Therefore uj(ao,--- ,ak) is a constant function, and we see that uj is in the image of i. Next, let I > 0, and assume that cj € Ak,l{U) satisfies du = 0. Then, by applying Corollary 3.16 on each contractible manifold UaQ n • • • n Uak, we see that the second sequence is exact. ¦ Proof of Theorem 3.19. We prove the theorem using the commutative diagram C.13). First of all, we shall construct a candidate for the isomorphic correspondence C.17) ip : HlDR{M) —¦ H\M\U) which the theorem claims. For a given de Rham cohomology class x € HlDR(M)t we have to define tp(x) € Hl{M\U). For it, we choose a closed form u e Al(M) on M representing x, and put r(u>) = u>q € A°'l(U). Now, since duj0 = d(r(u>)) = r{du>) = 0, by Proposition 3.20
3.4. PROOF OF THE DE RHAM THEOREM 125 there exists an tj0 G A°'l~l(U) such that dfjo = oi0- Then we put <*>i = 6vo- Since du>i = dErjo) = 6(dr)o) = 6u>q = 6(r(u)) = 0, again by Proposition 3.20, there exists an 771 G A1,l~2(U) such that dr)i = u>i. Then we put u>2 = 6rji. Now assume that w< G Ai,l~i(U) has been constructed successively by similar operations. Then, since du>i = 0 inductively, there exists r)i G Al'l~x~l(U) such that u>i = di)i. Then we put vi+l = Stj, g •4i+u~i~1(^0 (see diagrams C.18), and C.19)): <T ••• —i— A^-'iU) C.18) d| >li'/-<-1(W) 0 4 <5 6 A C19) 4 4 4 In this way, we finally reach ^ G Al,0{U). Since du^ = 0, by Proposition 3.20 there exists c G <?*(?/) such that uji = i(c). Since 5c = 0 by construction, c is a cocycle. Then we put <p(x) — [c] G Hl(U). Now we shall prove that (p(x) is uniquely determined, independently of various choices made in the definition. Assume that we start from another closed form u/ G Al{M) representing x, and by the same discussion as above u>[ G Ax,l~x{U) and 77^ G Ax'l~x~l{U) are determined and finally we reached a cochain d G Cl{U). We have to prove that c and d are cohomologous. Now there exists an element 70 G Al~l(M) such that u/ = <*> + d^o. Therefore we have u>'0 = ujq + r(^7o) = wo + d(rGo)). Now we shall show inductively that for an arbitrary i = 0, !,-•• }l, there exists an element Ai+1'l-{{U) A
126 3. THE DE RHAM THEOREM 7» e A1-1'1-*-1^) such that C.20) u/J = u)i + d{6-yi). (Here if i = 0, 6 is replaced by r and ^-1-'_1(W) stands for Al~l{M)\ and if i = I, d is replaced by i and «4'~1,_1(ZY) stands for Cl~l(JA).) Since the case of i = 0 is already shown, we assume that it is true up to i and prove the case of i + 1. Since u>i = drji, u[ = dr^ by definition, we have d{Vi — Vi ~ ^7t) = <*>( — Wj — <2(<$7i) = 0. Thus, by Proposition 3.20 there is an element ji+i e Ai'l~i~2 {U) such that d7j+i = tj • — t?j - 57^. Then we have wi+i = Srj'i = Srji + <5(d7i+i) -u)i+i +d(^7i+i), and we see that C.20) holds. Now in C.20), if we put i = I, we see that there exists 7* 6 Cl~x{U) such that u/[ = 07 + i(<S7f). On the other hand, since u/{ = t(c') and w/ = i(c), we have c' = c + $7;, so that c' and c are indeed cohomologous. In this way we have defined a map <p : H*DR{M) - H*{M;U). In the above discussion, starting from .4'(M) we followed a zigzag course down and to the right in the diagram C.13) and reached Cl{U). Conversely, if we start from Cl(U), and by a similar discussion as above we follow a zigzag course up and to the left in the same diagram C.13), we can reach Al{M). By this operation, a map 1>:&*{M-tU)—*HhR{M) is defined. Then it is easy to see that the above two operations are inverse to each other. Hence <p and ip are each other's inverses, so that C.17) is an isomorphism. ¦ (c) Proof of the de Rham theorem. In the previous subsection, we proved that for an arbitrary con- tractible open covering U of a C°° manifold M, there exists a natural isomorphism C.21) <p:H*DR(M)*H*(M;U) (Theorem 3.19) between the de Rham cohomology group and the Cech cohomology group. However, the integral of differential forms, which is the essence of the de Rham theorem, is hidden in this proof. In this subsection, we supply this point and prove the de Rham theorem in the form stated in §3.3(b). We prove Theorem 3.12 for C°°
3.4. PROOF OF THE DE RHAM THEOREM 127 triangulated manifolds, rather than Theorem 3.11 in the general form. This is because the idea of the former is easier to understand, and also because Theorem 3.11 can be deduced from Theorem 3.12, as was already remarked after Theorem 3.12. Let M be a C°° manifold and t : \K\ —» M a C°° triangulation of it. Hereafter we identify \K\ with M. Let V be the set of vertices of K; as in the proof of Proposition 3.18, if we put U = {0(v)\ v G V}, this is a contractible open covering of M. Furthermore, the Cech cohomology group H*(M;U) of M with respect toU can be naturally identified with the real cohomology group H*(K;R) of the simplicial complex K and so further with H*(M\R): H*{M\U) = H*{K;R) = H*(M;R) (see Example 3.17). Then recall that the degree I part <p : HlDR{M) —* Hl{M-U) = Hl{K;R) of the isomorphism C.21) is given explicitly as follows. For x G HlDR{M), we first choose a closed form u> e Al(M) representing x. Next, we choose 770 € A°'l~l(U) so that cfyo = r(u>), and, inductively, we choose r)i € At'l~'t~1(U) so that ^_i = drji. Finally if we choose c 6 Cl(U) — Cl(K\'R) such that <5t7j_i = i(c), c becomes a cocycle and we define ip(x) = [c] <E Hl{K;R) = Hl{M;R). On the other hand, the map / : HlDR(M) -» Hl(K;R) in Theorem 3.12 is given as follows. As above we assume that x € HlDR(M) is represented by a closed form u>. Then, for an arbitrary oriented /-simplex (cr) = (v0 • • • vi) of K, if we put co((<7» = / w, ho) Co € Cl{K; R) becomes a cocycle. Then, we define I{x) — [co]. Now to prove Theorem 3.12, it is enough to show that if we naturally identify Hl{M\U) and Hl(K;R)y the two maps / : HlDR{M) - Hl(K\R) and ip : HlDR(M) —* Hl{M\U) essentially coincide. The following proposition guarantees this fact, and with it the proof of Theorem 3.12 will be finished.
3. THE DE RHAM THEOREM Proposition 3.21. The diagram HlCD{M) —L- Hl(K;R) HlDR{M) —^ Hl{M-M) is commutative up to sign. That is, if we put €i = (-1) a , we have I = ei<p. PROOF. As above, if x € HlDR{M) is represented by a closed form u> € Al(M), then I(x) is the cohomology class represented by a cocycle cq e Cl{K;R) and <p(x) is the cohomology class represented by a cocycle c G Cl(K;R). Therefore to prove the proposition, it is enough to show that Co and ejc are cohomologous. First, we can see by the Stokes Theorem 3.8 that, for an arbitrary oriented /-simplex (vq- • -vi) of Kf co{(v0 ¦¦¦v[))= u)= I drH(v0) J(v0-vi) J(v0---vi) C.22) ,-n J(v0 ¦¦Oi-vi) i=0 J{v0 •¦¦Oi-vl) Now we shall define a cochain do € CZ_1(A";R) by putting d0({v0vi •v/_i)) = / 77o(t>o) J{vQvi -v,_i) for an arbitrary oriented (I — l)-simplex (voi>i • • ¦ v<-i)- Then, we have i 6d0((v0---vl))=Yl(-iydo((vo---Vi---vi)) i=0 = / VoM + Y^i-lY J Voivo)- J(ViVt) i=1 J{VQ--Vi--Vl)
3.4. PROOF OF THE DE RHAM THEOREM 12 Hence c0{(vo---vi))-Sd0{(vo---vi))= VoM- *7oM J(vi-vt) J(vi---vi) = - drji{voVi) J{Vi--Vt) 1=1 J{v0—X>i~-Vi} The reader should check this by referring to Figure 3.5. Figure 3.5 Next we define a cochain d\ € Cl~l(K\R) by putting di{(v0vi---vi-i)) = / Vi%vi) J(vi---vi-i) for an arbitrary oriented (/ - l)-simplex (vo^i ¦¦•vi-i) of K. Then we have 6di{(v0 ¦ ¦ ¦ vi)) = ^2(-iydi{{v0---Vi---vi)) i=0 = / m{v\V2) - / mivov*) J{v2-vi) J(v2-vi) i=2 J(v0-Vi-vt)
130 3. THE DE RHAM THEOREM Therefore we have (c0 -Sd0 - 6di){(v0 ¦ ¦ • vt)) = / Vi(voV2) - mivovi) - r)i(viV2) drJ{voViv2) L {Vi-Vl) = -^(-i)' / m{vQViv2). i=2 J(v2- Vi- vt) In general we define a cochain di € C'-^K'jR) (i = 0,1, •••,/- 1) by putting didvovi • • • vi-i)) = / ^(v0vi---Vt = / '¦ JiVi-Vl.x)) for an arbitrary oriented (/ — l)-simplex (i>oUi • ¦ • v/-i) of if. Then, repeating the above discussion, we see finally that i-i CO + J^Ct+l&Ji = €jC. i=0 Here tj = (-1) 2~^. We wish the reader might try the next step where di appears. Then one may understand how signs are attached in the above formula. In this way, it is shown that the two cocycles co,c are surely cohomologous up to sign, and the proof is finished. However, strictly speaking the above discussion has a problem. That is, while we used the Stokes theorem in formula C.22), the differential form rH is only defined on an open star 0(v) of each vertex v, so that it is not clear at all whether the values of such integrals / drH(v0), / Tto(v0) J(v0-Vt) J {Vo -Vi-Vl) exist. The situation is the same later in the discussion, using r)x. However, this problem can be easily solved as follows. First, instead of the open star 0(v) of each vertex v we choose a slightly larger open set 0'(v) containing 0(v) such that O,{v0)n---nO,{vk) are always contractible. Next we argue in the same way using U' = (O'(v)} instead of the open covering U = {0(v)}. Then the values of integrals as above are determined, and the Stokes theorem can be applied without any problem. To show that we can choose U' =
3.4. PROOF OF THE DE RHAM THEOREM 131 {0'(v)} satisfying the above properties, we need a discussion using an operation called the barycentric subdivision of a simplicial complex. However, we omit it here. It is not hard to see intuitively that such a matter is possible. ¦ (d) The de Rham theorem and product structure. As we saw in §3.3(a), the exterior product of differential forms induces a natural product structure on the de Rham cohomology of manifolds, and equipped with this H*DR(M) becomes an algebra. On the other hand, a product called the cup product is denned on the ordinary (singular) cohomology H*(M;R). If M is triangulated by t : \K\ —> M, the cup product is defined also on H*(K\R). The de Rham theorem has a good consequence with respect to this product structure. Namely, the following theorem holds. THEOREM 3.22 (de Rham theorem concerning the product). Let M be a C°° manifold. Then the isomorphism I:H'DR(M)*H'(M;R) in Theorem 3.11 preserves the product structure. Hence it becomes an isomorphism of algebras. If a triangulation t : \K\ —> M is given on M, the isomorphism I : H*DR{M) * H*{K;R) = H*(M;R) in Theorem 3.12 is similarly an isomorphism of algebras. Proof. Here we shall prove only the latter half of the claim, because in this case the idea of the proof seems easier to understand than in the general case. First let us recall the definition of the cup product of the cohomology of the polyhedron \K\. If a, r are two arbitrary simplices of K, their product \a\ x |r| is naturally a cell of the product space \K\ x \K\. A cellular decomposition of \K\ x \K\ is obtained by considering all such cells. Now let c,d € C*(K;R) be cochains of K of degrees k and / respectively. Then their cross product ex c\ defined below, is a degree k +1 cellular cochain of the cell complex K x K. Practically, if we let (a), (r) be a, r equipped with orientations, the product (a) x (r) becomes an oriented cell of \K\ x \K\ and for this we put c x c'((a) x <r» = c((a))c'((r)). Concerning the boundary operator, it is easy to check that the equality S(c x d) = 6c x d + (-l)kc x 6c' holds. Therefore, if both c, d are cocycles, c x d is also a cocycle. Then the cohomology class
132 3. THE DE RHAM THEOREM [c x c') 6 Hk+l(\K\ x \K\]R) is called the cross product of [c] and [c'\ and is denoted by [c] x (c']. Now let d : \K\ -+ \K\ x \K\ be the diagonal map, that is, a map defined by d(p) = (p,p) (p ? |/f|), and let d* : H*{\K\ x |tf|;R) -»'/f(|if|-;R) be the homomorphism induced by d. Then recall that the cup product [c] U [c'] € Hk+l(\K\\R) of two cohomology classes [c], [c'] is defined by [c]u[c'j=<f([c]x[c']). With the above preparation, we shall prove the claim of the theorem. We assume that two de Rham cohomology classes x G HpR(M), y e HlDR(M) are represented by closed forms u),r) € A*(M) respectively. Then their product xy 6 H^Rl(M) is represented by the closed form u> A 77. On the other hand, with respect to the isomorphism I:H*DR(M)—*H'(K-R) of the de Rham theorem, I(x) € Hk{K;R), I{y) € Hl{K;R) are represented by the following cocycles c, d € C*(if;R) respectively. That is, for an arbitrary oriented fc-simplex (a) and an {-simplex (r), c((a)) = / c, </«r» = / r,. Next we consider the product manifold M x M. If we let pi : M x M —* M (i = 1,2) be the projection to the ith component, then p\u> A P27? is a (fc + /)-form on M x M, and obviously it is a closed form. We denote it by u) x t] and call it the cross product of u and 77. The cross product induces a homomorphism Hhn(M) ® H*DR(M) B (H, [rj]) H[uq]6 JFT^CAf x M) of de Rham cohomologies. We call this the cross product of de Rham cohomology and write [a; x 77] = [uj] x [77]. Now if we let d : M —> M x M be the diagonal map as above, obviously <T(pJu>Ap2*?) = ^A77, because Pi o d = id. Therefore, we have d*([w) x [77]) = [uj A 77]. Here let us consider the following commutative diagram: H*DR(M x M) -i— JT(|ff| x |ff|;R) C.23) d*{ |d* ^dh(M) —y- H*(\K\;R) The commutativity of this diagram follows easily from the de Rham theorem in its general form, that is, Theorem 3.11. Now we can choose the following cocycle as a cocycle c of the cell complex K xK
35. APPLICATIONS OF THE DE RHAM THEOREiM 133 representing I([uj} x [77]). That is, an arbitrary oriented {k + l)-ce\\ of K x K has the form of (o) x (r) (a, r e K) and the value for this is given by C.24) c((a) x <r» = / w x 77. J(a)x(r) This fact can be seen easily if we subdivide the cell complex K x. K so that it becomes a simplicial complex and recall the definition of I for the manifold M x M. Now the integral of the right-hand side of C.24) is obviously 0 if the dimensions of o, r differ from k, I respectively, and otherwise it is exactly equal to / w tj, J{<r) J(t) namely c x d{{a) x (r)). Therefore, we have c — c x d. By the commutativity of the diagram C.23), we have I{\u) A 77]) = d*([c\) = d*{\c x c'\) = [c] U [c7] = 7(H) U 7((t?]), and the proof is finished. ¦ It might be a little hard to understand the definition of the cup product of ordinary cohomology compared with the natural definition of the product of de Rham cohomology. This is because, although the direct product K x K of a simplicial complex K naturally becomes a cell complex as mentioned above, a kind of artificial operation is necessary to make it be a simplicial complex. This is a theoretically inevitable fact. Actually, a deeper structure than the cup product, called cohomology operations, can be defined on cohomology from it. The so-called Alexander-Whitney map is one of the above artificial operations, and by it the cup product is defined at the cocycle level. It is possible to prove Theorem 3.22 using this, and it is enough to follow the discussion of the previous subsection keeping in mind the cup product of cocycles. Although it is necessary to pay attention to the sign, it should not be so difficult, and we hope the interested reader will try it. 3.5. Applications of the de Rham theorem (a) Hopf invariant. For an arbitrary C°° map / : S3 —> S2 from the 3-dimensional sphere to the 2-dimensional sphere, we define a real number H(f) € R as follows. First we choose a 2-form 6 € A2(S2) on S2 such that Js2 $ = 1. Then since d{f*6) = f*{d$) = 0, f*6 is a closed 2-form on S3. Since H2{S3;R) = 0, by the de Rham theorem
134 3. THE DE RHAM THEOREM there exists a 1-forrn 77 e A1{S3) such that f*9 = dr\. Then we define H(f) = [ r,Adrj. Js* H{f) is called the Hopf invariant of /. Theorem 3.23. (i) The value of H{f) is determined independently of the choices of 9 and 77, and thus it depends only on f. (ii) The value of H(f) depends only on the homotopy class of f. That is, if two C°° maps /o, f\ : S3 —» S2 are homotopic, thenH{fQ) = H{h). (iii) Let h : S3 — S2 be the Hopf map (§i.3, Example 1.27). Then H(h) = 1. PROOF. First we prove (i). Let 9' be another 2-forra on S2 such that Jsj 9' — 1 and let f"B' = dn'. We have to show that C.25) / rjAdr) = f n' A dn'. Js3 Js* Now since H2(S2\R) = R, by the de Rham theorem, there exists a r € A1{S2) such that 9' = 9 + dr. Then since d(n' - 77 - f*r) = f*{9' - 9 - dr) = 0 and also H1^3;*) = 0, again by the de Rham theorem, there exists a g E A°(S3) such that V = r] + f'r + dg. Therefore, we have 77' A dr) = G7 + /*r + dp) A (dr? + f*dr) = 77 A dn + 77 A d(/V) +/'(tA@ + dr)) + d(s(dr7 + f*dr)) = 77 A dT? + 77 A d(/*r) + d(p(d77 + f'dr)). The last equality follows because S2 is a 2-dimensional manifold, so that the 3-form r A (9 + dr) on it becomes 0. Next, since 77 A d{f*r) = -d(n A fr) + dn A /*r = -^G7 A /*r) + /*@ A r)) = -dfa A /*t), we have 77' A dr)' = 77 A dn + d(—77 A fr + 5(^77 -f- fdr))} and C.25) is shown. Next we prove (ii). By the assumption, there exists a continuous map F: S3xR-* S2 such that F{p, 0) = /0(p), F(p, 1) = /i(p) (p € 53). As we know, a continuous map can be approximated by a C°° map, so we may assume that F is of class C°°. Since H2(S3 x R; R) =
3.5. APPLICATIONS OF THE DE RHAM THEOREM 135 0, there exists a 1-form rj e A^S3 x R) such that drj = F*9. Let to : 53 x {0} -* 53 x R, tx : S3 x {1} -» S3 xR be the natural inclusion maps and let itf = 770, ij77 = 771. Then drH = f?9, drji = f{0. If we apply the Stokes theorem to the manifold S3 x [0,1] with boundary and the 3-form rj A drj on it, we have / d(rj A drj) = I rj A drj JS3x\0,l) Jd(S*x[0A)) = / ViKdoi- T?0 Adrjo. ^S3x{l} «/S3x{0} Since ^G7 A drT) = F*9 A F*0 = F*(9 A 0) = 0, we have / r?o Adr;o = / 771 Adr/i, ./S3x{0} 7s3x{i} and (ii) is proved. Finally, we prove (iii). Here, instead of arguing directly, we shall give a proof using the Euler class which will be introduced in Chapter 6. The Hopf map has the structure of a principal S1 bundle and its Euler class is exactly -1 € H2(s2;Z) = Z, as will be shown there. On the other hand, the Euler class of principal S1 bundles can be computed using differential forms as follows. First a 1-form u € A1(S3) called the connection form is defined, and then we can show that there exists a 2-form CI € A2(S2) called the curvature form such that Since the Euler class is defined as the de Rham cohomology class represented by the closed 2-form - — fi, in the present case we have 2tt 2ir jS2 /n- Js* Therefore, to compute the Hopf invariant H(/i), we can put 9 — — Cl. 2ir Then, since if we put 77 = — w, drj = h*9, we have 2n H{h)= I 77 A d-q Js* *2 JS* 4tt2 = 1
136 3. THE DE RHAM THEOREM and the proof is finished. Here we used the fact that the integral of the connection form u; on the fiber is equal to the length of Sl, namely 2ir. fl (b) The Massey product. The differential forms contain complete information on the real cohomology of manifolds, as the de Rham theorem shows. More precisely, this information includes both the vector space structure of real cohomology and the algebra structure defined by the cup product. In this subsection, we show that differential forms can measure a deeper structure defined on cohomology. While this structure contains an infinite sequence of higher degree products called Massey products, here we will introduce only the simplest one. Let M be a C°° manifold. We assume that certain de Rham cohomology classes x,y,z of degrees kj,m respectively are given on M and they satisfy the relation xy = yz = 0. Then a cohomology class (xty,z,)eHkD^+m-\M)/I(x,z)y called the Massey triple product of x, y, z, is defined as follows. Here I(x,z) = x • Hl+?-\M) + z • HkD+?-\M) expresses the subspace of Hkj?+m~l(M) consisting of all cohomology classes of the forms xu + zv {u e Hl^~l(M), v e tf??_1(M)). Now let us choose closed forms a, 0, 7 € A*(M) on M representing x, y, z respectively. Then by the assumption, there exist differential forms A,/z e A*{M) such that a/\0 = dX, P A 7 = dfx. Then, since d{Xj - {-l)ka A^)=aA/9A7-aA/0A7 = O, A7 - (-l)fco; A [i is a closed form. Now we set <x,y,*) = [A7-(-l)fcaA/z]. It is easy to verify that this cohomology class is uniquely determined as an element in the quotient space by the subspace /(x, 2), independently of the choice of differential forms used in the definition. We will leave it to the reader to verify the details.
3.5 APPLICATIONS OF THE DE RHAM THEOREM 137 Example 3.24. Let M be an S1 bundle on the 2-dimensional torus T2 such that its Euler class is 1 € H2(T2;Z) (see Chapter 6 for the terminology). M becomes an oriented 3-dimensional closed manifold. Practically we can also construct M in the following way. We define a 3-dimensional Lie group N by N has u,v,w as its (local) coordinate and can be identified with R3. Let T be a subgroup of N consisting of all elements whose entries are all integers. Then T acts naturally on N by the multiplication of matrices, and we see that this action is free and properly discontinuous. Therefore, by Proposition 1.52 of §1.5, the quotient space N/T becomes a 3-dimensional C°° manifold, and this is the above M. A simple computation shows that the left-invariant 1-forms on N are generated by du, dv, 7 = dw + udv. Therefore, they can be thought of as 1-forms on M. In this case, we see that HXDR{M), HpR(M) are 2-dimensional, and we can choose x = [du], y = [dv] and [du/\-y], [dv/\*f] as their generators respectively. Since dj = du A dv, we have xy = 0. Therefore the Massey product (x,x,y) € H2DR(M) /I{x,y) is defined. Obviously I{x,y) = 0 and (x,x,y) = [duA1]^0eH2DR(M). Thus this is a simple example where the Massey product is non-trivial. (c) Cohomology of compact Lie groups. Let G be a Lie group and g its Lie algebra. By §2.4(b), we can identify the dual space g* of g with the set of all left-invariant 1-forms on G. Particularly, there is a natural inclusion map i : g* —> A1(G). By taking exterior products of these left invariant 1-forms, we see that i induces a homomorphism i:AV — A"(G) (we use the same symbol). If we consider the values on the tangent space TeG at the identity of G, we see that i is an injection and furthermore its image coincides with the set of all left-invariant differential forms on G. Hereafter, we consider that A*g* = the set of all left-invariant differential forms on G.
138 3. THE DE RHAM THEOREM The exterior differentiations of left-invariant differential forms are also left-invariant. Practically it is computed by the Maurer-Cartan equation B.45). Therefore A*g* is closed in the de Rham complex A*(G) of G with respect to the operation of exterior differentiation. That is, it becomes a subcomplex. In general the homomorphism t* : JT(AV) —> H*{A*(G))^Hm{G\R) induced by i is neither an injection nor a surjection. However, in the case where G is compact, the following theorem holds. Theorem 3.25 (Cartan-Eilenberg). Let G be a connected compact Lie group. Then the natural inclusion map i : A*g* —> A*(G) from the set of all left-invariant differential forms A*g* on G induces an isomorphism H*(A*q*) = H*(G;R). The proof can be given by averaging differential forms in the same way as the proof of Exercise 3.10 using the Haar measure on G. However, here we omit it. Example 3.26. Since the n-dimensional torus Tn can also be regarded as the n-fold product of 50B), it is a connected compact Lie group. Its Lie algebra is commutative and can be naturally identified with Rn. Therefore, by Theorem 3.25, we have /T(rn;R)^A*(R)\ (d) Mapping degree. First of all, we shall consider how the orientability of C°° manifolds can be expressed by means of differential forms. Let M be an n-dimensional C°° manifold. Then, since dim hnT*M = 1, if an n- form u> which does not vanish at any point on an open set U of M is given, an arbitrary n-form on U can be uniquely expressed as a multiple of u; by a function. In particular, the "ratio" wi/u>2 of two n-forms ui, c^2 that are not 0 at any point is defined as a function on U. Therefore, if U is connected, we can classify all those n-forms into two classes. That is, if oj\/uj2 has positive value all over U, we let w\ and u>2 belong to the same class. Now the n-form u> = dx\ A • • • A dxn on Rn obviously does not become 0 on Rn. Let C/, V be connected open sets of Rn, and let <p : U —* V be a diffeomorphism. Then, by §2.1 B.4), we have (p'cj = Jy> uj. Here J<p stands for the Jacobian of </?. Therefore, a necessary and sufficient condition for two n-forms (p*u> and u> on U to belong to
35. APPLICATIONS OF THE DE RHAM THEOREM 139 the same class in the above sense is that the Jacobian of <p is always positive on U. Prom the above observation, the following proposition may be thought to be natural. PROPOSITION 3.27. A necessary and sufficient condition for an n-dimensional C°° manifold M to be orientable is that there exists an n-form that does not vanish at any point on M. Proof. First let us assume that there exists an n-form u> that does not vanish at any point on M. Then, the collection of local coordinate systems ([/, f) of M such that <p*{dx\ A • • • A dxn) belongs to the same class as u; is obviously an atlas of M, and the Jacobians of its coordinate changes are always positive. Therefore M is orientable. Conversely, assume that M is orientable. Then there exists an atlas S = {{Ua> V?a)}a€v4 sucn that the Jacobians of coordinate changes are always positive. By taking a locally finite refinement if necessary, we may assume that the open covering {Ua} consists of countably many elements and there exists a partition of unity {/Q} subordinate to it having the same subscripts. Then, if we put u> = 2_J fa<p*Q(dxi A • • • A dxn), this is an n-form that does not vanish at any point on M. ¦ It is obvious from the above proof that, if we give an n-form which does not vanish anywhere on an n-dimensional C°° manifold, an orientation associated to it is determined; and conversely, on an oriented n-dimensional manifold, we can take a nowhere vanishing n-form that is coherent to its orientation. Such an n-form is called a volume form. Now let M, TV be oriented connected closed manifolds of the same dimension (say n), and let / : M —> N be a C°° map. Let /* : Hn(M; Z) —> Hn(N\ Z) be the homomorphism induced by / on n- dimensional homology groups. By assumption, Hn(M;Z), Hn(N\Z) are infinite cyclic groups generated by the fundamental classes [M], [./V] of M, N respectively. Therefore, there exists an integer d € Z such that f.([M}) = d\N}. This is called the mapping degree of / and denoted by deg /. Intuitively, the mapping degree counts the number of windings of M round TV by /.
140 3. THE DE RHAM THEOREM Example 3.28. A) Let Sl = {ze C; \z\ = 1} and define fn : S1 - Sl (n € Z) by /n(z) = *n- Then deg/ = n. B) Let tt : S'2n+1 —> L(p;gi,--- ,qn) be the natural projection from S2n+1 to a lens space (see §1.5, Example 1.53). Then deg-n- = p. The following proposition gives the relation between mapping degree and de Rham cohomology. As the proof is not so difficult, we leave it to the reader as an exercise. Proposition 3.29. Let M, N be n-dimensional oriented connected closed manifolds, and let f : M —* N be a C°° map. Let u € An(N) be an arbitrary n-form on N. Then / f*u> = degf / u>. Jm Jn In particular, if u is a volume form of N such that Jnuj = 1, we have deg/= / f*u). Jm (e) Integral expression of the linking number by Gauss. Assume that two disjoint knots (that is, embedded S1 's) K, L are given in R3. Such a set-up is called a link with two components. As in Figure 3.6, there are various cases. If we try to explain the difference between (a) and (b) intuitively, we can say that in case (a) the two S^'s are intrinsically linking, while in case (b) they are not. We shall formulate this fact mathematically. (a) (b) Figure 3.6. Various links First we give orientations on K, L, and let f:Sl—+ R3, g : S1 —> R3
35. APPLICATIONS OF THE DE RHAM THEOREM 141 be their expressions by parameters. Then a map F : S1 x Sl —> R3 is defined by F{s,t) = g{t) - f{s) (s,t € Sl = R/Z). Here g(t) - f(s) expresses the subtraction as vectors of R3. If we denote the space obtained from R3 by removing the origin by R3 - {0}, then the image of F is contained in R3 - {0} by assumption. Now we define a projection n : R3 - {0} -* S2 by n(x) = x/||x||, and put F = n o F. Then we set Lk(tf,L)=degF and call this the linking number of K and L. Since the mapping degree is a homotopy invariant, we see that the above number is determined independently of the choice of parameters of K, L. We intend to measure the way of linking of two knots by the mapping degree, which is a fundamental invariant of mappings. We shall look for a concrete expression of the linking number in terms of an integral. First of all, we put oj = ,, {x\dx2 A dx3 - x2dxi A dxz + x^dx\ A dx2), \\x\\* and furthermore we denote the restriction of u to S2 by u>o- Then by a simple computation we see that du = 0 (Chapter 2, Exercise 2.7), and that / u>q = 47r (Exercise 3.8). Therefore, we have C.26) Lk(K,L) = ^- f F*u. 47T yT2 Now /, g are 3-dimensional vector valued functions defined on S1; let /i> 9i (i = 1,2,3) be their components. Then we have C.27) F'dxi = fi(s)ds + g\{t)dt (i = 1,2,3). From the formulae C.26) and C.27), a simple computation finally yields the following theorem. Theorem 3.30. The linking number of two disjoint oriented knots K, L in R3 is given by the integral Lk(K, L) Here f, g are the expressions of K, L by parameters, respectively.
142 3. THE DE RHAM THEOREM The above integral expression of the linking number was obtained by Gauss. Also the linking number can be defined more generally for two disjoint oriented submanifolds in an oriented n-dimensional closed manifold, such that the sum of their dimensions is exactly n - 1. Summary 3.1 The homology of a simplicial complex is the quotient of the module consisting of ail cycles by the submodule consisting of all the boundaries. 3.2 An arbitrary C°° manifold has a C°° triangulation. 3.3 The n-dimensional homology group of an n-dimensional connected orientable closed manifold is isomorphic to Z. A generator of this group, which is determined if an orientation is specified, is called the fundamental class. 3.4 Assume that an (n - l)-form u> with compact support is given on an oriented n-dimensional C°° manifold M. Then the integral of dw on M is equal to the integral of uj on dM. This is called the Stokes theorem. In particular, the integral of exterior differentiation of an arbitrary (n - l)-form on an n- dimensional closed manifold is always 0. 3.5 For an arbitrary (k - l)-form u; and an arbitrary singular k- chain c on a C°° manifold, the integral value of do; on c is equal to the integral value of u> on dc. This is called the Stokes theorem (with respect to chains). In particular, the integral of an exact form on a cycle is always 0. 3.6 The quotient space of the space consisting of all closed forms on a C°° manifold divided by the subspace consisting of all exact forms is called the de Rham cohomology group. Equipped with a product induced by the exterior product of differential forms, the de Rham cohomology group becomes an algebra. This is called the de Rham cohomology algebra. 3.7 The map from the de Rham cohomology of a C°° manifold to the singular cohomology induced by the integral of differential forms on chains is an isomorphism of algebras. This is called the de Rham theorem. Exercises 3.1 Referring to the proof of Theorem 3.4, show the following facts.
A) Let M be an n-dimensional closed C°° manifold and suppose that it is connected and unorientable. Then, we have tfn(M;Z) = 0. B) Let M be an n-dimensional compact C°° manifold with boundary and suppose that it is connected and orientable. Then we have#n(M,dM;Z) = Z. C) If M is a connected n-dimensional closed manifold, then we have#n(M;Z2) = Z2. 3.2 Show that the fundamental theorem of calculus: r f(x)dx = f(b) - f(a) can be regarded as a special case of the Stokes theorem. 3.3 A) Let M, N be oriented n-dimensional C°° manifolds and / : M —> N an orientation-preserving diffeomorphism. Then for an arbitrary n-form u> with compact support on N, show that / /*U>= / w. JM JN B) Let M be an oriented n-dimensional C°° manifold, and u> an n-form with compact support on M. Prove that / uj = J-M u> (here —M is the manifold M with the reversed orien- JM tation). 3.4 Let M be a connected n-dimensional closed manifold and uj € Ak(M), T) e An~k{M) closed forms on M. Prove that if a; An is not 0 at any point on M, the de Rham cohomology class [u;] € H^R{M) represented by u> is not 0. 3.5 Prove that HkDRm = i% fc = 0' DRK } \0, Jfc>0, by the definition of de Rham cohomology. 3.6 Prove that ^*(s') = {R' fc = 0,1' DRK } @, fc>l, by the definition of de Rham cohomology.
44 3 THE DE RHAM THEOREM 3.7 We denote the space obtained from R2 by removing the origin byR2-{0}. Then, compute H*DR{R2-{0}). Also, find a closed 1-form that represents a non-zero element of HpR(R2 — {0}). 3.8 For a 2-form u> = x\dx2 A dx% — X2dx\ A dx$ + x$dxi A dx^ on R3, find the value of the integral Js* Here S2 is the unit sphere in R3. 3.9 Let M be a connected oriented n-dimensional closed manifold. Then, prove that for an arbitrary integer d, there exists a C°° map / : M —» Sn whose mapping degree is d. 3.10 Assume that a finite group G acts freely on a C°° manifold M. Then, show that the de Rham cohomology of the quotient manifold M/G is given by HhR(M/G) s* HhR(Mf. Here the right-hand side denotes the set of all the invariant elements of H*DR{M) under the action of G.
CHAPTER 4 Laplacian and Harmonic Forms In this chapter we discuss differential forms on Riemannian manifolds. Roughly speaking, a Riemannian manifold is a differentiable manifold such that the length is defined for each tangent vector. The notion of differentiable manifold originates from a surface in R3 as a model. But the notion is abstract, and thus it may not be imbedded or immersed in Rn. Hence the concept of length or that of curvature for a curve on the manifold is not defined. Rather it was Riemann's epoch-making idea to define these notions as structures on a differentiable manifold. To speak more concretely, we simply assume that we have the notion of length of each tangent vector; then as consequences we get the notions of volume of the manifold and its shape, including the way it is curved. In this age, such ideas are accepted as commonplace, but at the time such facts were recognized, by Gauss for surfaces in R3, they were startling discoveries even for Gauss himself. Now that the length of tangent vector is defined, the magnitude of a differential form is also determined. By making use of this fact we can give a more precise statement to the theorem of de Rham. That is, from the point of view of magnitude, we may prove that within the set of all closed forms representing a de Rham cohomology class, there is one and only one differential form that has the best shape. Such a form is called a harmonic form, and it can be characterized by using a differential operator called the Laplacian. This is the theory of harmonic forms (or harmonic integrals) due to Hodge. In this book our aim is to explain the ideas that will lead up to this theory. For the details, the reader is referred to [War] or [deR] in the References. 4.1. Differential forms on Riemannian manifolds (a) Riemannian metric. We now state the definition of a Riemannian manifold. 145
146 4. LAPLACIAN AND HARMONIC FORMS Definition 4.1. Let M be a C°° manifold. If for each point p e M the tangent space TPM is provided with a positive-definite inner product gp : TPM x TPM -* R in such a way that gp is of class C°° in p, we say that g = {gp\ p € M} is a Riemannian metric on M. We also say that M is a Riemann- ian manifold. More precisely, an inner product on a real vector space V (it is TPM in the present case) means a symmetric bilinear map /i: V x V —* R; it is said to be positive-definite if /x(t>,t>) > 0 for all v € V and if fj.(v, v) = 0 implies that v = 0. Whenever we say an inner product from now on, we shall assume that it is positive-definite. Let (U; x\,..., xn) be a local coordinate system for M. If we set ^ = *(?'?) (pe?/)' then pi, is a function of X\,..., xn. We say that g is of class C°° if the functions p^ are of class C°° in all local coordinates. (In fact, we shall see in a moment that this is the case if it holds for any single local coordinate system around each point p.) Indeed, suppose (V; V\ > • • • > Vn) is another local coordinate system around p. Then the formula in Proposition 1.34 of §1.3 gives rise to from which it follows that hke{p) are also of class C°°. We may also write ds2 = V^ gijdxidxj, t,j=i which is a classical way of expressing the Riemannian metric as a quadratic form, that is, a symmetric tensor of degree 2. Here for another coordinate system we have the same expression ds2 - ^jT hijdyidyj.
4.1. DIFFERENTIAL FORMS ON RIEMANNIAN MANIFOLDS 147 Example 4.2. Relative to the canonical coordinates x\,... , xn, in Rn, ds2 = dx\ H h <?r2 is a Riemannian metric. This space Rn with the metric is called an n-dimensional Euclidean space. In other words, ds2 is a metric such that dx\' " ' dxn is an orthonormal basis. Example 4.3. The upper half-plane H2 = {(x.y) e R2;y > 0} provided with a Riemannian metric ,2 dx2 + dy2 ds = 5 y2 is called the hyperbolic plane. Using this, we can develop non- Euclidean (hyperbolic) geometry. EXAMPLE 4.4. An arbitrary submanifold of a Riemannian manifold has the induced Riemannian metric. It is called a Riemannian submanifold. Let (M,g) be a Riemannian manifold. For each tangent vector X € TPM, its length ||X|| is defined by \\X\\=y/j(X^C). Since the metric is positive-definite, we see that ||X|| is 0 if and only if X = 0. Suppose a curve C of class C°° on M is given by a C°° map: c : [a, b] —* M. Then its length L{C) is defined by L(C) = [ \\c(t)\\dt Ja This means that the length attained is given by integrating the speed (i.e. the length of the velocity vector). We can easily verify that the integral is independent of parametrization. The next proposition will show that any differentiable manifold admits a Riemannian metric. There is a lot of freedom in choosing a Riemannian metric, as can be imagined by changing around the shape of a surface (without changing the differentiable structure). PROPOSITION 4.5. On an arbitrary C00manifold there exists a Riemannian metric.
148 4. LAPLACIAN AND HARMONIC FORMS Proof. Let J7» (i = 1,2,...) be a locally-finite open covering with a partition of unity {/*}. Since we can regard each Ui as an open submanifold in the Euclidean space Rn, there is an induced metric, say &, on it. By setting g = ?V /t0t, we can verify that it is a Riemannian metric on M. ¦ (b) Riemannian metric and differential forms. Let {M,g) be a Riemannian manifold. For each point p 6 M, there is a positive-definite inner product gp : TPM x TPM -* R. By setting gp(X)(Y)=gp(X)Y) we get a linear map gp:TpM-+T;M. To show that this linear map is injective, assume that gp{X) = 0. Then we have gp(X)(X) = 0, that is, g(X,X) = 0, showing that X = 0 because g is positive-definite. On the other hand, we have dim TPM = dim T*M. Hence gp is an isomorphism, that is, by means of the metric gp we can identify the tangent space TPM and the cotangent space T*M. We may then extend this identification to the space X(M) of all vector fields on M and the space A1(M) of all differential forms of degree lonM. For example, for each C°°function / on M, df is a 1-form. By the isomorphism Al(M) = X(M), there is a unique vector field, denoted by grad/ and called the gradient of /, such that g(gcaAf,X)=df(X)=Xf, for every vector-field X on M. For a function / = /(xi,... ,xn) on the Euclidean space Rn, we have (See Exercise 4.3.) Again in the general case, suppose grad/ has no zero point in M. Then we have the notion of "level hypersurface", namely, a hypersur- face on which / is constant and grad/ is perpendicular to each level hypersurface (see Figure 4.1). As we showed earlier, the Riemannian metric induces an isomorphism Al{M) = X(M). For any two 1-forms u and rj, we have their inner product (u>p,-qp) at each point p, and so we have a function (u>,77) on M. We shall generalize this fact to the case of k-forms.
4.1. DIFFERENTIAL FORMS ON RIEMANNIAN MANIFOLDS 149 Figure 4.1. grad/ (Here we are using the symbol (,) to distinguish it from (,) that will result by integration of (,) in §4.2.) We need to go through some linear algebra. Let V be a vector space. Using a positive-definite inner product given on V, we may induce an isomorphism V = V* and hence an inner product in V* as well. Let k > 1. For two elements of the form ai A • • • A ak and Pi A • • • A 0k {oci,0j 6 V*), we define the value of their inner product to be (qi A • • • A ak, Pi A • • • A pk) = det ((a*, Pj)). That this value is independent of the way the two elements are represented follows from the properties of exterior product and determinant. We now extend the inner product so defined to the whole space A^V* by linearity. If t\,..., en is an orthonormal basis of V and 9\t... ,$n the dual basis, then all the elements of the form 0iyK--A0ik, 1 <ti < ••• < tfc <n, form an orthonormal basis of AfcV*, as the reader may easily verify. In this way, for any two fc-forms u> and 77 on M, we have the inner product {up,r}p) for each p ? M, and thus the function @^,77) on M. Note that in the special case k = 0, we define the inner product between functions / and g at each point p simply as the product of the values of / and g at p. We also define the inner product between two differential forms of different degrees to be 0. Example 4.6. For two 2-forms on R3 a; = a dxi A dx2 + b dx2 A dx$ + c dx$ A dxi, 77 = edxi A dx2 + f dx-i A dxz + gdx^ /\dx\,
150 4. LAPLACIAN AND HARMONIC FORMS we have (u;, 77) = ae + bf + eg. (c) The *-operator of Hodge. Let M be an n-dimensional C°° manifold. For any integer k @ < k < n), hkT*M and hn^kT*M have the same dimension as vector spaces, and so they are isomorphic. If M has a Riemannian metric and also is oriented, there is a natural isomorphism *: kkT;M ^An~kT;M, for each point p e M. By varying p e M, we get a linear isomorphism * : Ak{M) - An~k{M), where Ak{M) (resp. An~k{M)) is the vector space of all k-forms (resp. (n - fc)-forms) on M. The first part of definition (for a fixed point p) follows the same line of linear algebra. So write V instead of TPM. The inner product in V induces an inner product in V* and also in AfcV*. Furthermore, a given orientation in M induces a natural orientation in V and V*. Let 0i,...,0*,&k+i, • • • 0n be an arbitrary positively oriented orthonormal basis. Then we can get a linear map * : AkV* -» An~kV by setting In particular, we have *1 =01 A---A0„, *@i A---A0„) = 1. In this way, * is defined as a linear isomorphism in the natural fashion. (The reader may want to verify our assertion.) Now the Hodge operator * : Ak{M) —* An~k{M) can be defined globally. In other words, if u € Ak{M), then *w is an element in An~k(M) whose value at p is equal to *up for every p. Only one thing is not quite obvious. If u> 6 Ak{M), how does one know that *u> is of class C°°? To clarify this matter, let us find a concrete expression for *uj. Let {U;x\,... ,xn) be a local coordinate system that is positive relative to the orientation. Thus we may assume that X{ = ?- form a positive frame field. Take the Gram-Schmidt orthogonalization process and get an orthonormal frame ei,...,en.
4.1. DIFFERENTIAL FORMS ON RIEMANNIAN MANIFOLDS 151 Specifically, we let e\ = Xi/\\Xi\\ and inductively define Yi = Xi-Y^g{Xl>ej)eh « = VIMI- for i = 2,3,..., n. We let 6if 1 < i < n, be the dual basis of e*, 1 < i < n. Now if "= ? /ii-iAA-'-A^, then we have D.1) *u; = ? sgn(I,J)fix...ik6hA---A9jn_k. Here ji < • • • < jn_fc is the rearrangement of the complement of «i < • • • < ijt in the set {1,..., n} in ascending order, and sgn(J, J) is the sign of the permutation ti,...,ijt.ii • • • Jn-k- The fact that *u> is of class C°° follows from this representation. We have *1 G An(M), which is called the volume form or volume element and which will be denoted by vm- A concrete expression is given by t>M = 0i A • • • A 0n. In terms of the metric components gij we have vm = \/det(pij)(ixi A • • • A dxn} which the reader might want to verify (as Exercise 4.4). For a domain D C M, JD vm is the volume of D. In particular, if M is compact, fM vm is called the volume of M. PROPOSITION 4.7. The *-operator of Hodge has the following properties. For any f and g in C°°(M) and for any u and n in Ak{M) we have (i) *{fuj + gn) = f*u>-rg*r). (ii) **u = {-l)k(n-Vu>. (iii) o> A *r) = n A *w = (u;, t])vm ¦ (iv) *(oj A *rj) = +G7 A *uj) = (uj,n). (v) (*uj,*v) = (w,T7).
152 4. LAPLACIAN AND HARMONIC FORMS PROOF. It suffices to prove each identity at every point p. (i) is obvious. To show (ii), let 9\,..., 9n be a positive orthonormal basis in T*M and assume that u = 9\ A • • • A 9k- In this case, we have *u>p = 6k+i A • • • A 9n. Applying * once more, we get * * up = (~\)k^n~k^u>p. Next we show (iii). Since each term is linear in 77, we may assume, in the argument above, that r)p = #», A • • • A 9ik. In this case, we have *r)p = sgnG, J)9jx A • • • A 9jn_k by using the notation in D.1). Therefore the only time u>p A *t?p ^ 0 holds is when {ii,..•,i/J = {1,..., k}, and then we have u>p A *r)p = sgn I9\ A ... 0n, where sgn J is the sign of the permutation i\}.. -ik- On the other hand, we have {ujp, r)p) ^ 0 only when {ii,..., t*} = {l,...,fc}, in which case we have (up,r)p) = sgn/. We have thus proved that u> A *t? = {u),tj)vm- The remaining equality can be proved in a similar way. The proof of (iv) can be done by applying * to (iii) and using *vm = 1. Finally, (v) follows from (ii) and (iv). ¦ Now let M be an oriented Riemannian manifold. For any X € X(M)} let ux be the 1-form corresponding to X by the isomorphism X{M)^Al{M). We set divX = *d *u>x- It is called the divergence of X. If M comes with opposite orientation, * changes to (-*). Since * appears twice in the definition of div, it follows that divX is defined independently of the choice of orientation. Example 4.8. On a Euclidean space Rn, to X = J2ifi& tnere corresponds ux = J2i fidxi. Hence we get by direct computation. The physical meaning of divergence is this. If for n = 3 we assume that a vector field X represents a motion of incompressible liquid, then divX expresses the amount of spring-up at each point. If D is a domain with smooth boundary in the space, then the total amount JD divXdv of spring-up within D is equal to the total amount JdD(X,n)da that is going out of the boundary 3D. Here n is the
4.2. LAPLACIAN AND HARMONIC FORMS Figure 4.2. The Gauss formula outward unit normal vector field on dD, and dv and da represent the volume elements of R3 and dD, respectively. (See Figure 4.2.) This fact is tied to the Gauss formula. More generally, we have Theorem 4.9. Let M be an oriented compact Riemannian manifold. If X is a vector field on M, then we have the equality / (MvXvm = / {X,n)vdM, Jm JdM where n is the outward unit normal vector field on dM. In particular, if M is a closed manifold, then / div Xvm — 0. Jm There is a nice proof based on the theorem of Stokes (§3.2, Theorem 3.6). We leave a proof as Exercise 4.5 for the reader. 4.2. Laplacian and harmonic forms In this section, let {M,g) be an oriented Riemannian manifold, compact and without boundary. We note, however, that compactness is needed only when the formula D.2) below is used. In part (b) of the preceding section, we defined the inner product {ojp,rjp) at each point for two A;-forms on M. We now introduce the inner product in Ak(M) by integrating the function {ujp,t)p) over M, that is, D.2) (W|i7)= / (uj,t])vm,
154 4. LAPLACIAN AND HARMONIC FORMS where vm is the volume element of M. The following properties hold, (i) (linearity) (au> + bu',77) = a(u;,T?) + 6(u/, 77). (ii) (symmetry) A^,77) = G7,u>). (iii) (positive-definite) (w,a>) > 0; (w,u;) = 0 if and only if w = 0. This is certainly an inner product on the vector space Ak(M). In particular, the length ||u>|| = y/(uj,uj) is defined. According to Proposition 4.7 (iii), the inner product D.2) can also be written in the form D.3) (w,t|) = / u> A *r? = / r)/\*u). Jm Jm We have also, from Proposition 4.7 (v), (*u>,*77) = (a;, 77), which means that the Hodge operator * : Ak(M) —> An~k{M) is isometric relative to the inner product above. By convention, we define the inner product between differential forms of two different degrees to be zero, so that the entire space A"{M) is provided with an inner product. Next we study how exterior differentiation d : A*{M) —* A*{M) is transformed by the Hodge operator. For this purpose, we define a linear operator 6 = (-l)fc *~x d* = (-i)»(k+D+i * d* by requiring that the following diagram be commutative: Ak{M) —-i—> An~k(M) 4 I- Ak~l(M) > An~k+1{M). (-Dfc* From the definition, we immediately see that *6 = {-l)kd*> 6* = {-l)k+l * d, 6o6 = 0. PROPOSITION 4.10. Relative to the inner product ( , ) in A*{M), 6 is an adjoint operator of exterior differentiation d; that is, we have D.4) (dw,i?) = (w,ET7). Conversely, d is an adjoint operator of 6.
4.2. LAPLACIAN AND HARMONIC FORMS 155 PROOF. It suffices to prove D.4) when u> and rj are k- and (fc + 1)- forms, respectively. In this case, we have duj A*r) = d(u> A +77) - (-l)fcu; Ad*r) = d(u> A *rj) + uj A *6rj. Integrating each side over M, we get from D.3) {dw,r}) = / d{u A +77) + (u>,6rj). Jm By the theorem of Stokes (§3.2, Corollary 3.7) we get JM d{ujA*r}) = 0 and hence D.4). ¦ Here we are using the term adjoint operator. It can also be called the conjugate operator. In general, for any linear operator T : V —* V of a vector space V, the linear operator T* satisfying (Tv,w) = (v,T*w) (v,w€V) is called the adjoint of T. In the case where T* = T, it is said to be self-adjoint or self-conjugate. Definition 4.11. For a Riemannian manifold M, the operator defined by A = dS + 6d : Ak{M) -» Ak{M) is called the Laplacian or Laplace-Beltrami operator. A form uj € A*{M) such that Au; = 0 is called a harmonic form. In particular, a function such that A/ = 0 is called a harmonic function. Example 4.12. Let us compute the Laplacian on an n-dimen- sional Euclidean space Rn. It is sufficient to compute Acj for a fc-form written uj = fdxi1 A • • • A dxik, where x\,...,xn are the ordinary coordinates in Rn. First, choose ji,... ,jn-k such that dxix A • • • A dxik A dxjx A • • • A dxJn_k = dx\ A • • • A dxn. Then we get *u> — fdxji A • • • A dxjn_k, by the definition of *. By carrying out computation following the definition:
156 4. LAPLACIAN AND HARMONIC FORMS we get v1- df d*u) = > ——dxii A dXjl A • • • A dxJn_k, 5=1 la 6u> = ^(-l)s—^-dxi, A ••• AdxT, A---/\dxik. 3 = 1 ** Therefore we obtain D.5) dSu = - ^ -—- d^i, A • •• A dxik s=l ** + EE(-1M^-?-^Acte^A,-Ad^A--Adx'- 5=i t=i oxXtaxjt On the other hand, we have dw = V^ dxjt A d^i, A • • • A Xifc, and thus 1-* Or *du; = ^(-l)'14^-^ A-'-AdXj-, A---Adijn_fc. 5=1 •*« Further, we have n~k q2 r d*da; = ^(-l)fc^fdxJ1A.-.ACbjn_(c 5 = 1 ¦?» + E ^(-l)^3 ^ dxit A dlj, A • • • A dxU A • • • A dXjn_k. 5=1 t=i C2rj,^a:it With careful attention to signs we can compute D.6) n~k d2f 6du) = - E "aT"^*1 A " " A ^i* s=i dxi> n~k k rp f + J2 Y] °i {-l)t+1dxjs A dxu A • • • A dxlt A • • • A dxik. s=l (=1 °XJ*dxU
4 2. LAPLACIAN AND HARMONIC FORMS 157 Finally, adding D.5) and D.6), we arrive at \^ & j 5=1 ox* In this way, we have seen that for a differential form on a Euclidean space the Laplacian acts (with opposite sign) on the coefficients just like the classical operator Yl*=i ^F5"- ^ *s rather the other way around; that is, the Laplacian is an extension of the classical Laplace operator to the case of a general Riemannian manifold. PROPOSITION 4.13. The Laplacian A has the following properties: (i) *A = A * . If u is a harmonic form, so is *u>. (ii) A is self-adjoint, that is, (Aw, r/) = (uj,Arj) for allu>,r) e A*{M). (iii) A necessary and sufficient condition for Aw = 0 is that dw = 0 and 5w = 0. Proof. The proof of (i) is easy and is left to the reader, (ii) follows from the fact that d and 5 are adjoint to each other (Proposition 4.10). Let us prove (iii). If du> = 5u> = 0, then clearly Aw = 0. To show the converse, we need the assumption that M is compact. In this case, the equality (Aw, w) = ((dS + 6d)uj,u>) = (<5w, 6u) + (duj, du) = 0 shows that Acj = 0 implies ||<M| = I l*M I = 0, that is, du> — 5u> = 0. Corollary 4.14. Suppose M is a connected, oriented, compact Riemannian manifold. Then a harmonic function on M is a constant function. If n = dimM, then a harmonic n-form is a constant multiple of the volume element vm ¦ PROOF. If a function on M satisfies A/ = 0, then (iii) of the proposition above' implies df — 0. Hence / is a constant function if M is connected. Now any n-form on M must be a function times the volume element: u> = /%. If Aw = 0, (i) of Proposition 4.13 shows ¦u; = *(fvM) = / is a harmonic function, hence a constant / = c, that is, w = cvm. ¦
158 4. LAPLACIAN AND HARMONIC FORMS Now let M be an oriented n-dimensional compact Riemannian manifold. If r is the number of connected components, then both H°DR{M) and /fpfi(M) are naturally isomorphic to the direct sum of r copies of R (see Theorems 3.4 and 3.11 of Chapter 3). From Corollary 4.14, it follows that, for k = 0 and for k = n, every element of HqR is represented by a uniquely determined harmonic form. As a matter of fact, this fact remains valid for every k, as we see in the theorem of Hodge in the next section. 4.3. The Hodge theorem In this section we continue the assumption that M is an oriented compact Riemannian manifold without boundary. (a) The Hodge theorem and the Hodge decomposition of differential forms. We consider the set Ak{M) of all k-forms on M and denote it simply by Ak. Denote by Hfc(M), or simply Hfc, the set of all harmonic /c-forms on M, that is, Uk{M) = {u e Ak{M)\ Au> = 0}. Since every harmonic form is closed by Proposition 4.13 (iii), we get a linear map Hfc(M) - HkDR(M) by taking the de Rham cohomology. Lemma 4.15. The map Uk(M) - HkDR{M) is an injection. PROOF. It is sufficient to prove that if a harmonic form u) is exact, then u> = 0. Now if u> = dn, then by Proposition 4.10, we get (w,w) = (cty.w) = (r?,<M = G7,0) = 0. Hence o> = 0. ¦ It follows from the theorem of de Rham that HpR(M) is isomorphic to Hk(M;R)\ hence HqR(M) is finite-dimensional. Combining this and the lemma above, we see that Hk(M) is also finite- dimensional. As a matter of fact, we have indeed the following result.
4.3. THE HODGE THEOREM 159 Theorem 4.16 (Hodge theorem). An arbitrary de Rham coho- mology class of an oriented compact Riemannian manifold can be represented by a unique harmonic form. In other words, the natural map Mk{M) —> H^R{M) is an isomorphism. The essence of this theorem lies in the assertion on the existence of a harmonic form, and existence theorems are generally difficult. A complete proof of the Hodge theorem requires more preparations from analysis and, unfortunately, cannot be included in this book. We shall instead give the reader an outline of the proof. First, choose an orthonormal basis fii,..., 0r for Mk(M) and define the projection H :Ak{M)-+Mk{M) by setting t=i Ifw € Hfc, then Hu - u>. Lemma 4.17. The following three subspaces of Ak(M) are orthogonal to each other: Hk,dAk~1,SAk+l. Therefore we have the direct sum Mk 0 dAk~l © 8Ak+1 C Ak{M), and dAk~i (&5Ak+l is contained in KerH. Furthermore, an element of Ak orthogonal to the direct sum above must be 0. Proof. For u e Ek,r] e Ak~\0 e Ak+\ we have (w.dr?) = (<5w,77) = 0, (w,<S0) = {<L>,9) = 0, {dn,69) = {d\$) = 0. The first half of the lemma has been proved. Next, if u> e Ak is orthogonal to dAk~l, then for any rj e Ak~l we have 0 = {u,dr)) = {5u>, 77) and hence 6u> — 0. Similarly, if u> is orthogonal to <5-4fc+1, then we get du> = 0. Thus if u> is orthogonal to dAk~l 0 &Ak+l, then dw — Su = 0, that is, u) G Mk. If u is also orthogonal to Hfc, then (uj,u) = 0 and hence u> = 0. The second half of the lemma is now proved. ¦ If we could assume that Ak{M) is finite-dimensional, then the lemma above would imply that Ak(M) coincides with the direct sum of the three subspaces. But since Ak(M) is infinite-dimensional, we need hard work before we can actually prove the following result as formulated by Kodaira and de Rham.
160 4. LAPLACIAN AND HARMONIC FORMS Theorem 4.18 (Hodge decomposition). On an oriented compact Riemannian manifold, an arbitrary k-form can be uniquely ivritten as the sum of a harmonic form, an exact form, and a dual exact form; %n other words, Ak{M) =Mk{M)®dAk-\M)®5AkJr\M). Proof of the Hodge theorem based on the Hodge decomposition. It suffices to show that the natural map Uk(M) -» HpR(M) is surjective. Let u> e Ak(M) be any closed form and let oj = (jjh + dn + 59 be the Hodge decomposition of a;. Of course, a>// = Hu. By assumption, we have 0 = aw = d59. Therefore 0 = {d59,9) = {69,59) implies 59 = 0. In this case, we have u> = uh -r dn. Thus cj is cohomologous to the harmonic form o>//, as we wanted to show. ¦ (b) The idea for the proof of the Hodge decomposition. For any ui e Ak(M), u> - Hu is clearly orthogonal to Hfc, and we expect u-HujedAk-x®5Ak+l. We start with any element cjq € (H*)-1 and ask for a solution of the equation for n D.7) A77 = u)q. We know that it has always a solution, say n. Thus for uio = uj — Hu the solution n gives rise to uj-Hu = d{5n) + 5{dn) e dAk~x © 5Ak+\ namely, the Hodge decomposition for u>. We can furthermore modify v. First, let nx = n - Hn e (H*I. Then A^ = Afa - Hr)) = At?l = u> - Hu. Second, it is easy to see that 771 satisfying 771 6 (Mk)± and u) - Hu = Ar/i is uniquely determined. For this reason we can define a map G : Ak{M) - (H*I = dAk~l © 5Ak+1
4.3. THE HODGE THEOREM 161 in such a way that for any u € Ak(M) its Hodge decomposition is given by w = Hu + A(Gw) = Hu + d6{Gu;) + 5d{Gu>). This map G is called Green's operator. Clearly, KerG = Hfc. We also see that G : (H*I = dAk~l © 6Ak+i -> (H*I = dAk~l © 6Ak+1 is bijective and equal to A-1. The projection operator H : Ak —> Mk is self-adjoint and satisfies H2 — H. It also has the properties dH = Hd = 0, 6H = H8 = 0, AH = //A = 0. The proof is left to the reader. We also have Proposition 4.19. Green's operator G commutes with A, d, and 8. The proof is left to the reader. Thus we have seen that the fundamental problem is to solve the equation D.7), which is a typical example of what is called an elliptic partial differential equation of the second order. Several methods that apply to this problem are now available; the reader is referred to the references [deRh] and [Wa]. Here we shall only explain the meaning of the term "elliptic". In general, we consider a partial differential operator of order k that acts on Cm-valued C°°functions defined on Rn ' tj=i... where r- *¦' It is said to be of elliptic type if for every ? = (flf... ,?n) ^ 0 € Rn, the matrix *@)(O= ( X>a(*K° \|a|=fc
162 4. LAPLACIAN AND HARMONIC FORMS where f^ff1 ¦¦¦?-. is non-singular. Let us note that cr(D) depends only on the highest order portion of D. It is called the symbol of D. For example, the classical Laplace operator Yli ~§p ys clearly of elliptic type. We can see that for D to be of elliptic type, it is necessary and sufficient that at every point x the following condition holds: for all functions / : Rn -- Cm(/(x) ^ 0) and all h : Rn -» R (/i(x) = 0,dhx ? 0), we have D(hkf){x) ^ 0. For if we set dhx = ^dxx + • • • + Zndxn, then Da{hkf){x) = k\?af{x) and hence D{hkf)(x) = k\o{D){?)f{x). From this observation we see that ? € T?Rn. In this way, the symbol gives a linear map a(D)(?) : Cm —» Cm for each point. To say that D is of elliptic type means that these maps are isomorphisms for ? ^ 0. The notion of partial differential operator can be generalized so that it acts on sections of a complex vector bundle over a manifold. The definition of ellipticity is also naturally defined; that is, such notions can be reduced, by using local trivialization, to the case of a vector-valued function on a Euclidean space. If we use a formulation without using coordinates as in the discussions above, everything becomes transparent. For example, we easily see that the Laplacian A is a partial differential operator of second order acting on Ak(M) (see Example 5.10). We also see that it is elliptic by showing that the symbol of the Laplacian at p € M a(A)@:Afcr;(M)-Afcr;(M) is an isomorphism for each ? / 0. Just as in the case of functions on Rn we treated, this condition is equivalent to A(h2u>)p ^ 0 for each u) e Ak(M) (with u>p ^ o) and each function h € C^iM) (with h(p) = 0,dhp t? 0). We can verify this by concrete computation based on the definition of the Laplacian. 4.4. Applications of the Hodge theorem (a) The Poincare duality theorem. Let M be a connected, compact oriented n-dimensional C°° manifold. For each k @ < k < n), we define a bilinear map HkDR{M) x HnD-Rk(M) - R
4.4. APPLICATIONS OF THE HODGE THEOREM by setting Jm >-* u> A 77, where u> and 77 are closed k- and (n - fc)-forms, and [u;] and [77] the de Rham cohomology classes represented by uj and 77, respectively. This map is obviously bilinear. That the image is independent of the choice of closed forms representing the de Rham cohomology classes follows from computation for Corollary 3.7 of Stokes' theorem, namely, / Jm A0 + da) A G7 + dE) = / WA77+ / d(af\r}+{-l)kuj A0 + a/\dC) Jm Jm u) A 77. Jm Theorem 4.20 (Poincare duality theorem). For a connected, compact oriented n-dimensional C°° manifold, the bilinear map HkDR(M)xHnD~Rk(M)^R defined above is nondegenerate and hence induces an isomorphism Hl-Rk{M)~HkDR{M)\ Proof. Non-degeneracy of the map means that for any nonzero cohomology class [w] € HdR(M), there exists a certain [77] € Hp~R(M) such that JM u A 77 ^ 0. In order to prove this, let us choose a Riemannian metric. By Theorem 4.16 we may assume that u> is a harmonic form relative to the metric that is not zero identically. If 77 = *u>, then Proposition 4.13 (i) shows that 77 is also a harmonic form, which is closed. Since /. u,A77 = |MIVO, M we conclude the proof. Here we derived the Poincare theorem from the theorem of Hodge. As the name suggests, the essential content of the theorem was found by Poincare. Back then there was no concept of cohomology group, so the result was formulated only within the framework of the homology group of a triangulated manifold.
164 4. LAPLACIAN AND HARMONIC FORMS (b) Manifolds and Euler number. Suppose a figure K is triangulated with cti as the number of i- dimensional simplices. Then the alternate sum is an invariant regardless of the way K is triangulated. Behind this fact there is a long history, going back to Euler in the 18-th century. Based on the work of Betti, Poincare' formulated this result together with the foundation of homology groups. That is, the invariant mentioned above is equal to the alternate sum of Betti numbers, namely XW = ?(-!)'&, Pi=dimHi(K;R). Here x(/C) is usually called the Euler number or Euler characteristic or Euler-Poincare characteristic. For an n-dimensional C°° manifold M, we have x(M) = ?>l)Mim/WM). The next theorem is a simple application of the Poincare duality theorem. THEOREM 4.21. The Euler characteristic of an odd-dimensional closed manifold is 0. PROOF. Although this theorem holds for any topological manifold, we shall prove it for C°° manifolds. It is clear that we have only to prove it for a connected manifold. So let M be a Bn + 1)- dimensional connected closed manifold. If M is not orientable, let M be the set of all pairs (p,<r), where p is a point on M and a is an orientation in the tangent space TVM. Then we can see that M naturally becomes a connected and orientable C°° manifold and the natural projection re : M —> M is a double covering map. Now, using the triangulation of M induced from that of M, we find that X(M) = 2X(M). Therefore we may assume that M is oriented. In this case, by the Poincare duality theorem 4.20, we find that there is an isomorphism C'"W = A(M))*
4.4. APPLICATIONS OF THE HODGE THEOREM 165 for each k. Therefore we get dim HKDR(M) = dim(//?R(M)r = dimg?l-*(M), from which we conclude that 2n+l X(M)= ?(-l)MimflWM) = 0. i=0 (c) Intersection number. Let us now have a second look at the Poincare duality theorem from a different angle. Let M be an n-dimensional, connected, oriented closed C°° manifold. The natural identification Hk(M;R) = {Hk{M)Y induces an isomorphism Hk{M\W) = {HkDR{M\R)*. By composing this isomorphism and the isomorphism due to Poincare's duality theorem we have D.8) Hk{M;m)^H?-Rk{M). Now let N C M be a fc-dimensional, oriented, closed submanifold of M. Then the fundamental class of N determines a A;-dimensional homology class [N] e Hk{M;Z). Denote by [N]* e H%~R{M) the cohomology class that corresponds to [N] by the isomorphism D.8) above. In this case, what shape is the closed (n - fc)-form on M that represents [N}*1 For example, if we take a Riemannian metric on M, we can pick a harmonic form relative to the metric. But it does not necessarily follow that such a form represents the geometric properties of [N\*. As an answer to these questions, it can be proved that for any open subset containing N there is a closed form representing [N)*have that has its support inside U. For the details, the reader might consult [BT]. Next, let N\ and N2 be oriented closed submanifolds of dimensions k and n — k, respectively, in M. Then the number [Ni\ ¦ [N2] = [HiY U [N2]m € HnDR(M) = R is called the intersection number of N\ and N2. This number is actually an integer that is the number of intersections of the submanifolds N\ and iV2 together with signs. See Figure 4.3.
4. LAPLACIAN AND HARMONIC FORMS Figure 4.3. Intersection number By using intersection numbers, we can define an important invariant of a 4/c-dimensional connected, oriented, closed manifold M. The map H2*(Af;R) x H2k(M;R) 3 (x,y) ~ x • y 6 R is a symmetric bilinear map (or a quadratic form) on H2k(M\R), which we call the intersection form of M. If we express this form by a symmetric matrix relative to a basis, the number of positive eigenvalues minus the number of negative eigenvalues is called the signature of M and denoted sign M. Summary 4.1 If the tangent space at each point of a C°° manifold M is given a positive-definite inner product, we say that a Riemannian metric is given on M and that M is a Riemannian manifold. 4.2 For a Riemannian manifold we may identify the tangent bundle and cotangent bundle by using the metric. 4.3 For an n-dimensional, oriented, Riemannian manifold we have a linear operator, called the Hodge star operator, that maps fc-forms into (n — fc)-forms. 4.4 For a Riemannian manifold, one can define a self-adjoint operator A, called the Laplacian or Laplace-Beltrami operator, that generalizes the classical Laplace operator. 4.5 A function / is called a harmonic function if A/ = 0. A differential form a> is called a harmonic form if Au; = 0. 4.6 On a connected closed Riemannian manifold a harmonic function is a constant function.
4.7 For a connected, closed Riemannian manifold, every de Rham cohomology class is represented by a unique harmonic form. This is called the Hodge theorem. 4.8 For an oriented n-dimensional closed manifold, the fc-dimen- sional de Rham cohomology group and the (n-/c)-dimensional de Rham cohomology group are dual to each other. This is the Poincare duality theorem. 4.9 The Euler number of an odd-dimensional closed manifold is 0. 4.10 The signature of the intersection form of a 4fc-dimensional closed manifold is called the signature of the manifold. Exercises 4.1 Show that the space of all Riemannian metrics on a C°° manifold is connected. 4.2 Let D = {ze C;|z| < 1}, H = {z e C,lmz > 0}. H is naturally identified with the hyperbolic plane H2 of Example 4.3. Prove that the map H3z^ — eD 2 + Z is a diffeomorphism and that the Riemannian metric induced on D by this diffeomorphism is given by ¦ 2 _ 4|«fa)' -(i-M2J' This metric is called the Poincar? metric and D the Poincare' disk. 4.3 On an n-dimensional Euclidean space Kn, prove that the gradient of a function / is given by At STdf d 4.4 Show that the volume element vm of a Riemannian manifold (M, g) can be expressed by vm = \Jdet(pij) dxi A • - • dxn
168 4. LAPLACIAN AND HARMONIC FORMS where (U; X\,..., dxn) is a local coordinate system and gij is the local representation of g (or the components of g). Compute also the volume element of the hyperbolic plane H2 (Example 4.3). 4.5 Let M be a compact Riemannian manifold. Assume that X is a vector field on M and n a unit normal vector field on the boundary dM. Show that / divXvM = / (X, n)vdM Jm JdM 4.6 For a function / on a Riemannian manaifold M, prove that A/ = -divgrad/. 4.7 Let M be an oriented closed C°° manifold. Assume that relative to a certain Riemannian metric the exterior product of any two harmonic forms is also a harmonic form. Show that Massey products are all 0. (See §3.5 (b).) 4.8 Let M and N be oriented compact Riemannian manifolds. Introduce inMxiV the direct product orientation and direct product Riemannian metric. Show that for a harmonic form u> on M and a harmonic form 77 on N} the exterior product tx\lj A n^V is a harmonic form on M x N. Here n\ and txi are projections M x N -> M and M x N —> N. Using this result, prove that the map ? tf{^(M) ® #i*(A0 - HkDR(M x N) is injective. As a matter of fact, the map is an isomorphism (called the Kunneth formula). 4.9 Let M be an odd-dimensional compact manifold. Show that X(M) = |x@M). 4.10 Show that sign S2 x S2 = 0, sign CP2 = 1.
CHAPTER 5 Vector Bundles and Characteristic Classes In this chapter we study vector bundles. For the study of dif- ferentiable manifolds, the tangent space at each point plays a fundamental role, as we have seen from Chapter 1 on. Neither vector fields nor differential forms could be defined without tangent spaces. On the other hand, a look at a manifold immersed in a Euclidean space clearly shows that the tangent spaces do not exist in isolation but move smoothly as points move on the manifold. Thus it is natural to consider the set of all tangent spaces put together - this is called the tangent bundle of a manifold. A vector bundle is a generalization. Roughly speaking, a vector bundle is what we get by lining up vector spaces of a fixed dimension for all points of the manifold. For a given manifold, we construct not only the tangent bundle but also other vector bundles in order to study the structure of the manifold itself. For research in the theory of vector bundles the characteristic classes are important because they express, in the language of cohomology, the way the bundles are curved. We put off the study of characteristic classes in general until Chapter 6, and study the characteristic classes of vector bundles in this chapter. 5.1. Vector bundles (a) The tangent bundle of a manifold. Let M be a C°° manifold and consider the set of all tangent spaces at all points of M: TM = (J TPM. p€A/ We define the projection 7T : TM — M 169
170 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES by sending X G TPM to n(X) = p. Of course, we have 7r_1(p) = TPM. The bundle TM is the most fundamental example of a vector bundle, and is called the tangent bundle. Let us verify that TM admits a natural structure of C°° manifold such that the projection n is of class C°°. In the case where M = Rn, TM is naturally identified with the product manifold Rn x Rn. Also in the case where M is a submanifold of Rn, we may write TM = {{p,v) e TRn = Rn x Rn;p € Myv e TPM C TpRn}, from which it is easy to see that TM is a manifold of class C°°. In the general case, we proceed as follows. Let S be an atlas on M and ([/, <p) a local coordinate system belonging to S. We have <p(U) C Rn, where n is the dimension of M. For any tangent vector v E TPU, its image ip.{v) can be written in the form MwHai_ + ... + flB_ Now define a map (lp:7r-1(f/)-<^(^)xRncR2n by setting, for v E TPU, V(v) = MP),a,,...,on) € <p(U) x Rn. Obviously, <p is one-to-one and onto. Now we introduce a topology in TM by declaring that each n~l(U) is open and <p a homeomor- phism. The collection S = {(n~l(U),^(U^)eS} is then an atlas for TM. Furthermore, from the transformation formula for tangent vectors (§1.3, Proposition 1.34) the coordinate transformations are all of class C°°. It follows that TM is a C°° manifold. In the case where M is an n-dimensional complex manifold, the tangent space TPM at each point p € M is an n-dimensional complex vector space, and the bundle TM has the structure of a complex vector bundle of complex dimension n. (b) Vector bundles. With the tangent bundle of a C°° manifold as a model we now define an n-dimensional vector bundle as follows.
5.1. VECTOR BUNDLES 171 Definition 5.1. Let M be a C°° manifold. By an n-dimensional real vector bundle f = (E,n,M) over M we mean that tt : E —* M is a C°° map from a C°° manifold E onto M that satisfies the following conditions: (i) for each p 6 M, 7r_1(p) has the structure of an n-dimensional real vector space; (ii) local triviality: for each p e M there are an open neighborhood U and a diffeomorphism <pu : ir~l{U) = U x Rn such that for each point q € U its restriction to ^~l(q) gives a linear isomorphism: <py : n~l{q) —» {<?} x Rn. If we replace R by C in the above definition we get what is called an n-dimensional complex vector bundle. If n = 1 in the definition, we get what is called a line bundle. For a bundle f, we call E, n, and M the total space, the projection, and the base space, respectively. Also 7r-1(p) is called the fiber over p and is often denoted by Ep. Sometimes, we write ¦n : E —> M or simply E and call it a vector bundle. For any sub- manifold N of My not necessarily an open subset, a diffeomorphism <^n : n~l(N) = TV x Rn that satisfies the condition of local triviality is called a trivialization over N. Now suppose there are given two open subsets UQ and Up with trivializations ipa : it~l{Ua) = Ua x Rn and <pp : tt~1(U0) = Upx Rn. We can see that the composite map <Pa ° Vp1 ¦ (UQ n Up) x Rn ¦-¦ (Ua n Up) x Rn can be written in the form <Pa o<Ppl{p,v) = (p,ga0{p)v) {peUanUp,v6Rn). Here gap : Ua^Up —» GL(n\R) is a certain C°° map that expresses the shift of the two trivializations on UaC\Up. It is called the transition function. If we take another trivialization <^7 on an open set t/7, we obtain transition functions gp^ and ga~,. It is easy to verify that gap(p)9p'r(p) = 9a-y(p) for p € UQnUpC\Uy, which is called the cocycle condition. Conversely, given an open covering {UQ}a?A of M and a family of mappings {ga0}a,0€A that satisfies the cocycle condition, then by patching together Ua x Rn we can construct a vector bundle. We shall give the details in §6.1, Proposition 6.2, Chapter 6. If rr : E —» M and ir' : F —» jV are both n-dimensional vector bundles, a bundle map is, by definition, a C°° map / : E —* F such
172 5 VECTOR BUNDLES AND CHARACTERISTIC CLASSES that the diagram E —L-> F -i I"' M > N f is commutative and such that for each point p € M the map / : Ep —* Ff(p) is a linear isomorphism. In this case, / is a bundle map over /. For example, if / : M —> N is a diffeomorphism, its differential /« : TM —> TN is a bundle map over /. Two vector bundles ?t = (Ei,7Tt, M), i = 1,2, over the same base M are said to be isomorphic if there is a bundle map E\ —» E2 over the identity map of M. In this case, we write f i = ?2 or E\ = E-2- The product M xRn is obviously a vector bundle over M, which we call the product bundle. A vector bundle isomorphic to the product bundle is called a trivial bundle. For example, the tangent bundle TRn is clearly a trivial bundle. We can simply observe that being isomorphic defines an equivalence relation in the set of all n-dimensional vector bundles over M. Given a manifold M, the set of all isomorphism classes is denoted by Vectn(M). For example, if m = dimM, then [TM] ? Vectm(M), where [TM] denotes the isomorphism class of the tangent bundle TM. It is important for the study of M to consider not only the tangent bundle but also various elements of Vectn(M) for varying n. Definition 5.2. For a given vector bundle 7r : E -> M, a C°° map s : M —> E such that 7r o s = id^ is called a section. In other words, a section associates to each p a certain point s{p) € Ep that is of class C°° relative to p. The section s such that s(p) = 0 € Ep for every p is called the zero section. A section s such that s(p) ^ 0 is said to be non-zero or non-vanishing. By using the term section it is easy to explain local trivialization of a vector bundle. For example, giving a trivialization </? : ix~l (U) = U x Rn on an open subset U is equivalent to choosing sections Si : U —> E,i = 1,..., n, over U such that for any point p € U, $\(p),..., sn(p) form a basis of Ep. Such a set of sections S* is called a frame field over U. We denote by T(E) the set of sections of a vector bundle E. We can make it into a vector space by defining addition and scalar multiplication as follows:
5.1. VECTOR BUNDLES FIGURE 5.1. Sections of a vector bundle For s,s' ? T{E) and a € R, (s + s'){p) = ${p) + s'{p)\ {as){p) = as{p). For 5 € T(B) and for / e C°°{M), we can define (/s)(p) = f{p)s(p), making T(B) into a module over C°°(M). Let us remark that if E is a complex vector bundle, we can make T(E) into a complex vector space as well as a module over C°° complex-valued functions. Example 5.3. Sections of the tangent bundle TM of a C°° manifold M are nothing but vector fields on M. Hence X(M) — T(TM). (c) Several constructions of vector bundles. Restrictions and induced bundles Let 7r : E —» M be a vector bundle over a C°° manifold M. For an arbitrary submanifold N, we set E\n = ¦n~1(N) and define a projection 7r : E\n —> AT to be the restriction of 7r to E\n. Then we get a vector bundle over N, which is called the restriction of E to TV. Again start with a vector bundle -n : E —> M and let / : N —> M be a C°° map. Set /^=={(P,«)€MxE;/(p) = *(*)} and define the projection it : /*E —> iV by 7r(p, u) = p. In this way, we get a vector bundle over N (verify this in Exercise 5.1). This bundle is called the bundle induced by / or the pull-back. The natural map f*E 3 (p,u) —» u G E is a bundle map over f : N -* M. In particular, if TV is a submanifold and i the inclusion map: i : N —> M,
174 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES then the pull-back i*S is naturally isomorphic to the restriction E\w, as we can easily see. Subbundles and quotient bundles Let 7r : E —> M be a vector bundle over M. A vector bundle 7r : F —+ M is called a subbundle if F is a submanifold of i? such that, for each point p e M, the fiber Fp is a vector subspace of the fiber Ep of F. Example 5.4. Let Af be a submanifold of M. Then T7V is a subbundle of TM. Now suppose that F is a subbundle of a vector bundle -k : E —¦ M. For each point p € M, consider the quotient subspace Ep/Fp and set E/F= \JEP/FP. p€M We can verify that the natural projection ir : E/F —¦ M is a vector bundle, called the quotient bundle of E by F. (See Exercise 5.2.) If n and m are the dimensions of E and F, then the dimension of E/F is n - m. Example 5.5. By Example 5.4, 7W is a subbundle of TM|^. The corresponding quotient bundle is called the normal bundle of NinM. Example 5.6. We construct a complex line bundle L over the complex projective space CPn as follows. First consider the trivial (n + l)-dimensional complex vector bundle CPn x Cn+1. An arbitrary point I on CPn is a complex line through the origin of Cn+1, that is, a 1-dimensional complex subspace. Now we set L={(^)GCPnxCn+1;z^}. By simple observation we see that L is a 1-dimensional subbundle of CPn x Cn+1 and hence a complex line bundle over CPn. This bundle is sometimes called the Hopf line bundle . By the same construction we get the Hopf line bundle over RPn. Complexification of a real vector bundle An n-dimensional complex vector space is naturally a 2n-dimen- sional real vector space. It follows that an arbitrary n-dimensional
5.1. VECTOR BUNDLES 175 Figure 5.2. Normal bundle complex vector bundle can be regarded as a 2n-dimensional real vector bundle. This corresponds to the fact that GL(n; C) can be realized as a Lie subgroup of GLBn;R). Conversely, given any n-dimensional real vector bundle E we can construct an n-dimensional complex vector bundle T5<g>C by complexifying each fiber Ep (that is, by taking Ep <S> C). This is called the complexification. The situation is similar to the fact that GL(n\ R) is naturally a Lie subgroup of GL(n; C). The definition of a quotient bundle we gave above is somewhat abstract and difficult to understand. By using the idea of a metric in a vector bundle we can obtain a more intuitive picture. For example, when TV is a submanifold of a C°° manifold M, the normal bundle of TV in M (see Example 5.5) can be described as a subbundle of TM\s by making use of a Riemannian metric on M. That is, for each point p e TV, if (TpNI is the orthogonal complement of the subspace TPTV in the space TPM, then U cw1 is a subbundle of TM\^. On the other hand, there exists a natural isomorphism {TpNI = TPM/TPN that gives an isomorphism of the vector bundle above to the normal bundle of TV. (See Figure 5.2.) If we recall the definition of Riemannian metric on a C°° manifold (§4.1 (a)), it will be natural to give the following definition. Definition 5.7. By a Riemannian metric on a vector bundle ¦n : E —> M we mean that each fiber Ep,p G M, is given a positive- definite inner product gp : Ep x Ep —¦ R in such a way that gp depends
176 5 VECTOR BUNDLES AND CHARACTERISTIC CLASSES on p in a C°° manner. We often simplify and say a metric or an inner product. For a complex vector bundle, we use a Hermitian inner product. Here we say that gp depends on p in a C°° manner if the following holds. For a frame field Si : U -» E A < t < n) on an open subset U, 9PMp),Sj{p)) is a C°° function for every pair i,j. The following proposition can be proved in a manner similar to the case of a Riemannian metric (Proposition 4.5). So the proof is omitted. PROPOSITION 5.8. An arbitrary vector bundle admits a metric. Various constructions such as direct sum, tensor product, dual vector space, exterior algebra, etc., for vector spaces can be extended to the case of vector bundles. We shall take up some of them. Whitney sum Suppose two vector bundles 7Tj : Ei —* M, i = 1,2, over the same base space are given. Then the set Ei®E2 = {{uuu2) eEiX E2\iri(ui) =7r2(u2)} with the projection 7r : Ei © E2 3 {u\,u2) »-» tti(ui) € M is the Whitney sum of E\ and E2. If dimEi = niy i — 1,2, then dim(?'i © E2) = n\ + n2. Example 5.9. Let E be a vector bundle and F an arbitrary subbundle. Then there is an isomorphism E = F © E/F. Dual bundle and exterior power bundle Let 7r : E —> M be a real vector bundle. If we set p€M then this is a vector bundle whose dimension is the same as that of E. Here ?* is the dual vector space of Ep, namely, Hom(?p,R). We call E* the dual bundle. If we pick a Riemannian metric, then for
5.1. VECTOR BUNDLES 177 each point p € M there is an induced isomorphism E* = Ep. Hence E* is isomorphic with E. For a C°° manifold M, the dual bundle of the tangent bundle TM is written T*M and called the cotangent bundle. For a complex vector bundle n : E —> M, we get E* = (J Homc(^p,C), which is a complex vector bundle of the same dimension and is called the dual bundle. We should remark, however, that in general E* and E are not isomorphic. This is due to the fact that the Hermitian inner product Ep x Ev —> C is complex-linear in one component but conjugate-linear in the other component. Indeed, the dual bundle L* of the Hopf line bundle L on CPn is not isomorphic to L, as can be verified by using Chern classes in §5.5. From a vector space V we may get the fc-th exterior power AfcV (see §2.1 (c)). By applying this process to each fiber of a vector bundle, we get the fc-th exterior power bundle. Example 5.10. A differential form of fc-th degree over a manifold M of class C°° is nothing but a section of the exterior power bundle hkT*M of degree fc, where we recall that T*M is the cotangent bundle. In other words, we have Ak{M) = r(AfcT*M), as we already stated in §2.1(d). The tangent bundle of the projective space In order to get used to various constructions of vector bundles, we shall now obtain a concrete description of the tangent bundle of the projective space. First we deal with the real projective space RPn. Let L be the Hopf line bundle over RPn from Example 5.6. The usual inner product on Rn+1 x Rn+1 defines a Riemannian metric on the trivial bundle RPn x Rn+1. Since L is a subbundle of this trivial bundle, the orthogonal-complement bundle L1 (with fibers that are the orthogonal complements to the fibers of L) turns out to be an n-dimensional vector bundle over RPn. Now recall that each point of RPn is represented by a certain line t that goes through the origin of Rn+1. Then for the Hopf line bundle L its fiber over the point I, say Le, is I itself, and Lj- is nothing but the orthogonal complement ?L. Now the line t meets the n-dimensional sphere Sn at two points, say,
178 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Figure 5.3. Tangent vectors to the projective space x and -x. In other words, RPn is the space obtained by identification of pairs of points that are symmetric to each other with respect to the origin. Therefore the tangent space TeRPn is obtained by identifying TxSn and T-xSn by the correspondence TxSn 3 v *-* -v e T-xSn. If we regard TxSn and T-xSn as linear subspaces of Rn+1 by parallel displacement, then any tangent vector X 6 T(RPn is expressed by the pair {(x, v), (—x, —v)}, x € 5n, v € ?L. Such pairs induce a linear map by the correspondence ? 3 ax >-* av € tL (a ? R). Conversely, the linear map fx determines a pair {(x,d), (-x, -v)} uniquely. In this way, we may write TeRPn = Hom(?J±). By moving the point ? on RPn we obtain the following proposition. Proposition 5.11. Let L be the Hopf line bundle over the real projective space RPn and let L1 be the orthogonal complement of the subbundle L in the product bundle RPn x Rn+1. Then we have a natural bundle isomorphism TRPn^Hom(LyL-L).
51. VECTOR BUNDLES 179 Corollary 5.12. Let e be the trivial line bundle over RPn. Then there is an isomorphism TRPn © e ? L © • • • © L (the Whitney sum ofn+l copies of L). Proof. For any line bundle f, Hom(?,f) is trivial. In fact, using a non-vanishing section, a trivialization of the bundle can be constructed. Together with Proposition 5.11 this implies TRPn © c ss Hom(L, LL) © Hom(L, L) S Hom(L, L^ © L). Since LL © L is clearly the (n + l)-dimensional trivial bundle, we get Hom(L, L1 © L) ^ Hom(L, e © • • • © e) S Hom(L, c)©---© Hom(L, c). As we have seen before, the dual bundle of a real vector bundle is isomorphic to the original bundle. In particular, Hom(L,e)=*L. Hence TRPn © e S* L © • • • © L, as we wanted to show. ¦ If we apply the argument above to the complex projective space CPn, we get the following. PROPOSITION 5.13. Let L be the Hop/line bundle overCPn and € the trivial complex line bundle. Then there is an isomorphism TCPn © e ^ V © • • • © L* (the Whitney sum of(n+ 1) copies of L*). The difference from the case of the real projective space is that the dual bundle of a complex vector bundle is not necessarily isomorphic to the original bundle. These results, in particular Proposition 5.13, will later be useful for Theorems 5.48 and 5.49.
180 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES 5.2. Geodesics and parallel translation of vectors In this section we review various well-known facts about a surface in R3, and motivate the definition of a connection in a vector bundle that will be given in the next section. (a) Geodesics. The shortest curve that joins two given points in the plane is obviously the segment joining the two points. We obtain the notion of a geodesic naturally if we consider a surface or a more general Rie- mannian manifold. To keep our discussions simple, we shall consider only a surface lying in a 3-dimensional Euclidean space. It will be relatively easy to extend our considerations to the case of a general hypersurface in Rn. Let M be a surface in R3. Then M has an induced Riemannian metric from R3. We shall consider an appropriate condition under which a given curve c : (a, 6) —* M can be thought of as a generalization of a line segment on the plane. We think of riding a certain vehicle moving along c. It is desirable if we can avoid any side-swing and if we can keep constant speed. However, it would be difficult to avoid vertical shift, due to the curved surface. From physics, these things can be expressed in terms of the acceleration vector. First, the velocity vector c(t) of the curve (as motion) is contained in the tangent plane at c{t), that is, 6{t)eTc{t)M, te{a,b). But the acceleration vector c(t) = dc/dt goes out of the tangent plane at c(t) in general. Denoting by Np the orthogonal complement of the tangent space TPM in TPR3, we have TPR3 = TpM®Np. Accordingly we get c{t) = (Dhc)(t) + (Dvc)(t), where Dhc and Dnc denote the components of c{t) in Tp and in 7Vp, respectively. In fact, we have Dh6 = c — {c,n)n, where n is a unit normal field at c(t). Definition 5.14. A curve c:(o,b)-+Mona surface M in R3 is said to be a geodesic if its acceleration vector is perpendicular to the tangent plane to M, that is, (Dhc)[t) = 0.
52. GEODE5ICS AND PARALLEL TRANSLATION OF VECTORS 181 Figure 5.4. Decomposition of the acceleration vector In terms of our ride, if the acceleration vector is always in the normal direction, then there is no side-swing. In this case, it is easy to verify that (c(t),c(t)) = 0, which in turn implies that (c,c), and hence the speed of c(i), is constant. As simple examples, a great circle c on a sphere M is a geodesic in the sense of Definition 5.14; on the other hand, a small circle on M has constant speed but is not a geodesic. (b) Covariant derivative. To continue the preceding discussions, we proceed as follows. A) In the case where we view R3 as a manifold, let X be a tangent vector at a point p. Then for any vector field Y defined in a neighborhood U of p, the covariant derivative Dx Y is the tangent vector at p defined as follows. Let (Y\, V2, V3) be the components of Y relative to any Cartesian coordinate system (x 1,22.?3) in R3. Then DxY is the tangent vector with the components {XY\, XY2, XY$). Here each XYi is the derivative of the function Y{ taken in the direction of X at p. It is easily verified that DxY is independent of the choice of the coordinate system. This idea of covariant derivative is just an extension of differentiation of a function / in the direction of a tangent vector at p. B) Let M be a surface in R3 and let p € M. If X € TPM and if Y is a vector field defined on a neighborhood U of p in M, then we have the covariant derivative DxY at p as in the above, and its
182 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES orthogonal decomposition DxY = {Dh)xY + {Dn)xY, where {Dh)xY € TPM is the eovariant derivative within the surface M and (Dn)XY is a normal vector to M, that is, a vector in the normal space Np. If we are interested only in the tangential component, we denote it by V XY. Note that if X and Y are tangent vector fields on an open subset, then the eovariant derivative VXY makes sense as a vector field on U. C) Generally, given a manifold M, eovariant differentiation is a mapping {X,Y) e X{M) x X{M) ~ VXY e X(M) that satisfies the following properties: (I) Vx1+X3y = vXly + vXly, v/xy = /vxy; (ii) VX{YX + Y2) = VxKi + Vxy2, Vx(/y) = fVxY + (Xf)Y, where / is any differentiable function on M and X,Y>Xi,X2,Y\,Y2 are vector fields on M. PROPOSITION 5.15. Covariant differentiation defined on a surface M C R3 satisfies the properties (I) and (II) above. The proof is left as an exercise to the reader. D) We may further define a somewhat more general concept of covariant differentiation. Let / : M —> M be a differentiable map between two manifolds M and M. By a vector field along the map / we mean a differentiable map Y : M —* TM such that *(Y(P)) = f(p), PeM. We shall denote by 3t/ the set of all vector fields along the map /. In this set-up, suppose M has covariant differentiation denoted by V. Then we can obtain covariant differentiation along the map / which essentially gives rise to the covariant derivative of each vector field Y e Xj relative to a vector field X on M. Actually, the discussion we had on a curve being a geodesic depends on the general concept of covariant differentiation. Suppose c(t) is a curve on a surface M in R3. Then Dh6 is nothing but the co- variant derivative Va/©tc(t), whose vanishing defines a geodesic. We
5.2 GEODESICS AND PARALLEL TRANSLATION OF VECTORS 183 >K> FIGURE 5.5. Parallel transport of tangent vectors may also consider a vector field, say, Y(t) on M along a curve c(t). If the covariant derivative (Va/atY)(t) is identically zero, then we say that Y is parallel along c(t). Definition 5.16. A vector field Y along the curve c(t) on M is said to be parallel along the curve if V^^Y = 0. Thus a geodesic is a curve whose velocity vector is parallel along the curve. (c) Parallel displacement of vectors and curvature. We continue to deal with a curve c(t) on the surface M in R3. How do the tangent planes TPM move and relate to each other as the point p moves on the curve c{t)l If M were a plane, this would be just parallel transporting of a vector at c@) along the curve c. See Figure 5.5. We now define the notion of parallel displacement. For any two points p and q on M, let c(t),0 < t < 1, be a curve joining p — c@) and q = c(l). We define a linear map r : TPM —> TqM as follows. Take any Yq 6 TPM as initial value, construct a parallel vector field Y(t),0 < t < 1, on M along c, and set r{Y0) = ^A). The parallel vector field exists by the unique existence theorem for the solution of the kind of ordinary differential equation of first order mentioned in Theorem 1.41. Explicitly, we can first assume that the curve c(t),0 < t < 1, lies in a coordinate neighborhood U in M with local coordinates u\,U2- Then c{t) = (ui(t),U2{t)) The equation D^Y = 0 (or more generally V^/dtY = 0) can be expressed by using
184 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES FIGURE 5.6. Parallel displacement and curvature the functions $ijk such that 2 where Xi = ¦?-. Now the equation that expresses that Y(t) is parallel along c is dYk 2 -IT + E *«>(«i@.t*2(t))(dtiiM) V*@ = 0 (fc = 1,2). By linearity of this equation it also follows that r is a linear map. It is also isometric. If the curve c(t) is not contained in a coordinate neighborhood, we divide c(t) into a finite number of coordinate neighborhoods and get the result in a piecewise fashion. Along a geodesic con M, the velocity vector c(t) is parallel along c. For any parallel vector Y(t), the angle between Y and c is constant. This leads to a geometric construction of parallel vector fields along a geodesic. For any curve on the surface, parallel displacement depends on the choice of a curve between the two points depending on the way the surface is curved. See Figure 5.6, and observe how parallel displacement behaves depending on whether the surface has positive or negative curvature.
5.3 CONNECTIONS IN VECTOR BUNDLES AND CURVATURE 185 5.3 Connections in vector bundles and curvature (a) Connections. Motivated by the properties of covariant derivative in §5.2, we are naturally led to the following. Definition 5.17. By a connection in a vector bundle n : E —> M over a C°° manifold M, we mean a bilinear map V : X{M) x T(E) -> T(E) satisfying the following conditions: @ V/X.s = /Vxs and (ii) Vx(/s) = /Vxs + (X/M, where / 6 C°°{M),X e X{M),s e T{E). We call Vxs the covariant derivative of s relative to X. We shall see that any vector bundle admits a connection. First consider the product bundle M x Rn. Let x\>...,xn be the canonical coordinates in Rn. We take a frame field (s\}..., sn), where s,(p) = d/dxi, and set Vx^t = 0 (i — 1,... ,n), for every vector field X. For any s — ?V a^ and every vector field X, we set i=i For this connection, V^s is just the partial derivative in the direction of X if s is considered as an Rn-valued function on M. We call it a trivial connection in the product bundle. For an arbitrary vector bundle tt : E —» M, we take a locally finite open covering {Ua}a€A such that 7r-1([/a) is trivial and denote by V° a trivial connection in each n~l(Ua). Let {/Q} be a partition of unity for the covering UQ, and define V*5 = X>Vxs. It is easy to verify that this defines a connection in E. As is obvious from this construction, there are infinitely many connections. Recent research indicates that the study of the space of all connections in a given vector bundle is gaining in importance. We shall state just one property of the space of connections.
186 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Figure 5.7. Space of all connections Proposition 5.18. Let Vj A < i < k) be k connections in a given vector bundle. Then every linear combination X)i=i U^it where ti -H h tk = 1, is a connection. The proof is easy, and is left as Exercise 5.5. We may express the property above as follows. If we consider k connections as k points in the space of connections, then ail rectilinear figures generated by these points as vertices are contained in the space. (b) Curvature. Assume that a vector bundle n : E —* M has a connection V. Then for any vector field X € X(M)y there is an associated covariant derivation V* : T(J5) —» T{E). When V is the trivial connection in the product bundle, Vx is nothing but the action of X on Rn-valued functions. Hence for the trivial connection, from the definition of bracket we have obviously VxVy-VyVx=V[x,y]. For a general connection in an arbitrary vector bundle, the formula above does not hold, but the deviation from the above leads to the following. Definition 5.19. Let V be a connection in a vector bundle -n : E —* M. Then the map that assigns to a pair of vector fields {X, Y} the operator R(X,Y) = i{VxVy - VyV* - V,x,y]} is called the curvature of the connection.
5.3 CONNECTIONS IN VECTOR BUNDLES AND CURVATURE 187 Lemma 5.20. The curvature R has the following properties. For any X, Y € X{M), f,g,h€C°°(M), and s € T{E), we have (i) R{Y,X) = -R{X,Y); (ii) R(fX,gY)(hs) = fghR(X,Y)(s). Proof, (i) easily follows from [Y,X] = -[X,Y]. To show (ii), first assume that f,g are constant and equal to 1. Then E.1) VXVy(/ls) = VX(/lVyS + (Yh)s) = KVxVys + {Xh)VYs + (Yh)Vxs + (XYh)s. Similarly we get E.2) Vy Vx{hs) = Vy(/iVX5 + {Xh)s) = hVYVxs + {Yh)Vxs + (Xh)VYs + {YXh)s. We also have E.3) V\x,Y)(h8) = hV{x>Y]s + ([X, Y]h)s. Subtracting E.2) and E.3) from E.1), we obtain R{X,Y){hs) = hR{X,Y){s). Next, in the general case, we have E.4) 2R[fXtgY) = V/^Vjy - V9yV/x ~ V(/X>9yj = /V*((?Vy) - ^Vy(/VX) - V(/Xl,y] = /((X9)Vy + pVxVy) - g((Yf)VX + /VyVX) - V\fX,9Y)- Now, using the formula in Proposition 1.40 (iv) in §1.4, [fX,gY] = fg\X,Y] 4- f(Xg)Y - g{Yf)X, we obtain V[/x,<,y] = /5V,Xly] + /(*<?) Vy - g{Yf)Vx. Substituting this in E.4), we have 2R(fX,gY) = fg(VxVy - VyVx - V[x,y)) = 2fgR{XtY).
188 5. VECTOR. BUNDLES AND CHARACTERISTIC CLASSES This, together with the discussion of the first half, leads to R(fXygY)(hs) = fghR(X,Y)(s), completing the proof. ¦ (c) Connection form and curvature form. Let V be a connection in a vector bundle n : E —* M', and R its curvature. In this subsection we discuss how we can locally represent V and R by differential forms. Suppose we take a frame field s\,...,5„ G T(Eu), where U is a certain open subset of M. For any vector field X on U, we may write down n i=l with u^X) e C^iU). Since w)(fX) = fu){X) by Definition 5.17 (ii), it follows from Theorem 2.8 in §2.1 that each w] is a 1-form on U. In fact, these n2 1-forms contain all the information on the connection V on U. So, denoting them collectively as " = K), we call u> the connection form of V on U. We might consider oj as a 1-form on U with values in the set gl(n, R) of all real nxn matrices. We now look at the curvature R from the same point of view. For any vector fields X, Y on U, we define Q){X,Y) € C°°{U) by writing R(X,Y)(Sj) = jrn)(X,Y)si. By Lemma 5.20 we have Sl){Y,X) = -Q){X,Y), n)(fX,gY) = f9n){X,Y). Again by Theorem 2.8 we see that each QJ is a 2-form on U. Putting them together, we get a 2-form on U with values in g((n,R). We call it the curvature form. The following theorem describes the relationship between the connection form and the curvature form, and is called the structure equation.
5.3 CONNECTIONS IN VECTOR BUNDLES AND CURVATURE 189 Theorem 5.21. For a vector bundle the connection form uj = (ui'j) and the curvature form Q = (ft*) are related by da; = -wAw + fi. Componentwise, this is <S = -]Tu4.Au;;fc + ft}. Proof. From the definition of the curvature form we get E.5) 2R{X, Y)(sj) = ]? 2Q){Xy Y)Si. i=i On the other hand, the definition of curvature gives us E.6) 2R(X,Y)(Sj) = (VxVy - VyVx - V{XtY])8, = VxCtu)(Y)st) - Vy?W;(X)s,) - J2u)([X,Y})Si i=i i=i t=i = jr{Xu,){Y))Si + JT u*(YH(X)Si t=l i,fc=l - f)(yw}(X)Mi - ? u,${X)u>i(Y)8i - JTuililX.Y])*. t=l t,fc=l i=l Now if we substitute 2u,){xtY) = x*j(y) - ywj(X) -u,;([x,y)), 2u/fc A u/J(X, y) = u>UX)«>${Y) - L>i{Y)u,${X)f in E.6), we get E.7) 2H(x,y)Ei) = 2^{^(x,y) + ^^A^(x,y)}5t. i=l fc=l Now, comparing E.5) and E.7), we obtain the structure equation. ¦
190 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES (d) Transformation rules of the local expressions for a connection and its curvature. Let V be a connection in a vector bundle -n : E —> M. Given two open subsets Ua and Up in M and trivializations <pa:7r-1(C/a) = C/QxRn, <Pp ¦ n~\Up) = ^x Rn, let gap : UQ D Up —> GL(n\R) be the transition function (see §5.1 (b)). We denote by cjq, Cta; up, Sip the connection and curvature forms on Ua and Up relative to the frame fields induced by <pa and <pp. Then PROPOSITION 5.22. We have the transformation rules @ up = g'pWagop + 9llpdgap, (») fy = 9lp^a9ap- Proof, (i) Let slt... ,sn be the frame field on Ua induced by <pa, and ii,..., tn the field induced on Up by <pp . On Ua fl [/^ we have E.8) *j=X>5*' t=i where the components of gap are denoted by <?}. Applying V* to E.8), we obtain n n n E.9) $>(/?)j(x)«fc = J2d9)(x)si + ? $M«)i(*)*. where the components of wa and u>/j are denoted by u>(a)*- and <*>(/?)*. Substituting in t* of the left-hand side of E.9) the expression E.8), with j replaced by k, and comparing the coefficients of Si, we obtain X>(/?)J(X)ri = dg){X) + J2g^(aYk(X). Jt=i k=i Since this holds for arbitrary X and t,j, we can write gQpup = dgap + oJagap.
5.3 CONNECTIONS IN VECTOR BUNDLES AND CURVATURE 191 Multiplying by <?~j on the left, we obtain the desired equation. Now we prove (ii). From Theorem 5.21 we have Qp = dwp + up A u)p. Now we take the exterior derivative of each side of (i). Here functions and 1-forms appear as matrices, but their exterior derivatives can be easily handled by the usual rules. For instance, if we write g for gQp for simplicity, then exterior differentiation of g~*g = J, where / is the unit matrix, gives rise to dg~lg + g~1dg = 0, from which we get dg~l — -g~ldgg~l. Now we can compute as follows: ftp = du>p +u>p Aujp = -g~ldgg~l Awaj + g~ldu>ag - g~lua Adg - g~ldgg~x A dp + {g~lu)ag + g~ldg) A {g~luag + g~ldg) = g~l{dwQ+ujaAuja)g = g~l?lag. (e) Differential forms with values in a vector bundle. Our explanation of connection and curvature in a vector bundle is just about over. Some readers might think the description could be a bit cleaner, particularly in E.9) in the proof of Proposition 5.22. As a matter of fact, we may use the notion of differential form with values in a vector bundle and offer a logically cleaner presentation. First let us deal with a trivial line bundle E = M x R. In this case, a section of E is nothing more than a C°° function and hence T(E) = CCG{M). For a vector field X on M, we may set Vx/ = X/ (feT(E)) and obtain a trivial connection on E. On the other hand, if we use a 1-form df e Al(M), then we can write Vxf = Xf = df(X). Now suppose we are given an arbitrary vector bundle E with a connection V. We think of Vs as a 1-form on M whose value on X € 3t(M) is Vxs, which is a section of E. More generally, we have Definition 5.23. Let E be a vector bundle over a Cx manifold M. By a fc-form on M with values in E we mean a section of AkT*M<2>
192 5 VECTOR BUNDLES AND CHARACTERISTIC CLASSES E. If we denote by Ak(M\ E) the set of all fc-forms with values in ?, then Ak{M;E) = T{AkT*M®E). This is a generalization of the equation Ak{M) = T(hkT*M) stated in Example 5.10 of §5.1. Corresponding to Theorem 2.8 in §5.1, we have a natural identification Ak(M; E) = {X{M) x • • • x X(M) - T{E)}, where the map is alternating and multilinear relative to C°°(M)- modules X{M) x • • • x X{M) and T{E). An arbitrary element of Ak(M;E) can be written as a linear combination of elements of the form 6 <g> s {6 e Ak{M),s € T(E) = A°(M;E)). Also the usual exterior product induces a natural map Ak{M) x Ae{M- E) - Ak+e(M\ E). Now with the definition above let us review connections. First, for any section s G T(E) the map X{M) 3Xm Vxs e T{E) is linear as Coc(M)-modules by condition (i) in Definition 5.17. By writing V s for this map, we can consider it as an element of A1 (M; E). Thus a connection induces a linear map V:T{E)-^Al{M\E) for which condition (ii) for connections appears in the form V(/s) = df ® s + /Vs. Next, we consider the curvature R of a given connection. For any vector fields X, Y, and s 6 T(?), R{X,Y){s) is also a section. We take the fiber EndEp over a point p € M and consider a vector bundle EndE over M with fiber End?p, the set of all endomorphisms of Ep) for each point p € M. Then R induces a map R : X(M) x X{M) - T(EndE). From Lemma 5.20 we see that the map above is alternating and multilinear over C°°(M)-modules. Hence we can write R € A2(M; EndE). Recall that the differentiation d : C°°(M) -* Al{M) can be considered as a trivial connection on M x R. In this case, exterior differentiation Ak(M) —» Ak+1(M) is defined for all k. For a general
5.4 PONTRJAGIN CLASSES 193 connection V : T(E) -» Al{M\E), a linear map D : Ak{M\E) -» Ak+l(M; E) is defined if we set DF <g> s) = d0 ® s + {-l)k6 A Vs F> 6 Ak{M), s € T(E)). For k — 0, we have D = V. We leave it to the reader to verify that the definition of D is independent of the way elements of the form 0®s are written. This differential operator D is called the covariant exterior differential. Unlike the well-known formula d o d = 0, it is not true that D o D = 0. In fact, we have PROPOSITION 5.24. Let V be a connection in a vector bundle E and R G A2(M; EndE) its curvature. Then the linear map R:T{E)-^A2(M]E) coincides with the composite map D o V. The proof is left as Exercise 5.6. 5.4 Pontrjagin classes (a) Invariant polynomials. Let 7r : E —¦ M be an n-dimensional vector bundle. To study its properties we introduce a connection V and get the curvature R. Recall that this means that M has an open covering {UQ} such that V is trivial over Ua and R is expressed by a gl(n; Revalued 2-form fta = (n(a)j) over Ua. Furthermore, the curvature forms Qa are related by E.10) Ct0 = g:l0na9a0. Let us start with these forms and try to define a global form on M. If we succeed, we can integrate such a form on M and obtain a global invariant of E. The key lies in the formula E.10). That is, the forms ?la and H^ are similar to each other in UaC\U0. This means that if we consider only those polynomials in the components of Qa that are invariant by similarity, then we obtain consistent quantities in UanU0, thus getting a globally defined form on M. We state Definition 5.25. If a polynomial function /:gl(n;R)-R
194 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES is invariant by similarity, that is, f(X) = f(A-1XA) (X = (x*)egl(n;R)), / is called an invariant polynomial. The totality of invariant polynomials In is a commutative algebra relative to the usual operations on polynomials. It is easy to see that invariance is equivalent to the condition f{XY) = f{YX) for any two matrices X,Y. Simple examples include detX and TvX\ the former has degree n and is equal to the product of all the eigenvalues (counted with multiplicities), and the latter has degree 1 and is equal to the sum of all the eigenvalues (as well as to the sum of all the diagonal components). It is a well-known result on symmetric polynomials that an arbitrary symmetric polynomial in n variables can be uniquely expressed as a polynomial of elementary symmetric functions ct\,... ,<7„. If we denote by Oi{X) the i-th elementary symmetric function of the eigenvalues of a matrix X, then det(J + tX) = 1 + t<n{X) + t2a2(X) + ¦¦¦ + tnan{X), where / is the unit matrix. By writing this equation it becomes clear that cri(X) is an invariant polynomial. THEOREM 5.26. The algebra In of invariant polynomials is a polynomial ring generated 6y ai,... ,crn, that is, Proof. Let t be the Abelian Lie subalgebra of gl(n; R) consisting of all diagonal matrices of degree n. If the i-th diagonal component of a general diagonal matrix X is denoted by x^ then polynomials in xi,... ,xn can be thought of as polynomial functions on t. Since the restriction of any / € In is a polynomial function on t, it can be expressed as a polynomial of Xi,..., xn, which we denote by p(f). On the other hand, for any i, j and X e t, the matrix X' obtained from X by interchanging the (i,i) component and (j, j)-component is similar to X. Since f(X') = f(X) by definition of invariant polynomial, we have f(X') = f{X). Hence we see that p(f) is a symmetric polynomial. Accordingly, writing Sn for the totality of symmetric polynomials in X\,... ,xn, we see that the correspondence / >-» p(f) induces a homomorphism p:In-*Sn.
5.4 PONTRJAGIN CLASSES 195 By applying an invariant polynomial cr, to diagonal matrices, we see that p(ai) € S» is an elementary symmetric function on xi,..., xn. Therefore we conclude that p is onto. It remains to prove that p is one-to-one. For this purpose, it suffices to show that the map PC : 7„(C) - S„(C) over C is one-to-one. Here 7n(C) is the totality of C-valued, invariant polynomial functions on gl(n;C), and 5n(C) is the totality of symmetric polynomials with n complex coefficients. Then Jn and Sn are subspaces of /n(C) and Sn(C), respectively. Now assume Pc{f) = 0. Then / is 0 on an arbitrary complex diagonal matrix of degree n. By invariance of /, / vanishes on matrices similar to diagonal matrices, in particular, on any upper triangle matrix with all distinct diagonal components. Further, by continuity of /, / is 0 on an arbitrary upper triangle matrix. As is well-known, any square matrix is similar to an upper triangle matrix and hence / is 0 on arbitrary matrices. We have just proved that p is one-to-one. FYom the description above we see that an arbitrary invariant polynomial / e In is uniquely written as a polynomial of <7i,..., an. However, there are cases where it is more convenient to use a basis that we define now. For each i, let us define S; G In by si(X) = TrXi. Obviously, we have p(si) =x\ + -- + x'n. By simple computations we get 5i=^i, s2=aj-2a2, s3 = aj - 3<7i(T2 + 3a3. More generally, it is relatively easy to obtain what is called Newton's formula, namely, 5i-a1Si_i+a25j_2 \-{-l)l~lOi-iSi + (-iyi(Ti =0 (i = 1,. ..,n) Si -<7iSi_i +cr2Sj_2 1- (-l)nanSi-n =0 (t = n+ l,n + 2,...). For the proof, see Exercise 5.7. By induction on i, we see that s, can be written as a polynomial of a\,..., at, and conversely, a, as a polynomial of s\,..., s». Thus we conclude that In — R[si,..., sn\.
196 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Now for any invariant polynomial / € 7n of degree A;, /(fta) is a 2fc-form on Ua and f{?la) — /(fy?) on U^ nfi/j by invariance. Hence we have a globally defined B/c)-form, say /(fi) € A2k(M). PROPOSITION 5.27. For any invariant polynomial f G In of degree k, f(Q) ? A2k(M) is a closed form. Proof. For the proof, let us first compute the exterior differential of the curvature form H = (ft*). By the structure equation (Theorem 5.21), we have H = a\o + u> A u;. By taking the exterior derivative of each side we obtain d?l = du> Au> — u) Adu) = QAw-wAuAu-u)An + wAwAw, and therefore E.11) dn = QAu>-ujAfty which is called Bianchi's identity. In components it is expressed by In view of the discussion following the proof of Theorem 5.26, we have /»SR|«i 4 Thus it is sufficient to show that e//(Q) is closed for the case where / = Si. In this case, we have Si(Q) = Tr(Q,1). Using Bianchi's identity E.11), we compute dsl(Q) = dTr{ni) = TT{dQi) = Tr(dQ A ft*-1 + Q A dQ A ft1-2 + • • • + fii_1 A dQ) = Tr((ftu; - wft) A Q1 + Q A (fiu; - uQ) A fii-2 + --- + ni_1 A(fiw-^fi)) = TV(-wAn, + fi'Aa;)= 0. The last equality follows from componentwise commutativity of the two matrices u> and Q and the fundamental property Tr(XY) = Tr(YX). ¦
5.4 PONTRJAGIN CLASSES 197 (b) Definition of Pontrjagin classes. If we choose a connection V in a vector bundle -n : E —» M, then the curvature form ft of V is defined. If furthermore / is an invariant polynomial of degree k, then a 2/c-form /(ft) is defined. Since it is a closed form, we may consider the de Rham cohomology class [/(ft)] e HlkR{M). Proposition 5.28. The de Rham cohomology class H2^R{M) is independent of the choice of a connection V. Proof. The idea for the proof is simple. Suppose V° and V1 are two connections with curvature forms ft0 and ft1. The natural map 7rxid:ExR->MxR gives a vector bundle over MxR. On the bundle E x 1R we define a connection V as follows. For any arbitrary section s G T(E x R) that is independent of the coordinate t in the direction of R (thus s{p, t) = s{p) as a section of E —> M now regarded as a section of E x R -> M x R), we set (i) V?5 = 0, (ii) Vxs = A - t)V°xs + tVxs, where X e T{Pit)(M) x {*}. An arbitrary section of E x R can be written as a linear combination with functions as coefficients for vector fields independent of t. An arbitrary vector field on M x R can be expressed as a linear combination with function coefficients of the vector field ^ and of a vector field tangent to M x {t}. This means that we can define a certain connection V by the conditions imposed by (i) and (ii). Using the curvature form ft, we get the deRham cohomology class \m)\ € HgR. Using this setup, we define, for e = 0,1, a natural inclusion map it : M —» M x R, where i((p) = {p,e). By the definition of V, we obtain t*ft = ft*. Therefore we get iim) = m°), ii/(ft) = /(ft1). On the other hand, io and i\ are obviously homotopic, and from Corollary 3.15 in §3.3 we obtain im0)} = WW]) = t;([/(n)D = (/(ft1)], completing the proof. ¦
198 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES We have thus found that the cohomology class [f(Ct)] € HffR depends only on E and not on the connection we take. Hence we denote it by f{E) and call it the characteristic class of E corresponding to/. Proposition 5.29. (Naturality of the characteristic class relative to the bundle map). For a bundle map from a vector bundle E into a vector bundle E': E -i— E' ¦1 1* M » M' h we have f(E) = h*(f(E')) e HfR{M). In particular, for any arbitrary C°° map g : N —> M, we have f(g*E) = g*(f(E))eH2DkR(N). Proof. Let V be a connection on E'. Then we can define a connection /i*V on E, called the induced connection by a bundle map, in the following way. If s is a section on E', it naturally induces a section of E, which we denote by h*s. Any arbitrary section of E can be locally expressed as a linear combination of those induced sections where coefficients are functions on M. So for any tangent vector X e TPM we set (h*V)x(h*s)=h*(VhmXs) to define h*V. We leave it to the reader to work out the details showing that a connection can be defined in this way. If u = (u>*-) is the connection form for V, then h*u> = {h*u>j) is the connection form for h*V. Hence if 0. = (ft}) is the curvature form for ?", then h*Q, = {h*Q,lj) is the curvature form of E. The assertion in the proposition follows immediately. ¦ PROPOSITION 5.30. If an invariant polynomial f has odd degree, then[f{tl)] = 0eH2DkR{M).
5.4 PONTRJAGIN CLASSES 199 PROOF. Again from the discussion following Theorem 5.26 we recall that In = R[si,... ,sn]. It is thus sufficient to show that (sfc(ft)] = 0. For this purpose, we introduce a Riemannian metric (see Definition 5.7) in E. Then we shall construct a connection V in E that satisfies the condition E.12) XE,s') = (Vx5,5,) + (s,Vxs/) for all X € X{M) and all s,s' G Y{E). Such a connection V is said to be compatible with the given Riemannian metric, and is also called a metric connection. In order to obtain a metric connection, we use the idea of Gram-Schmidt orthogonalization and get a locally finite open covering {Ua} of M together with a C°° orthonormal frame field S\,... ,sn on Ua, as in §4.1, (c). The trivial connection Va on Ua x Rn satisfies E.12) on Ua- To see this, write n n 5 = Y^at5j, s' = y^6jSj; t=l t=l we have (s, s') = Y^i=i a»^- Therefore we get n E.13) X{s,s') = ]T((Xai)&i + ai(;r&i)). t=i On the other hand, we have n n E.14) V?s = ?(**)*, V?s'= ?(***)*• i=i t=i Therefore we get E.15) (Vaxs,s') = ^(Xoi)fci, E, Vaxs') = ^MX&i). i=l :=1 The equation E.12) for VQ follows from E.13) and E.15). Now by using a partition of unity {/Q} associated to the covering {Ua} with Va we can obtain V = X^/oV01, which is a connection defined globally that satisfies E.12). Next we show that the connection form u> — (w]) on UQ corresponding to V is a skew-symmetric matrix, that is, w] + <jj{ = 0. By
200 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES definition, we have Substituting s = Si,s' = Sj in E.12), we find that fc=l fc=l = wj'(X)+o;j(A:). We also see that the curvature form ft = (ftp is also skew-symmetric, as follows: n ft;=du) + ^wiAwj = -du>l +5^0/,* Au;J fc=i - _^ _ ^Z^i Awf = -nj. Now if X is a skew-symmetric matrix, so is Xfc for any odd integer k > 0, and Tr(JTfc) = 0. Hence sfc(ft) = TY(ft*) = 0, completing the proof. ¦ We are now ready to define the Pontrjagin classes. Definition 5.31. For an n-dimensional vector bundle E —> M, the characteristic class that corresponds to the invariant polynomial 1 r*2fc € In BttJ< is written Pk(E)eH4DkR(M) = H4k(M;R), and is called the Pontrjagin class of degree k. Using the curvature form ft, we may write [d€t G + 2^I = l+Pi(E)+P2(E) + ¦ ¦ ¦ +P[n/2)(E) € Hhi :(M),
5.4 PONTRJAGIN CLASSES 201 or briefly, p{E), and call it the total Pontrjagin class. The closed form that can be written using any particular connection is called the Pontrjagin form. The reader might be a little perplexed about the constant attached to the invariant polynomial for any Pontrjagin class. The reason is that, with our choice of constant, the Pontrjagin class can be well-defined topologically and also matches the Pontrjagin class Pk(E)eH4k(M;Z). (c) Levi-Civita connection. In §5.2 we observed that the tangent bundle of a surface in R3 has a natural connection. The argument there is certainly applicable to a general submanifold of Rn. We shall now deal with the general situation that the tangent bundle TM of an arbitrary Riemannian manifold has a uniquely determined natural connection called the Levi-Civita connection. It is also called the Riemannian connection. We first prove the following proposition. Proposition 5.32. Let M be a Riemannian manifold and U any coordinate neighborhood. Let s\,..., sn be an orthonormal frame field and 01,..., 8n e Al(U) the dual frame field. Then there is a unique l-formu) = (u>*) with values in gl(n\R) satisfying the conditions (i) u){ = -u)), (n) ^ = -e;=i^a^ Proof. First we can find the coefficients ax-k such that n n E.16) d$i = Yl a)k9j A 9k = 2 ? a)k9j A 0k j,k=l j<k subject to oj.• = — a*fc, 1 < i, j, k < n. We also write n E.17) 4=E65^' k=i where bljk are to be determined so as to satisfy the conditions of the proposition. From condition (i) it is necessary to have & = -bw
202 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES From E.17) we have - ? w< a & = - ? b)kek a & = JT b)kv a ek. j=l j.k=l j,k=l This equation must be equal to E.16). Comparing the coefficients of 0i A 9k, we must have E.18) 2a;fc = 6jfc-^. Interchanging i, j as well as i,k, we obtain E.19) 24 = bik- *?., as well as E.20) 24 = ^-6*.. Using 6jfc = -6jfc, we compute E.18)-E.19)+E.20) and obtain Now it is clear that u>* obtained from these bl-k is the desired form. ¦ Now we can define the Levi-Civita connection by using Proposition 5.32. For any coordinate neighborhood U and an orthonormal frame si,..., s„ on U we set i=l where a; = (a;j) is determined by Proposition 5.32. The connection so defined in TU is compatible with the metric if and only if condition (i) is satisfied. We shall now rephrase condition (ii) in Proposition 5.32 so as to make it independent of the choice of orthonormal frame field. In general, a connection V given in a vector bundle E induces a connection V* in the dual vector bundle E*. In the terminology of connection forms, if u> = (u;*) is the connection form for V, then V* has the connection form — lu> = (-u^) relative to the dual frame field. The reader should verify this (Exercise 5.9). In our case, we have
5 4 PONTRJAGIN CLASSES 203 Therefore condition (ii) is satisfied if and only if the composite map T(T*U) = A'iU) ^ T{T*U®T*U) A T{A2T*U) = A2{U) maps each 9l to its exterior derivative d$l. From this we can prove the following: Theorem 5.33 (Levi-Civita connection). The tangent bundle of a Riemannian manifold has a unique metric connection such that the composite map A o V* coincides with exterior differentiation d. The reader may wish to carry out the proof. In a slightly different approach, we can proceed as follows. Suppose that V is a connection in the tangent bundle TM of a manifold M. By setting T(X, Y) = VXY- VyX - (X, Y) (X, Y € X(M)) we see that T is bilinear over C°°(M) and hence it is a tensor field, called the torsion tensor. Now for any differential form 9 of degree 1, the exterior derivative is given by (i) d$(X,Y) = \{X9{Y) - Y9(X) - 9([X,Y})) as in the special case of Theorem 2.9 of Chapter 2. On the other hand, the covariant exterior differential of 9 is given by (ii) V0(X, Y) = X9(Y) - 9(VXY), and its alternation by (hi) i{(v-0)(x,r)-(v*0)(r,x)} = \{X{9(Y)) - YF(X)) - 9(VXY - VYX)} = Ud0{X, Y) - 9{[X, Y] + T{X, Y)} = d9(X,Y)-1-9(T(X,Y)). It follows therefore that d9 is equal to the alternation of the covariant differential of 9 if and only if 0(T{X, Y)) = 0 for all vector fields X, Y on M, that is, if and only if the torsion tensor of V is identically 0. Theorem 5.33 can be interpreted as asserting that, given a Riemannian metric on a manifold, there is a unique, metric linear connection with zero torsion.
204 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES 5.5. Chern classes (a) Connection and curvature in a complex vector bundle. It is possible to talk about connections, curvature, and Chern classes for a complex vector bundle in a parallel fashion to the case of a real vector bundle. However, Chern classes are more important than Pontrjagin classes in that they are indispensable in the study of complex manifolds. Here we will have brief discussions of both, even though there is some repetition. Suppose tv : E —> M is an n-dimensional complex vector bundle. The set of all sections T(E) is not only a module over the ring of all real-valued functions C°°{M) but also a module over the set of all complex-valued C°° functions, which we denote by C°°(M;C). Obviously, C°°(M;C) = C°°(M) <8>C. Also, for differential forms we set Ak{M;C)=Ak{M)®C and call its elements complex A;-forms. By definition, an arbitrary complex fc-form can be uniquely written in the form u + ir) (u;,t? € Ak{M)), where i is the imaginary unit. Exterior differentiation Ak{M;C) - Ak+l{M;C) is defined by simply extending ordinary d linearly over C. The cochain complex {Ak{M;C);d} is called the complex de Rham complex and its cohomology is denoted H*DR{M\C). By virtue of the de Rham theorem we clearly have HhR(M;C) = HhR(M)®C^ /T(M;C). Definition 5.34. Given a complex vector bundle E —> M, a connection is a connection V : 3i(M) x T(E) -» T{E) for the underlying real vector bundle E that furthermore satisfies the condition Vx(ts)=iVxs. The additional condition is equivalent to condition (ii) in Definition 5.17 being satisfied not only for every / € C°°(M) but also for every / € C^iM; C). If we use the description in terms of differential
5.5. CHERN CLASSES 205 forms with values in vector bundles (see §5.3 (e)), we can say that a connection V : T(E) —> Al(M; E) is a complex linear map such that V(/s) = df <g> s + /Vs, for all / € C°°(M;C) and for all s € T{E). The curvature of a connection in a complex vector bundle is defined by using the same formula (Definition 5.19) as in the case of a real vector bundle. Let us now consider the connection form and curvature form. If, in an open subset U, we are given a frame field si,... sn € T(E\u), then writing n vxs; = ;>>}(;o5l (XeX(U)) i=l we get complex 1-forms w* € Al{U;C) over U. Put them together, uj = (wj), and we obtain a 1-form on U with values in gl(n;C); we call it the connection form for V. Similarly, we can get the curvature form ft = (fij) as a 2-form with values in g[(n;C). The structure equation (Theorem 5.21) and the Bianchi identity hold in the same form. The transformation formulas (Proposition 5.22) for the connection and curvature forms remain the same except that the transition functions gap '• UaC]Up —* GL(n;C) now have values in GL(n;C). (b) Definition of Chern classes. A connection V in a complex bundle n : E —> M leads to curvature R; locally, we have the connection form u> = (a;*) and the curvature form Q = (ftp, which are related by transition functions gQp, namely top = 9aptoagap, in any non-empty intersection UQnUp. In order to construct differential forms globally on M, we make the same definition as in the case of a real vector bundle. That is, a polynomial function /:0l(n;C)-C such that f(X) = f{A~lXA)
206 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES for every A G GL(n; C) is called an invariant polynomial function on GL(n\C). The set of all invariant polynomials is denoted by 7n(C). As stated already in the proof of Theorem 5.46, there is an isomorphism /n(C) * 5n(C), where Sn(C) is the commutative algebra of all symmetric polynomials with complex coefficients in n variables. Again by a parallel process we can prove the following sequence of results. For any invariant polynomial / G /n(C) of degree fc, we have /(ft) G A2k(M\C), and it is a closed form (Proposition 5.27); the corresponding de Rham cohomology class [/(ft)] G H2k(M;C) is determined independently of the choice of the connection (Proposition 5.28). We call it the characteristic class of E corresponding to /, and denote it by f(E). The characteristic class is natural with respect to bundle maps; namely, if / is a C°° map g : N —» M and g*E is the induced bundle, then f(9*E) = g'(f(E))eH2k(N-X) (Proposition 5.29). The situation of a complex vector bundle is entirely different from the real case in that Proposition 5.30 does not hold. As we shall see in a concrete example in §5.7(b), the characteristic class corresponding to an invariant polynomial of odd degree is not trivial. Definition 5.35. For an n-dimensional complex vector bundle 7r : E —* M, the characteristic class corresponding to (_±)\€/„(C) is written ck{E) G H2k{M;R) and called the Chern class of degree k. In terms of the curvature form ft we have detG - -Ln)l = 1 + c,(?) + c2(E) + • • • cn(E) G /T(M;R), Ztci J which we call the total Chern class and denote by c(E). A closed form representing each Chern class corresponding to a chosen connection is called a Chern form.
5.5. CHERN CLASSES 207 From our discussions so far, the Chern class Ck(E) is an element of the complex cohomology group H2k(M\C). But as stated in the definition above, it can be defined as a real cohomology class, in fact, as a cohomology class with integral coefficients, namely, Ck(E) e H2k{M;Z). Proposition 5.36. Each Chern class Ck is a real cohomology class. Proof. The proof of this proposition is almost parallel to that of Proposition 5.30, for which we introduced a Riemannian metric in the real vector bundle, chose a metric connection, and made use of the facts that the connection form and the curvature form are both skew-symmetric matrices for any metric connection. In the present case of a complex vector bundle, we make use of a Hermitian metric. We recall here that a Hermitian metric is positive-definite in each fiber Ep and C°° in p € E. We also note that the Hermitian metric is conjugate linear in the second component, namely, (at;, bv) = ab{v, v') (a, b e C, v, v' € Ep). We can introduce a Hermitian metric on E and then construct a connection that is compatible with the metric, that is, X(s, s') = <Vxs, s') + E, Vxs') (X € ?(M), s, s' e T(E)). This can be done by an argument similar to the proof of Proposition 5.30. It is also easy to show that the corresponding connection form u = (u;j) and the curvature form ft = (ft*-) are both skew-Hermitian, namely, u,* + q{ =0, ft* + Ci{: = 0. Now if X is a skew-Hermitian matrix, then I - ^-{X is a Hermitian matrix, so that its determinant is a real number. If we use this fact in the defining equation of each Chern class, then we can easily verify that the differential form defining each d is real. ¦ (c) Whitney formula. The characteristic classes of the Whitney sum (§5.1) of two vector bundles is given by the following Whitney formula.
208 5 vector bundles and characteristic classes Theorem 5.37. (i) // E and F are complex vector bundles, then k c.(EeF)=^ct(E)cfc_t(F), t=0 namely, c{E®F) = c(E)c(F). (ii) // E and F are real vector bundles, then k Pk(EeF) = Y/ME)Pk-l(F), t=0 namely, p(E © F) = p{E)p{F). Proof. We first prove (i). Clearly, T{E®F) = T{E) x T(F). It follows that if V and V are the connections of E and F, then there is a natural direct sum connection V 0 V on E © F. If fi and Q' are the curvature forms of V and V, then the curvature form f2 of V © V is the direct sum matrix of H and Ct': Ho ?)¦ Hence c(E©F) = det[/-^] = det [/ - —ft] det [/ - -±-n'\ , proving (i). The proof of (ii) is similar. ¦ (d) Relations between Pontrjagin and Chern classes. If E is an n-dimensional real vector bundle, its Pontrjagin class p(E) € H*(M;R) is defined. On the other hand, since the complex- ification E (g> C of E is an n-dimensional complex vector bundle, its Chern class c{E <8> C) G H*(M;R) is defined. There is a close relationship between these characteristic classes. (It is possible to use the relations to define the Pontrjagin classes in terms of the Chern classes.)
5.5. CHERN CLASSES 209 PROPOSITION 5.38. Let E be a real vector bundle and ?<8>C its complexification. Then pk(E) = (-l)kc2k(E®C) € H2k(M;Z). PROOF. Our proof depends on using differential forms and is limited to the real case. A connection V in E naturally induces the connection V <8> C. The connection and curvature forms u> and Cl extend to the corresponding forms for V ® C. Therefore we have Pk(E) = [(-^JW@)] = (-1)* [(^JWn)] = (-l)kc2k(E®C). If k is odd, then o2k = 0. Thus we get ck{E <8> C) = 0 € H2k{M; R) (fc = 1,3,...). Next let E be an n-dimensional vector bundle. We may think of it as a 2n-dimensional real vector bundle. How are the Chern classes of E and the Pontrjagin classes of E related to each other? For this purpose, let us define what is called the conjugate bundle - a complex vector bundle of the same dimension, denoted E. On each fiber Epy p € M, we define multiplication by a complex number a + hi € C (a, 6 6 R) by setting (a + bi)v = av — biv for every v € Ep. We take Ep as a new complex vector space and define E = IJ Ep as the conjugate bundle. Lemma 5.39. The conjugate bundle E of a complex vector bundle E is isomorphic to the dual bundle E* of E. PROOF. Introduce a Hermitian metric in E and consider on each fiber the map v € Ep >-> ?{v) 6 E*; ?{v) :u6 Ep ^ l(v)u = (u, v) € C. It is easy to verify that we get an isomorphism E = E*. ¦
210 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES PROPOSITION 5.40. The Chem classes of the conjugate bundle E of a complex vector bundle are given by ck{E) = (-l)kck(E). We have also for the dual bundle ck(E*) = (-l)kck(E). Proof. A connection V for E remains a connection for E. If V has the curvature form ft, then V for E has ft as curvature form. On the other hand, we may assume that ft is skew-Hermitian from the proof of Proposition 5.36. Hence ft = -'ft. By putting this into the definition of the Chem class, we get the formula we want. Combining it with Lemma 5.39, we get the second formula. ¦ Proposition 5.41. Let E be an n-dimensional complex vector bundle. Then, writing pi for Pi(E) and Ci for Ci(E), we have I-P1+P2 + (-l)nPn = A + Ci + C2 + ¦ • ¦ C„)(l -C1+C2 + (-l)UCn)- For example, we have pi=ci2-2c2; pi = c22 -2dc3 + 2c4. PROOF. We write ?r when E is regarded as a real vector bundle. Then J5r <g> C is a 2n-dimensional complex vector bundle, and there is a natural isomorphism ER <g> C ^ E © E. To see this we consider the correspondence for each fiber {Ey^<S>C)p^u + v®i (u -f iv u — iv\ —5-,—2~J eEp®Ep (u,v,iveEp = Ep). Because i(u+v ® i) = -v + u <S> i (-v + iu -v — iu\ .fu + iv u-iv\ we see that the correspondence above induces an isomorphism over C.
5.6. EULER CLASSES 211 By Proposition 5.38 we have pk = (-l)*C2fc(?it ® C). On the other hand, by applying the Whitney formula (Theorem 5.37) to the isomorphism above, we obtain c(?r <g> C) = c(E ®E) = c{E)c{E). Now we can complete the proof by using Proposition 5.40. ¦ 5.6. Euler classes (a) Orientation of vector bundles. We defined the notion of orientation of a manifold based on the idea of orientation of the tangent space at each point, namely, equivalence classes of ordered bases. We can generalize this idea to define orientation for a vector bundle. Definition 5.42. Let it : E —> M be an n-dimensional vector bundle. To define the notion of orientation for E> we go back to Definition 5.1. In addition to condition (i), we assume that for each point p G M the fiber it~1(q) is given an orientation (as a real vector space). Furthermore, in the notation in condition (ii), we assume that the linear isomorphism <pu : it~l(q) —» {q} x Rn takes the orientation in Tt~~1{q) to the fixed orientation in Rn (standard fiber). (In the last statement, it does not matter how the standard fiber Rn is oriented.) When we can orient all fibers Eqyq € M, in the manner above, we say that the vector bundle E is orientable. If E is orientable and oriented, we can reverse the orientations in the fibers and get an opposite orientation for E. The notion of orientation is the same for a C°° manifold M and for its tangent bundle TM. Any complex vector space V, as a real vector space Vr, is orientable and, in fact, it has a natural orientation. If we take any basis v\,..., vn of V, then an ordered basis t>i, iv\,..., vn, ivn for V over R defines an orientation. Moreover, if w\,..., wn is another basis of V over C, then the transformation between vi,ivi,..., vn,ivn and w\,iw\,... ,Wn,iwn has positive determinant. Thus Vr has a natural orientation. It follows that every complex manifold is naturally oriented. (b) The definition of the Euler class. Let 7r : E —* M be a real oriented vector bundle. If we introduce a Riemannian metric in E and select a metric connection, then
212 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES the curvature form is represented by a skew-symmetric matrix. In particular, the Pontrjagin class of the highest degree is given by E.21) pn(E) = jdet (J-njj = (~5^[detn] € H4n(M : R). The conclusion will be that if E is oriented, it is possible to get the "square root"of pn{E); that is to say, we can find a certain cohomology class e{E) = H2n(M;R) such that e(EJ = pn{E). We call e{E) the Euler class of E. In order to find the square root, we start with an alternating matrix X — (xlj) (that is, lX = -X) of degree 2n. For example, if then detX = x2. In general, it is known that if we set where a runs through the permutation group &2n on 2n letters, then detX = Pf(XJ\ Pf{X) is called the Pfaffian of X. It satisfies Pf{T-1XT)=detTPf{X), where T G 0Bn) is arbitrary. In particular, Pf is invariant by T € SOBn), because detT = 1. We shall, later in Chapter 6, consider the algebra of invariant polynomials 1(G) for a general Lie group G. Here Pf € J(SOBn)). By applying the invariance property of Pf to the curvature form, we can construct a 2n-form on the whole of M: Pfip) e A2n{M). More specifically, we proceed as follows. In a neighborhood U of each point, we choose an orthonormal frame field s\,..., sn € T(E\u) with the given orientation. Then for the corresponding curvature form 0 = (ft})> we have
5.6. EULER CLASSES 213 Since P/(ftJ = detfi, we set «*(n) = ~p/m and obtain from E.21) E.22) pn(E) = [eu(ftJ]. We call eu(Q.) the Euler form. As we have shown in §5.4 (a) and §5.5 (b), all differential forms obtained by substituting the curvature in arbitrary invariant polynomials of the general linear groups GL(n; R) and GL(n; C) are closed. We naturally expect the same result for the Euler form and can actually prove it by using the Pfaffian, but we are not going to proceed in that manner, because in Chapter 6 the reader will find that the result above is a very special case of a much more general result (Proposition 6.46). The reader might appreciate the fact that from the accumulation of concrete individual results a powerful general theory is established with a higher point of view. We now define the Euler class e(E) by setting e{E) = [eu(n)]eH2n{M;R); then from E.22) we get pn(E) = e(EJ. That this definition depends neither on the choice of a Riemannian metric nor on that of a compatible connection can be seen as follows. First, suppose a metric g is fixed. If two connections Vo and Vi are metric, then A - t)Vo -f tV\ is also metric. By arguments similar to those for Proposition 5.28, we can complete the proof. Second, if go and pi are two metrics, they can be joined by a family A - t)go + tg\ of metrics (see Exercise 4.1). Using similar arguments as before, we can finish the proof. It should be clear at this point that the Euler class is natural with respect to an orientation-preserving bundle map; that is, for an arbitrary C°° map g : N —> M we get e(g*E) = g'(e(E)) provided we give g*E the orientation induced by E. Actually, it is known, as with Chern and Pontrjagin classes, that the Euler class can be defined over integer coefficients, namely, e(E) €
214 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES H2n(M\ Z). Roughly speaking, if S{E) denotes the set of all vectors of length 1, then the projection ?r : S(E) —> M is an oriented fiber bundle with S2n~l as fiber (see §6.1 for terminology). Since 52n_1 is Bn-2)- connected and since 7r2„_i(S2n-1) = Z, the primary obstruction class is defined as an element of H2n (M; Z) (see §6.2). This is how the Euler class is defined topologically. If n is an odd number, an n-dimensional oriented vector bundle as well as an oriented fiber bundle with 52n-1 as fiber have a topologically defined Euler class in Hn(M;Z). That class, however, has order 2 and is thus a torsion element; it restricts to 0 as a class with real coefficients. (c) Properties of the Euler class. An n-dimensional complex bundle E is a real 2n-dimensional vector bundle with a natural orientation, as we saw in (a). Hence it has both Chern classes and the Euler class, which bear a simple, clear relationship to each other. To emphasize that we look upon E as a real, oriented vector bundle and we denote it by ?r as before. PROPOSITION 5.43. For an n-dimensional complex vector bundle E we have e(Ex) = Cn(E). Proof. The proposition is valid with coefficients in Z, but here we prove the assertion over R by using differential forms. We introduce a Hermitian metric in E, consider a compatible connection V, and denote by R the curvature of V. If S\,..., sn is a local orthonor- mal frame field, then the equation E.23) R(sk) = f2nisi j=i determines the curvature form ft = (ft^), which is skew-Hermitian. On the other hand, we can think of the Hermitian metric on E as a Riemannian metric on Ejt and the connection V as a Riemannian connection in Er. Furthermore, S\,isi,... ,sn>isn becomes a local orthonormal frame field with positive orientation. Now writing and observing that R is linear over C, we get from E.23) R(sk) = a\s\ + 6fct5i + h a%sn + b?isn, R(isk) = -&fcSi + o?isi + 6?sn + a^iSn.
5.6. EULER CLASSES It then follows that the curvature form Qr of ?r is given by -b\ ... aln -bl \ a] ... bi al 1 b) \  \b? -K Here we have e(?R) = [eu(nR)], cn(^) = [det f-^")] • prove is ^P/(fiR) = det(-^n). So the formula to prove is E.24) This is, however, a purely algebraic identity between Pf and det as polynomials with matrix entries as variables. So suppose Q from here on is an arbitrary skew-Hermitian matrix and ffo the corresponding real alternating matrix. Then, as is well known in linear algebra, there is a unitary matrix U such that U~lQU is the diagonal matrix with ib\,..., ibn as diagonal entries. Then we obviously have E.25) det(-^n) .^(-irto) = g)V-6„. On the other hand, if Ur € SOBn) is the orthogonal matrix naturally induced by t/, then Uxln*uR = /0 -&i 6i 0 , 0 Vo 0 bn 0 \ -bn 0 / By invariance of P/, we get E.26) iSF'-'iw-iSr "<".-'">¦*>' (sy Our proof is complete in view of E.24), E.25), and E.26). The next proposition can be proved similarly, and is left as Exercise 5.8.
216 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Proposition 5.44. Let E and F be two oriented vector bundles over a C°° manifold M with dimensions 2m and 2n, respectively. Then the Whitney sum E © F is an oriented vector bundle of dimension 2(m + n), and its Euler class is given by e{E®F) = e{E)e{F) e //2(m+n)(M;R). 5.7. Applications of characteristic classes (a) The Gauss-Bonnet theorem. The following result was originally proved independently by Al- lendoerfer [A] and Fenchel [F] in the case of Riemannian manifolds embedded in a Euclidean space, and by Allendoerfer and Weil [A-W] in the general case. See also Chern's paper [Ch] for a modern proof. Here we shall give a proof from a topological point of view. Theorem 5.45 (Gauss-Bonnet theorem). Let M be an oriented 2n-dimensional closed C°° manifold. Then we have (e(TMI[M)) = X(M). Hence for the curvature form Q of a connection compatible with any Riemannian metric in TM (in particular, for the Levi-Civita connection on TM for a Riemannian manifold M ), we have ( eu(n) = X(M). Jm In order to prove this theorem we prove the following general fact. Lemma 5.46. Let ix : E —» M be an oriented, 2n-dimensional vector bundle. If there exists a section s € T(e) that is never 0, then e{E) = 0. Proof. Introduce a Riemannian metric, and assume ||s(p)|| = 1 at every point by dividing by the length of s. We can easily construct a metric connection V in E such that Vs = 0. Whenever we choose a local orthonormal frame field si,..., s„, we may assume si = s. Then the corresponding connection form has entries 0 on the first row and on the first column. From the structure equation the curvature form Q has the same form. It follows that Pf{Q) = 0, and hence the Euler form is identically 0. Now the assertion is clear. ¦
5.7. APPLICATIONS OF CHARACTERISTIC CLASSES 217 Figure 5.8 Proof of the Gauss-Bonnet theorem. We shall provide an intuitively easier proof at the expense of rigor. Let t : \K\ —> M be a triangulation. We want to construct a vector field X over M that has a zero (also called a singularity) only at the barycenter of each simplex. Let K' be the complex obtained by barycentric subdivision of K and let K" be the result of applying the procedure twice. We define a simplicial map <p:K"-> K' as follows. An arbitrary vertex v of K" lies in the interior of a uniquely determined simplex a € K. We denote this simplex by a(v) and its barycenter by 6<7(v). From the set V(K") to V{K') (the set of vertices of K') we have a map V{K") 3vm <p(v) = ba(v) e V{K'). It is easy to show that this is a simplicial map, that is, for any simplex {vo, ¦ ¦ ¦, vi} of K"y {<p{vo),..., f{ve)} is a simplex of K'. This is the map <p : K" —* K' we want. The continuous map <p : \K"\ = M —> \K'\ is also denoted by <p. The action of the map on the 2-simplex is depicted in Figure 5.8. Around each of the three vertices, it is an expanding map with each vertex as center, and all points near the barycenter go to the barycenter. Also around the midpoint of any of the three edges the behavior is somewhat complicated, that is, points on the segment go to the corresponding midpoint but other points go to the barycenter. The situation appears mostly the same in the case where dimensions of simplices are general. The set of all points that are fixed by
218 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Figure 5.9 ip coincides with the set of all barycenters of simplices of K, namely, the set V(K'). Using this fact we shall define a vector field X. For each point p E M, the segment pip{p) that joins the two points p and (p(p) within a simplex of K has an image by the map t that is a curve through t{p). Now we define Xp to be the velocity vector of the curve at t{p). As parameter of the curve we normalize the length of the segment as 1, independently of p. (In this way, as p approaches a singularity, the length of Xp approaches 0.) By setting Xp = 0 at each point p € V{K'), we obviously have a continuous vector field. However, it is not C°°. Here we need the well-known fact that continuous maps between C°° manifolds, and more generally, continuous sections of vector bundles, can be approximated by C°° objects. In the present situation we may deform X slightly and make it C°° without changing singularities. In the 2-dimensional case, X appears as in Figure 5.9 Next we observe the situation around each singularity. A singularity is the barycenter ba of a certain simplex a G K. Assume dim a = i. If i = 0, that is, if q is a vertex of K, then X is a vector field that diverges away from the center q. If i = 2n, that is, if q is the barycenter of a simplex of K of highest dimension, A" is a vector field that converges toward the center q. For the general case
5.7. APPLICATIONS OF CHARACTERISTIC CLASSES 219 Figure 5.10 0 < i < 2n, X converges in i directions tangent to a and diverges toward the barycenters of some 2n-dimensional simplexes in the remaining 2n - i orthogonal dimensions. Figure 5.11 will show how X behaves. From the discussions above we find that E.27) Yi = -(*, dx\ . + *, —) + xt+1 dxi+i ¦+X2: d ldX2n is the model for a vector field with isolated singularity q. More precisely, as we deform X to be C°°, as we mentioned before, the situation around the singularity looks like the restriction of Yi above. The vector field Yi is the gradient field ^grad/t (see §4.1 (b)) for /i = - -*i+xt •+*2n> - an important function that appears in Morse theory. Actually, we must take orientation into consideration, but this causes no problem since reversing orientation for the model is compensated by reversing the orientation of the vector bundle. Now let us denote by S the set of all singularities of X (previously denoted by V{K')). We shall show that if we take a nice connection in TMy then the corresponding Euler form eti(fi) will be identically
220 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES 0 except in a neighborhood U(q) for each q € «S, and furthermore E.28) f eu(fi) = (-l)*«. JU(q) Here, of course, iq is the dimension of the simplex of K of which q is the barycenter. When all this is proved, we shall get / eu(n) = W cii(ft) = ?>l)l« = ?(-l)dimcr = X(M), Jm qzsJu^ qes aeK which will complete the proof. In §6.2 (f) we shall interpret the quantity in E.28) as the index ind(X, q) for X at the singular point q. The number iq is called the index at the critical point (i.e. the origin) of the standard form of Morse function. We begin by introducing a Riemannian metric in TM. Each point p e S has a neighborhood U(q) that can be identified with a neighborhood of the origin of R2n, say, an open unit disk D(l) of radius 1. We introduce the metric induced by the Euclidean metric of R2n and extend it to M. We set M' = M\[jU(q) qes and modify X so that ||X|| = 1 on M'. Then we introduce a connection V on TM compatible with the metric above and such that VX = 0 on M'. The construction is the same as that in Lemma 5.46. We now take care of E.28) in the 2-dimensional case, the essential result that will imply the general case. There are three vector fields to consider on R2, namely, Vb>^i»^2- First normalize Yq to length 1 and write x d yd r0 — "—I—IT"' r ox r ay where r = yjx2 + y2. We are interested in constructing a connection V on TR2 compatible with the ordinary Euclidean metric and satisfying VVq = 0 away from the origin. Consider the connection form u> relative to ^, ?~. Since u> is alternating (skew-symmetric), we can consider u>\ = -w2 only. In concrete form we can write ^d , d „d , d
5.7 APPLICATIONS OF CHARACTERISTIC CLASSES Then we have *«-'G)«»5 + ;*5+'(?K + ?*1 o*^+H)«l+(^ dy xydx x ,\ d " ^ a-- ay Therefore, we get a unique solution . .1 _ -y dx+ - * x2+y2 x*+y* * Passing to the polar coordinates tan 9 = *, we obtain uj\ = dd. This expression of the connection form is valid away from the origin. The curvature form fi is also alternating, and only Vt\ nas to De dealt with. The structure equation shows that From the theorem of Stokes (Theorem 3.6 in §3.2) we find that / Q\= [ d9 = 2tt, and hence / «»(fi) = / ^p/(fi) = / i-nj = i. JD(\) JD(i)Zn jd(i) Zn So the equality E.28) is valid. For the vector fields Y\ and V2 the computation above goes through with sign changes, and the values are -1 and 1. Hence the proof for the 2-dimensional case is complete. In the general case of dimension 2n we proceed as follows. We have to show that for a connection V on TR2n satisfying the condition Vy/ = 0 away from the origin for Y- — n~nYi> the value of the corresponding integral an,i = / eu{Q) J?>{\) is (-1I. First, the value of the integral does not depend on the choice of V, as we see by using the fact that if connections satisfy the condition above then a linear combination of them (in the sense
222 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Figure 5.11 of Proposition 5.18) also satisfies the same condition. Now for any n, we can prove that a>n,i = (-l)laTi,o> as we already saw in the case where n = 2. For the proof, we use the following fact. For example, the correspondence dx\ dxi defines an automorphism of TR2n (over the identity map of R2n) that preserves the metric and reverses the orientation. The curvature form ft' of the pull-back connection satisfies Pf{Sl') = —Pf(Q). In view of this, it suffices to prove that an,o = 1. Let Z be a vector field on S2, as shown in Figure 5.11. So Z has singularities at the north and south poles only. We take the direct product Mn = S2 x---xS2, and denote by iXi : Mn —> S2 the projection onto the i-th component. Let & denote the pull-back of the tangent bundle TS2 by 7^; then obviously we have TMn =fi0 ••¦0?„. Therefore we can define a vector field Zn by Zn{Pu...,Pn) = Z{Pl)®---(&Z{pn). Clearly, Zn has singularities only at the 2n points that are the north and south poles of each component sphere S2. Furthermore, the
5.7. APPLICATIONS OF CHARACTERISTIC CLASSES 223 vector field Y2j provides a local model for Zn at those singularities. Therefore we get E.29) (e(TMn), [Mn]) = / eu{ty = 2nan>0. On the other hand, by repeatedly using the formula for the Euler class of the Whitney sum (Proposition 5.44) we obtain (e(TMn)y \Mn)) = Gr*e(T52) • • • <e(rS2), [Mn]> E.30) =(e(TS2),[S2))n = 2n. Here we used the already proven fact the for n = 2 the contribution to the integral from each of the north and south poles is 1. We have an.o = 1 by E.29) and E.30), thus completing the proof. ¦ (b) Characteristic classes of the complex projective space. For the tangent bundle of a C°° manifold M of dimension n we write the Pontrjagin classes in the form p(M) = 1 +Pl(M) +p2{M) + ¦ ¦ ¦ +p(n/4)(M) € iT(M;Z). If M is an n-dimensional complex manifold, we write the Chern classes of the complex vector bundle TM in the form c{M) = 1 + Cl (M) + c2(M) + • • • + Cn(M) € H* (M; Z) and call it the Chern class of M. The determination of Pontrjagin or Chern classes is an important step for the study of the structure of the manifold. Here we take up an important example, namely, the complex projective space. First we state simple facts on the cohomology of the n-dimensional projective space. As is well-known, CPn can be decomposed into cells by picking one each for i = 0,2,..., 2n. Thus the homology group (and the cohomology group) is Z at each even dimension and 0 at odd dimensions. We take a generator of the 2-dimensional cohomology group xeH2{CPi;Z)^Z as the one that is 1 on [CPl] e #2(CPn; Z) determined by the natural orientation. In this case, the cohomology ring of CPn is given by i/*(CPn;Z)^Z[x]/(xn+1). This fact can also be derived from what we did, as will be seen in the discussions below.
224 5 VECTOR BUNDLES AND CHARACTERISTIC CLASSES Proposition 5.47. If L is the Hopf line bundle overCPn, then cx{L) = -x. PROOF. It is easy to provide a direct proof starting from the definition. The reader should try (see Exercises 6.5 and 6.6). However, we provide a round-about proof so that we can see relationships between various results obtained so far. We hope it will serve as a good review. Let i : CP1 —* CPn be the natural inclusion; then clearly i*L = L. Hence it suffices to prove the case n = 1, by virtue of the natu- rality of the pull-back of the Chern class. Since C\{L") = -c\{L) by Proposition 5.40, it is now sufficient to show that C\{L*) = x. Setting C\(L*) = kx, we shall show that k = 1. By Proposition 5.13, we have TCP1 © c 5* L* © L*. Hence by Whitney's formula (Theorem 5.37), we obtain dOCP1) = c^TCP1 © e) = ci(L* © L*) = 2kx. On the other hand, Proposition 5.43 and the Gauss-Bonnet theorem imply c^CP1) = e(TS2) = X{S2)x = 2x, which implies k = 1. ¦ Theorem 5.48. The Chern classes of the complex projective space are given by the following formulas: c(CPn) = (l + i)n+1, cfc(CPn) = Proof. By using Proposition 5.13, Whitney's formula (Theorem 5.37), and Proposition 5.47 above we get c(CPn) = c(TCPn © c) = c({n + 1)L*) = A + x)n+1. From this theorem we get what we announced at the beginning of this section, that is, that xn is a generator of Hn(CPn,Z) = Z, as follows. First, from the theorem above we have cry- Cn(CPn) =(n+l)xn.
5.7. APPLICATIONS OF CHARACTERISTIC CLASSES 225 On the other hand, from Proposition 5.43 and the Gauss-Bonnet theorem, we obtain (cn(CPn), [CPn]) = <e(TCPn), [CP"]> = x(CPn) = n + 1. Therefore we have (xn,[CPn]) = 1, completing the proof. THEOREM 5.49. The Pontrjagin classes of the complex projective space are given as follows: p(cp») = (i+x2r+\ Pfc(cp") = (n +k iy>k. PROOF. Substitute TCPn for the complex vector bundle E within Proposition 5.41 and use Theorem 5.48. Then we get 1 - Pl(CPn) + p2(CPn) - • • • = A + x)n+1(l - x)n+1 = A - x2)n+1. Therefore p(CPn) = A + z2)n+1, which completes the proof. ¦ (c) Characteristic numbers. Let M be a 4n-dimensional, oriented closed manifold. While regarding each Pontrjagin class pk as a variable of degree k, suppose there is given a homogeneous polynomial /(Pi.P2,.-.) of degree n. If we substitute Pk{M) in each variable p/c, we obtain an element of the cohomology of M /(p1(M),p2(M),...)G//4n(M;R), which we denote by f{p(M)). The number that is determined as </(p(M)),[M]) is called the Pontrjagin number and simply denoted by f(p(M))[M]. Example 5.50. The Pontrjagin numbers of the complex projective space are easily found from Theorem 5.49. For example, Pi[CP2] = 3, P2[CP4] = 10, p2(CP4] = 25. For a complex manifold we can use the Chern classes instead of the Pontrjagin classes and define the Chern numbers.
226 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES Figure 5.12. Cobordant manifolds Example 5.51. The Chern numbers of the complex projective space are easily obtainable from Theorem 5.48. For example, c1[CP1] = 2, c2[CP2] = 3, c2[CP2] = 9. All these numbers are called characteristic numbers. They represent the global curving properties. Characteristic numbers played a central role in the work of the great topologist R. Thorn. In the early 1950s he used them to create cobordism theory, which is a way of classifying differentiable manifolds. We shall now briefly sketch the ideas of cobordism theory. Definition 5.52. Let M and N be two oriented closed C°° manifolds of the same dimension. We say that they are cobordant if there is a compact C°° manifold W of one higher dimension such that dW = M U -N (see Figure 5.12). Here II denotes topological sum and —N is N with reversed orientation. In particular, if dW = M, we say that M is null cobordant. Proposition 5.53 (Pontrjagin). Two cobordant closed manifolds have the same Pontrjagin numbers. In particular, a null cobordant manifold has all vanishing Pontrjagin numbers. Proof. By definition, the Pontrjagin numbers of —N are the same as those of N with the signs reversed. Therefore it suffices to show that dW = M implies that all the Pontrjagin numbers are equal to 0. Let i: M —> W be the inclusion map. In this case, it is easy to prove that there is an isomorphism i*TW ^ TM © e
5.7. APPLICATIONS OF CHARACTERISTIC CLASSES 227 by considering outward tangent vectors at each point of the boundary of W. It follows that i*p{W) = p{M). By using the Stokes theorem, we find, for any polynomial /, f(pi,P2,..-)[M)= f f(p(M))= ( i*f(p(W)) JM JdW = / df(p(W)) = 0, Jw which concludes the proof. Here p{W) and p{M) are the Pontrjagin forms relative to suitable connections in the tangent bundles TW and TM, respectively. ¦ Prom this theorem and Theorem 5.49 it follows that none of the spaces CP2n is null cobordant; that is, each of them cannot be the boundary of a higher-dimensional, oriented closed manifold. We had no time to discuss the Stiefel-Whitney classes, which are important together with the Chern and Pontrjagin classes. They are cohomology classes with coefficients in the cyclic group Z2 of order 2. For a closed manifold, the Stiefel-Whitney number is defined as an element of Z2. A particularly simple and clear result by Thorn now takes the following form, thanks to contributions by Milnor [M] and Wall [Wal]. Theorem 5.54 ([T], [M], [Wal]). Two oriented closed manifolds of the same dimension are cobordant if and only if all the Pontrjagin numbers and all the Stiefel- Whitney numbers coincide. In particular, a closed manifold is null cobordant if and only if its characteristic numbers are 0. Oriented, 4fc-dimensional manifolds have an important invariant, the signature, that we discussed in §4.4 (c). Hirzebruch, in the same year 1953 when Thorn published his cobordism theory, used it to prove that the signature can be concretely expressed in terms of Pontrjagin numbers. This result is now called the Hirzebruch signature theorem. Its content can be simply explained as follows. Consider *i,..., tk as indeterminates and form the formal power series k n t=i
228 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES It turns out that the homogeneous components are polynomials of the elementary symmetric functions a\,..., crfc of t\,..., ?*. For components of degree k, substitute pt in place of Oi and obtain the polynomial Lfc(pi,... ,pfc), which is called the L polynomial. For example, Ly = -pi, L2 = —Gp2 - p2), L3 = g^F2p3 - 13p2P! + 2p\). Theorem 5.55 (Hirzebruch signature theorem). Let M be a Ak-dimensional, oriented, closed C°° manifold. Then sign M = Lfc(pi,...,pfc)[M]. Example 5.6. Clearly, sign CP2k = 1. On the other hand, by Example 5.50 we have sign CP2 = iPl[CP2] = 1, L2[CP4] = ^Gp2 - p2)[CP4] = 1. So the signature theorem holds for these manifolds. Summary 5.1 A vector bundle comes with a manifold called the base space, to each point of which there is associated a vector space of a fixed dimension in such a way that locally it appears like a direct product of the base space and the vector space. 5.2 The tangent bundle is a vector bundle that is formed by putting the tangent spaces of a C°° manifold together. 5.3 A curve is called a geodesic if its acceleration vector is parallel along the curve. 5.4 To give a connection to a vector bundle E is to define the derivative Vxs of an arbitrary section s of E in the direction of an arbitrary tangent vector X to the base space. 5.5 The curvature is obtained by taking the covariant exterior derivative of the connection, and it measures the curving of the vector bundle. 5.6 A connection and its curvature are locally expressed by the connection form of degree 1 and the curvature form of degree 2 with values in gl(n,R); they are related by the structure equation. 5.7 A characteristic class of a vector bundle associates to the bundle a cohomology class of the base space satisfying the natu- rality condition relative to any bundle map.
EXERCISES 229 5.8 Substituting the curvature form into an invariant polynomial of degree k, we get a closed form of degree 2k on the base space; its cohomology class does not depend on the choice of a connection, and is called a characteristic class. 5.9 The Pontrjagin class is a characteristic class of a real vector bundle, and the Chern class is a characteristic class of a complex vector bundle. An even-dimensional, oriented, real vector bundle also has an Euler class as its characteristic class. 5.10 On a closed C°° manifold, characteristic numbers are obtained by integrating various polynomials of characteristic classes. 5.11 For an oriented, even-dimensional compact Riemannian manifold, the integral of the Euler form is equal to the Euler number of the manifold. This is called the Gauss-Bonnet theorem. Exercises 5.1 Let 7r : E —» M be a vector bundle over a C°° manifold M and / : N — M a C°° map. Set rE = {(p,u)eNxE;f(p) = w(u)} and show that the natural projection ir : /* E —> N is a vector bundle over N. {f*E is called the pull-back of E by /.) 5.2 Let it : E —¦ M be a vector bundle and F a subbundle. Show that the natural projection 7r : E/F = (J Ep/Fp -> M is a vector bundle. 5.3 Give a concrete description of the set Vecti (S1) of equivalence classes of 1-dimensional real vector bundles over Sl. 5.4 Let € be the trivial line bundle over the n-dimensional sphere Sn. Show that TSn © € is a trivial bundle. 5.5 Let Vj,l < i < k, be connections in a vector bundle. Show that for arbitrary real numbers A*,l < i < k, such that Yli=\ ^i = 1> X)i=i ^i^i ls a connection. 5.6 Prove Proposition 5.24; that is, if V is a connection in a vector bundle E, R its curvature, and D : Al(M; E) -* A2{M\ E) is covariant exterior differentiation, prove that i? = DoV. 5.7 Prove Newton's formula (see §5.4 (a)). 5.8 For two oriented vector bundles, show that e(E(BF) = e(E)e(F).
230 5. VECTOR BUNDLES AND CHARACTERISTIC CLASSES 5.9 Let V be a connection in a vector bundle E and u> = (cjj) its connection form. Prove that there is a natural connection V* in the dual bundle B* with -1uj = {-u){) as its connection form. 5.10 A) Find all the Pontrjagin numbers of CP2 x CP2. B) Find all the Chern numbers of CP1 x CP2.
CHAPTER 6 Fiber Bundles and Characteristic Classes It is hoped that the description of the preceding chapters has helped the reader understand that the tangent bundles play an important role in the analytic study of the structures of C°° manifolds. The tangent bundle has a unified structure in the arrangement of one vector space for each point of the manifold. As we discussed in detail in Chapter 5, we obtain the notion of a vector bundle by generalizing the tangent bundle. The way this bundle is curved can be expressed by the Pontrjagin and Chern classes. In this chapter we further generalize the notion of a vector bundle to that of a fiber bundle. A fiber bundle, roughly speaking, assembles one manifold for each point of the manifold. Of particular importance is a principal bundle, which assembles Lie groups that are controlled by the standard Lie group itself. The main theme of the chapter is the Chern-Weil theory, which is also the final goal of our book. This theory describes the way a principal bundle is curved by using the ideas of a connection and curvature in the language of de Rham cohomology. This is a general theory with a wide view, and it includes the characteristic classes of vector bundles. 6.1. Fiber bundle and principal bundle (a) Fiber bundle. Let F be a C°° manifold. A very simple example of copies of F attached to all points of another manifold B is the product B x F. If we denote by n : B x F —> B the natural projection, then for each b € B we have n~l(b) = F. We call this structure a product bundle. In general, we may fairly freely change the way we arrange copies of F, and that means we can construct various different figures. We give 231
232 6 FIBER BUNDLES AND CHARACTERISTIC CLASSES Definition 6.1. Let F be a C°° manifold. Suppose there are given C°° manifolds E and B and a C°° map tt : E -» B. We call ? = (E,-n,B}F) a differentiable fiber bundle (or a differentiable F bundle) if it satisfies the following condition: (local triviality) For each point b of B there are an open neighborhood U and a diffeomorphism <p : tt~1(U) = U x F such that for an arbitrary u € n~l(U) we have tt(u) = -nio<p(u), where -n\ : ?/ x F —> ?/ denotes the projection onto the first component. We call E the total space, B the base space or base, F the fiber, and -n the projection. We call Et, = tt-1F) the fiber over b. Instead of (?,7r,?,F) we may call n : E —> B or simply ? a fiber bundle. In the definition above, E, B, F are C°° manifolds. We may just assume that they are topological spaces and n is a continuous map and if a homeomorphism; then we obtain the definition of a fiber bundle in general. But in this book we shall always work with differentiable fiber bundles. Let & = (Ei,iTi,Bi,F) (i = 1,2) be two fiber bundles with the same fiber. By a bundle map from ?i to ?2. w« mean C°° maps / : E\ —> E2, f : B\ —¦ ?2 such that the diagram Ei * Ei -1 1- Bi > ?2 / is commutative (that is, fl^o/ = fom) and such that the restriction of / to an arbitrary fiber 7r—1F),6 € B\, is a diffeomorphism. If furthermore / is a diffeomorphism, so is / (and vice versa). Also (/-1, /-1) is a bundle map. The proof is easy. Two fiber bundles & = (Ei, 7rt, ?, F) over the same base space B and with the same fiber F are said to be isomorphic if there exists a bundle map / : Ei —> ?2 together with the identity map / : B —> ?. We write ?i = ?2- A bundle that is isomorphic to the product bundle B x F is called a trivial bundle. Let ? = (?\ 7r, B, F) be a fiber bundle. For any submanifold M of the base ?, the collection ^|M = Gr-1(M),7rU-1(M))M,F)
61. FIBER BUNDLE AND PRINCIPAL BUNDLE 233 is also a fiber bundle with fiber F. Here 7r|^-i(M) is the restriction of it to 7r_1(M). We call ?|m the restriction of ? to M. We also write E\m for 7r_1(M). If there exists an isomorphism E\m = M x F, we say that ? is trivial over M. By a cross section or simply a section of ? = (?, 7r, #, F) we mean a C°° map s : B —> E such that 7r o s = id. In other words, it is a map that associates to each point 6 of B a point in the fiber over 6. Whether a given fiber bundle has a cross section or not often becomes an important question. (b) Structure group. Let ? = (?\7r, B,F) be a fiber bundle. By definition, there exist an open covering {UQ} of the base space B and a trivialization Then the map ifia o if~l : (Ua nU0)xF* (Ua HU0)xF gives an isomorphism of the trivial F bundle over (UQ C\ Up). Assume be UanUC and p € F. Then <pQ o <p~l(b,p) is of the form (b,q), where q can be written as gap(b)(p). This means that there exists a map 9ap :Uar)U0-> DiffF such that ^oV?-1F,p) = F,9Q/3F)(p)) {beUanU0,PeF). Here gap is differentiate in the sense that the map (Ua nU0)xF3 (b,p) ~ 9ap(b)(p) e F is C°°. We call gQp the transition functions of the fiber bundle. The family of transition functions gap clearly satisfies F-1) 9a0{b)9fiy(b) = W>) (beUQr)Up(lUy), called the cocycle condition. Conversely, we have Proposition 6.2. Given an open covering {UQ} of a Cx manifold B and a family of differentiate functions {gaff} satisfying the cocycle condition F.1), there is a fiber bundle ? = (E,tt,B,F) with B as the base, with F as the fiber, and with gap o.s the transition functions.
234 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Proof. From the cocycle condition F.1) we find that 9*a(b) = id, 9(ja{b) = gap{b)-1. Let us assume that the open covering {Ua} is indexed by A = {a}. For each a € A consider Ua x F and let E be the topological disjoint union: UaUQ x F. Thus an arbitrary element of E can be represented .in the form (a,6,p), where a C A,b e Ua> p € F. We introduce an equivalence relation in E by saying that two elements (a,b,p) and (/?, c, q) are equivalent if and only if b = c and furthermore p = 9ap{b){q)- We denote by E the quotient space of E with respect to the equivalent relation above. For each a, Ua x F is naturally a subspace of E. It is easy to see that E admits a C°° structure such that Ua x F is an open submanifold. The projection n : E —* B defined by ir(a,b,p) = b is a C°° map. Now we see that ? = (F,7r,B, F) is the fiber bundle satisfying the conditions. ¦ We have seen that an arbitrary F bundle can be constructed by pasting together the product bundles Ua x F by means of the transition functions. Here we may use any element of the group Diff F of all difFeomorphisms of F; but Diff F is too large. Thus it is necessary to use smaller subgroups of DiffF. Most important are the cases where a certain Lie group G acts on F, namely, the cases where G is a Lie group whose action on F is a natural C°° map G x F —> F. For example, SO{n + 1) C DiffSn. Definition 6.3. Let (F,7r,B,F) be a fiber bundle. Suppose B admits an open covering {Ua} and each Ua admits a trivialization V?Q : 7T_1 (Ua) -+[/QxF. Furthermore, assume that the corresponding transition functions gap —> DiffF define C°° maps of UaC\Up into a Lie transformation group G C DiffF. In this case, we say that {Ua,<pa} defines a G structure in (E,tt,B, F) with structure group G. We denote the fiber bundle by (F, 7r, B, F, G) in this case. We make the following remark. First, suppose a family of transitions gap defines a G structure. We say that a trivialization </? : ¦n~l{U) —* U x F is admissible or compatible with the G structure if there is a map ga : U f) Ua —¦ DiffF such that <p o <p~l (b,p) = {b,9a{b){p)) {b € U fi Ua,P € F) and such that the image of every gQ is contained in G and the map ga : U nUa —* G is C°°. If we adjoin an admissible trivialization {U,<p), we get the same G structure. Thus we can talk about a maximal family of transition functions for a given
6.1. FIBER BUNDLE AND PRINCIPAL BUNDLE 2 33 G structure. This situation is analogous to what we have seen about a Cx structure by means of an atlas of charts. Various definitions pertaining to fiber bundles can be adapted to fiber bundles with structure group. We will briefly discuss them. Let ? = (?'t,7rt,Bj,F,G) {i = 1,2) be two fiber bundles with the same fiber F and the structure group G. By a bundle map from ?1 to ?2 we mean a bundle map as F bundles: -1 1- Si > Bi f that satisfies the following condition. If p : Ti]{U) ~* U x F is an arbitrary admissible trivialization and if xp : n^*(V) —> V x F is an arbitrary admissible trivialization, then, y\> o / o <f>~x being a bundle map from the product bundle (?/n/-1 (V)) x F to V x F, there exists a map h : U n f~l{V) -* DiffF such that Tpofo<p-l(b>P) = (f(b),h(b)(p)) (beUnf-](V),P€F). The condition we require is that the image of /i lies in G and the map ft : U n /-! (V) -* G is of class Cx. Suppose two bundles & = (jE^tt.^.F.G) B = 1,2) have the same base B, the same fiber F and the same structure group G that admits a bundle map / : E\ —> F2 over the identity map of B. Then we say they are isomorphic, and write ?1 = ?•>• Next let ? = (F, 7r, B, F, G) be a fiber bundle with structure group G. If the images of all transition functions pa,g lie in a Lie subgroup H of G, then we may regard ? as a fiber bundle with structure group i/. In this case, we say that the structure group G of ? is reducible to H. In this terminology we may state that a fiber bundle ? is trivial if and only if the structure group is reducible to the trivial subgroup (i.e. the identity subgroup). In the same way as Proposition 6.2 we may prove PROPOSITION 6.4. Given an open covering {?/«} °f a Cx manifold B and a G°° family of functions gatj : U0 O Uj -* G satisfying the cocycle condition F.1), we can construct a fiber bundle ? = (E, n, B, F, G) urith structure group G whose transition functions are exactly the given {gQp}.
236 6 FIBER BUNDLES AND CHARACTERISTIC CLASSES In this proposition, there is no condition imposed on the fiber F. If G acts on a different manifold F' as a Lie transformation group, then we have a fiber bundle ?' = (E,n,B,F',G) with the same transition functions but different fiber. In this case, we say that ? and ?' are mutually associated bundles. The most important associated bundle is the one with G as fiber on which the group G acts naturally (as left translations) on G. This is the idea of a principal bundle in the next subsection (c). As an application of Proposition 6.4 we shall define the notion of induced bundle, an important method of constructing new bundles. Let ? = (E, 7r, B, F, G) be a fiber bundle, {Ua} an open covering, y?Q : n~1{Ucx) -* Ua x F a trivialization over UQ, and gQp : UQ f) Up —> G the transition functions. Now let / be a C°° map M —> B. In this case, {f~l{Ua)} is an open covering of M. The map gap o / : f~l{Ua) n f~l{Up) -+ G is of class C°° and satisfies the cocycle condition. By Proposition 6.4, there is a fiber bundle over M with {pa/?} as transition functions. This bundle is called the induced bundle or the pull-back, and is denoted by /*(?)• By definition, there is a natural bundle map /*(?) —> ?. A concrete description is given by noticing that the total space f*E of /*? can be put in the form rE={(p,u)eMxE;f(p) = n(u)}. A detailed verification is left as Exercise 6.2. (c) Principal bundle. Among the fiber bundles a principal bundle plays the most important role. Definition 6.5. Let G be a Lie group. Then a fiber bundle (P, 7r,M, G,G) with fiber G and structure group G is called a principal bundle if the action of G on itself is left translation, that is, La : x —» ax, where a,x 6 G. It is also called a principal G bundle. Here we used P instead of E to emphasize " principal bundle". We shall use (P,rr, M,g), n : P —> M, or simply P, to denote a principal bundle. We have Proposition 6.6. Let? = (P, it, M, G) be a principal G bundle. Then we can define an action of G on the total space P to the right, that is, a map P x G 3 {u,g) ¦-» ug € P such that {ug)h = u(gh), where g,h e G. This action takes each fiber onto itself, and is free,
6.1. FIBER BUNDLE AND PRINCIPAL BUNDLE 237 that is, if ug = u for some u € P, then g = e, the identity of G. Further, the quotient space P/G is identified with the base space M. Proof. We define the action PxG-»Pas follows. Let {UQ} be an open covering of M with trivialization <p : ir~l(Ua) —* Ua x G, and gap • Ua n Up —* G the transition functions for the G structure. Now let u e P and g € G. Pick a such that n(u) € UQ, and assume that <pa(u) = (p, h) {p = ir{u),heG). We set ug = <^a1(P»/lP) and show that this element is independent of the choice of a such that ¦n{u) € UQ. It follows that the action of G on P to the right is defined, and the rest of the proposition is easy to prove. Now going back to the definition of ug above, suppose n(u) is also contained in Up, and set <pp{u) = {p,h'). Then by definition of transition functions we get h' = gpa(p)h and also ^>p{ug) = (p,gpa(p)(hg)). By associativity of multiplication in the group G we have 90<*{p){hg) = {gpa{p)h)g = h'g. Hence ug = Vpl{p,tig), which proves our assertion. ¦ Conversely, suppose we are given a C°° map tt : P —> M and right action of G on P. Assume that for any point p € M there are an open neighborhood U and a diffeomorphism <p : n~l(U) = U x G satisfying n{ug) = 7r(u), <p(ii0) = <p(u)$ (u € fr (?/),$ € G). Then we can make (P,n,M,G) into a principal G bundle. All this can be taken as another definition of principal G bundle. PROPOSITION 6.7. For a principal bundle to be trivial it is necessary and sufficient that it admits a section. Proof. Let ? = (P,7r,M,G) be a principal G bundle. If it is trivial, it clearly admits a section. Conversely, suppose a section s : M —¦ P exists. By Proposition 6.6, G acts on P to the right. From this construction of the action we see that, given two points u and v on one fiber 7r_1(p),p € M, there is a unique element g € G such that v = ug. Now define a map / : P —¦ M x G as follows. For each point u € P consider sGr(u)); the two points u and s(tt(u)) lie
238 0 FIBER BUNDLES AND CHARACTERISTIC CLASSES on the same fiber. Hence there is a unique element g € G such that u = s{n(u))g. Now set f{u) = (n(u),g). It is easily checked that / is an isomorphism as principal bundles. ¦ Example 6.8. Let t\ : P -* M be a principal G bundle. The bundle induced by the projection ir* P —> P is trivial. In fact, P 3 u »-> (u,u) e 7r*P is a section. (See Exercise 6.2.) (d) The classification of fiber bundles and characteristic classes. Given two Cx manifolds F and ?, it is a fundamentally important problem to classify all isomorphism classes of all fiber bundles with base B and fiber F. It is in general an extremely difficult problem. Complete solutions for an arbitrary manifold B are known only in a couple of cases including F = Sl, as we discuss in the following subsection. As is shown in Theorem 6.22, S1 bundles have Euler classes as characteristic classes and as complete invariants. We now define for an arbitrary F the characteristic classes of a fiber bundle with fiber F. Definition 6.9. Let A be an abelian group. Suppose for an arbitrary fiber bundle ? = (E,n,B,F) with fiber F an element a(?) of the cohomology group Hk(B\ A) of the base B with coefficients A is defined and is natural relative to a bundle map in the following sense. Then a(?) is called a characteristic class of the F bundle (of degree k with coefficients .4). Here naturality relative to a bundle map means that for any bundle map between two F bundles &(?»>^i, Bx, F) (i = 1.2) E) > ?2 ?i » B2 f we have c*(?,) = /•(<*(&))• By definition, any two isomorphic F bundles over the same base space have the same characteristic classes. However, the hope of classifying fiber bundles by characteristic classes is difficult to fulfill, due to the fact that Diff F is essentially infinite-dimensional. On the other hand, if we restrict structure groups to Lie groups, we get a relatively satisfactory theory concerning the classification
6.1. FIBER BUNDLE AND PRINCIPAL BUNDLE 239 of fiber bundles and the construction of characteristic classes. The characteristic classes of vector bundles (Pontrjagin classes and Chern classes) are typical examples. In the remaining part of this chapter we use differential forms to build the theory of characteristic classes of fiber bundles whose structure groups are Lie groups. Let us briefly mention the classifying space of fiber bundles. It is a little removed from the main theme of our book, but merely knowing that it exists might be useful. Let G be a Lie group. Then there exists a principal G bundle tt : EG —> BG, called a universal G-bundle, for which the following is valid. Theorem 6.10. Let M be a C°° manifold and ? an arbitrary principal G bundle over M. Then there is a differentiable map f : M —» BG, unique in the sense of homotopy, such that the pull-back by f of the universal G bundle is isomorphic to f. Hence there exists a one-to-one natural correspondence between the set of isomorphism classes of principal G bundles ocuie M and the set of all homotopy classes \M, BG) of the maps of M into BG. The space BG above is called the classifying space. As a matter of fact, BG and EG are not ordinary but infinite-dimensional manifolds, although we do not go into the matter. We only mention that the characteristic classes of a principal G bundle are nothing but elements of the cohomology group of the classifying space BG. (e) Examples of fiber bundles. Example 6.11 (Vector bundle). A real n-dimensional vector bundle is a fiber bundle with fiber Rn and structure group GL(n,R). Similarly, an n-dimensional complex vector bundle has fiber C" and structure group GL(n,C). By introducing a metric in each vector bundle, we may reduce the structure group to 0(n) and ?/(n), respectively. Example 6.12 (The tangent frame bundle of a Cx manifold M). Let M be an n-dimensional manifold. By a frame at p € M we mean an ordered basis [v\,..., vn] of the tangent space TPM. Denote by Fp the set of all frames at p. We also consider the space F(M) of all frames at all points of M. The general linear group GL(n; R) acts on each fiber Fp on the right: FpxGX(n;R)-»Fp;
240 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES namely, for a frame u = [v\,..., vn] e Fp and for a regular matrix g = (gi}) € GL(n\R) we set wi=5Z^iVJ' U0= [wi,...,twn]. 3 In particular, this action is free (that is, if up = u for some u G Fp, then g is the identity matrix) and the orbit space is a single point. The action F(M) x GL(n;R) —> F(M) is also free, and the quotient space is M. The natural projection -n : F(M) —¦ M is defined by 7r(tz) = p if u € Fp. It is easy to check that F(M) is naturally a C°° manifold as in the case of TM. That is, if ((/, xi,..., x„) is a local coordinate system and if for a frame u = [vi,..., vn] we write each V{ G TPM in the form ^ = L^ax~. j j and define the map tt-1(U)->UxGL{ti-R) by u € Fp —> (p, (pij)) € C/ x GZ/(n;R), then it is obviously bijective. Now it is easy to check that it : F(M) —¦ M is a principal bundle with structure group GL(n; R). We call it the tangent frame bundle. In other words, it is a principal bundle associated to the tangent bundle. If M is given a Riemannian metric, we can get a principal bundle with s structure group 0{n) by considering only orthonormal frames. If M is furthermore oriented, we get a principal bundle with structure group SO(n) by considering only positively oriented orthonormal frames. By taking a general vector bundle other than the tangent bundle we can obtain its associated principal bundle. Example 6.13 (Covering manifold). The covering map ir : N -* M of manifolds (see §1.5 (d)) becomes a fiber bundle with zero- dimensional manifold as fiber. In particular, the universal covering 7r : M —> M is a principal bundle with zero-dimensional Lie group 7TiM with discrete topology acting on M as universal covering group. 6.2 S1 bundles and Euler class Before we begin the theory of characteristic classes for general principal bundles we shall discuss the details on the Euler class for
6 2 S1 BUNDLES AND EULER CLASS 241 an S1 bundle, which is the starting point of all characteristic classes, so to speak. (a) S1 bundle. Definition 6.14. By an S1 bundle we mean a fiber bundle with fiber Sl. If we can orient each fiber in such a way that locally we get a unique sense, then we say that the Sl bundle is orientable. If we have chosen an orientation, we say that S1 is oriented. The structure group of S1 bundles is DiffS1 and that of an oriented S1 bundle is the subgroup Diff+S1 of DiffS1 consisting of all orientation-preserving difFeomorphisms of 5l. An S1 bundle that is not orientable is, for example, the so-called Klein bottle obtained from the cylinder Sl x I by identifying both boundaries with orientation reversed. An important example of oriented bundles S1 is, for example, a principal Sl bundle. Here S is the Lie group U(l) of all complex numbers of absolute value 1, which can also be expressed as the rotation group 50B) of the plane around the origin. PROPOSITION 6.15. Every oriented S1 bundle admits the structure of principal Sl bundle. PROOF. Let n : E —> B be an oriented S1 bundle. By introducing a Riemannian metric in the total space E we can induce a metric on each fiber and hence define the length of a curve along a fiber. For each point b G B we denote by ?b the length of the fiber Eb = n~l F). Now we define the right action of Sl on E} ExS1 3{u,z)^uze E{ze S1), as follows. If 7r(u) = 6 and z = ex6f then uz is the point we reach when we proceed in the direction of orientation a distance -^(.b on Et, from u. It is simple to verify that this action gives rise to the structure of principal Sl bundle. ¦ PROPOSITION 6.16. An oriented Sl bundle is trivial if and only if it admits a section. Proof. Direct from Propositions 6.15 and 6.7. ¦ (b) Euler class of an Sl bundle. In order to see if a given S1 bundle ? = (P,7r,M, S1) is trivial or not by appealing to Proposition 6.16, we shall try to construct a
242 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Figure 6.1. Constructing a section section. We consider a C°° triangulation t : \K\ —¦ M of the base M. From the general theory of fiber bundles, it is known that a fiber bundle over any contractible base is trivial. Therefore the restriction of ? to any simplex of K is trivial. This fact is understandable without the general theory by subdividing K sufficiently so that each simplex is contained in an open subset of an open covering giving local trivialization of f. Thus, at each vertex of K, pick any point from its fiber and call it s(v) 6 Pv. Next, for each 1-simplex |t>oi>i| use the trivialization vr-1(|voVi|) = |^oviI x Sl to choose a section s : |vo^i| —> n~1(\voV\\) C P that coincides with the previous one at both ends of \voV\ |. In this way, we can get a section on the 1-skeleton, that is, on the union of all O-simplices and 1-simplices. In the following we denote by K1 the 1-skeleton of the simplicial complex K. (See Figure 6.1.) Now let a = |t>oUit>2| be an oriented 2-simplex and choose a trivialization (/?„ : tt-1(G) x 51. If we strictly follow the notation in Chapter 3, we should write <pa ¦ 7r~J(|cr|) = \cr\ x Slt but we simply write a in place of \a\. We also write a in place of (a). Since the section s : da —¦ 7r-1(a) is already defined on the boundary da of a, we can compose it with the trivialization y>a and obtain So :da -> 7r'(a) =axS'.
6.2 S1 BUNDLES AND EULER CLASS 243 Figure 6.2. The sum of degrees from the 1-skeleton of r into S1 is zero Since a is oriented, its boundary do can be uniquely identified with Sl. Hence by composing the map sa above and the projection to the second component a x S1 we obtain the map sa:dcr = Sl -taxS1 - Sl. Let degsa be the degree of this map (see §3.5 (d)). If it is 0, we can extend the section s to all of a. Lemma 6.17. The correspondence that associates to each oriented 2-simplex a the number degsa defines a 2-cochain of K : cs € C2(K\Z). It is, in fact, a cocycle, that is, 6cs = 0. PROOF. We have only to show that cs(dr) — 0 for any oriented 3-simplex r. But this is clear from Figure 6.2. ¦ Now let s' be a different section of ? above on a 1-skeleton Kl of K, and let cs' G Z2(K;Z) be the corresponding 2-cocycle. Then we have Lemma 6.18. cs and cs> are mutually cohomologous. PROOF. We define a 1 cochain d € Cl{K\Z) as follows. First for an arbitrary vertex v we choose an oriented path ?v that joins two points s(v) and s'(v) on the fiber Pv within Pv. For example, start at s(v) and proceed in the given direction of Pv to s'(v) - that is the path tv. Now for any oriented 1-simplex k = |uoVi|, choose a
6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Figure 6.3. tK trivialization <fK -.TT'1^) ^KX S1 and get the composite map with the projection to the second component tK : 7t-1(k) = k x S1 -* Sl. Let us write ?K for the oriented path iVQ • s'(k) ¦ ?~* • s(k)~1. Here • denotes composition of paths, and _1 after an oriented path indicates reversing the path. (See Figure 6.3.) Now lK can be naturally identified with S, and the degree of the map tK —» S1 is well defined. We define a 1-cochain d by d(n) = degi*. By simple observation we can verify that cs>(a) = cs(a) + d{da). Therefore cs> — cs = 8d, completing the proof. ¦ We can prove the following lemma by slightly modifying the proof of Lemma 6.18. The detail is left as Exercise 6.4. Lemma 6.19. Let s be a section on K1 of an S1 bundle, and let c3 € Z2(K;Z) be the corresponding 2-cocycle. Then for an arbitrary 2-cocycle c € Z2(K;Z) that is cohomologous to cs there is a section s' on Kl such that cs> = c. Now since cs is a 2-cocycle of K by Lemma 6.17, its cohomology class [cs] € H2(K\Z) = H2{M\Z) is defined. Since this class does not depend on the choice of a section s by Lemma 6.18, it looks as if the classes would be determined by the bundle f itself. But there is a
6.2 S1 BUNDLES AND EULER CLASS 245 ^^ Figure 6.4. Decomposing a problem before we can actually finish. We have fixed a C°° triangulation t : \K\ —» M in our discussion, but of course a triangulation is not unique. However, this problem can be settled by using the uniqueness of triangulation in the following sense. That is, any two C°° trian- gulations of a C°° manifold admit a C°° triangulation as a common refinement (a result due to S.S. Cairns and J.H.C Whitehead (§3.1, Theorem 3.3). Now let L be a simplicial complex that is a refinement of a Euclidean simplicial complex K. Namely, the polyhedron \L\ determined by L coincides with \K\, and furthermore for an arbitrary simplex a of L there is a simplex r of K such that a C r. In this case, t : \L\ = \K\ -* M becomes a C°° triangulation of M. Under these circumstances we have Lemma 6.20. Let s be a section defined on a l-skeleton of K and extend it arbitrarily to a section s on the l-skeleton of L. Then the two cohomology classes [cs] € H2(K\Z) and [c$] e H2(L\Z) coincide under a natural isomorphism H*(K;Z) = H"(L;Z). PROOF. Let a be an arbitrary 2 simplex. By definition of a refinement, there exist 2-simplices t{ (i = l,...,r) such that a = Ui t%- If we assign one orientation to a, it induces orientation on each T{. Now a natural isomorphism H*(K;Z) = //*(L;Z) is induced by the chain map C*{K\Z) -> C.(L;Z) that is defined by C2{K;Z) 3 a ^Y^rx € C2(L;Z). (See Figure 6.4.)
246 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES On the other hand, by definition of cs, c$ and from Figure 6.4, we easily get C5(cr) = ^cs(rt), which implies the assertion of the lemma. ¦ These discussions also show that the cohomology class [ca] € H2(M;Z) does not depend on triangulation of M. Definition 6.21. For an S1 bundle ? = (P,tt, M, S1), the cohomology class defined as above is written e(Oetf2(M;Z) and called the Euler class of f. (c) The classification of S1 bundles. The next theorem will show that the Euler classes defined in the previous section are complete invariants in the classification of oriented S1 bundles. It is irrelevant which structure group, Diff+S1 or f/(l), we choose to consider. THEOREM 6.22. The Euler class is a characteristic class for an oriented 51 bundle; that is, if & = (P^tt^M^S1) (i = 1,2) are two oriented S1 bundles with a bundle map Pi -1— P2 Mx » M2 f between the two, then the equality /"(efo)) = e(?i) holds. In particular, two isomorphic Sl bundles have the same Euler class. Furthermore for any C°° manifold M, the correspondence between the isomorphism classes of Sl bundles over M and H2(M\Z), F.2) K]-e(fle#2(A#;Z), is one-to-one and onto.
6.2 S1 BUNDLES AND EULER CLASS 247 Proof. We prove the first part that the Euler class is a characteristic class of S1 bundles. Let t{ : \Ki\ -* M{ (i = 1,2) be a triangulation of the base space Mt. Let s be a section of ?2 over the 1-skeleton of K2 and cs e Z2(K2;Z) the corresponding 2-cocycle. Then |c,| = e(?2) € H2{M2\Z). As is well known in the homology theory of simplicial complex, we may take a sufficiently fine subdivision L of K\ and find a simplicial map g : \L\ —» \K2\ that is homotopic to / : M\ —> M2. Here by saying that g is a simplicial map, we mean that it is a continuous map that takes any simplex of L to a certain simplex of K2 by an affine map. In this case, by pulling back a section s of ?2 on the 1-skeleton of K2 by a simplicial map g we can define a section g*s of ?1 on the 1-skeleton of L. In other words, if g : P\ —» P2 is a bundle map over p, we have ff(p*s(p)) = s(g(p)) for any p € Mi. In this case, obviously, we get cp.(a)(a) = c4(^(a)), that is, Cg'S = <?*(c5). Therefore e«i) = [c,-.]=^([c.]) = /(c(ft)), which shows that the Euler class is a characteristic class of an S1 bundle. To prove the second part of the theorem, we first show that the correspondence F.2) is injective, that is, if two S1 bundles over the same base space have the same Euler class, then they are isomorphic. Assume & = (P^tt^M,.?1) (i = 1,2) satisfy efo) = e(fc)- By Proposition 6.15, we can'assume that f, are principal bundles. Let t : \K\ -> M be a C°° triangulation of M, and let Si (s = 1,2) be sections of & on the 1 skeleton of K. By assumption, the two cocycles cs, and c32 are mutually cohomologous. By suitably modifying s\ or s2 we may, by virtue of Lemma 6.19, assume that cSl coincides with cS2. Now an isomorphism ?il|KM - 6l|/<M on the 1-skeleton K1 of K can be defined by the correspondence itiWK1]) 3 Sl(p)z - s2(p)z € Tr^dff1!) (P € l^l.z € 51)- (See Figure 6.5.) Since we have cSl (a) = c52(cr) for any oriented 2-simplex a of /C', the bundle isomorphism ?i|da — ?2 |<9a on the boundary of a can be extended to an isomorphism ?1 \a = &U- Here we use the well-known result that two maps fi : S1 —¦ 51 (i = 1,2) are homotopic if and only if deg f\ = deg /2. Next let r be an arbitarily oriented 3-simplex.
248 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Figure 6.5. Isomorphism over Kl We already have an isomorphism </?: i\\ar — ^Idr- We try to extend it to the interior of r. We identify ^~l{r) with r x Sl by using the trivialization of (j on r. Then it is sufficient if we can construct an isomorphism (p so that the following diagram is commutative: TrfV) ?* r x S1 —i-> 7r^(r) 3? r x S1 5r x S1 » dr x Sl. Since r is oriented, we can identify the pair (r, dr) with (?>3, S2). Now let us define a map h : dr = S2 —» S1 by setting y>(p,i) = (p>Mp)) (PG52,1G51). It is well known that the 2-dimensional homotopy group ^(S1) is 0, that is, every continuous map S2 —» S1 can be extended to a continuous map D3 —* S1. Without knowing anything about the 2- dimensional homotopy group, we can prove the above fact using only the facts that S2 is simply connected and that the universal covering space of S1 is R. Therefore there exists h : D3 —> S1 such that h|5a = h. Hence by setting <p{p,z) = (p, h(p)z) (p € D3,z € S1), we get a desired isomorphism. By similar arguments using ^(S1) = 0 (i > 2) we can show that ?i and ?2 are isomorphic on the (? 4- l)-skeleton for every I. It now follows that ?1 =&•
6.2 S1 BUNDLES AND EULER CLASS 249 Finally, we shall show that F.2) is surjective. Let x e H'2(M;Z) be an arbitrary element, and choose a 2-cocycle c € Z2{K\Z). We start with the product bundle \Kl\ x Sl on the 1-skeleton. For each oriented 2-simplex a we identify its boundary da with Sl and paste a x Sl onto IK1! x S1 by using the isomorphism da x S2 = S1 x S1 3 (iu,z) -» (w, uTc(<7)z) e da x S1 c |K!| x Sl. Performing this procedure for every 2-simplex, we get an Sl bundle over the 2-skeleton K2 of K. Next let t be an oriented 3-simplex. We already have an S1 bundle on dr. Since c is a cocycle, the bundle is trivial, as we can see from the proof that F.2) is injective. Therefore using this triviality we can paste r x S1. In this way, we can get an S1 bundle over K3. For an oriented 4-simplex p, we have constructed an S1 bundle on dp^S3. Because #2(S3; Z) = 0, the proof of injectivity will show that the bundle is trivial. Continuing these discussions, we finally obtain an S1 bundle over M whose Euler class is equal to the element x € H2(M\Z) by construction. We have worked with S1 bundles in the differentiate category. Rigorously speaking, we need to use an appropriate open neighborhood in place of each skeleton Kl, and so on. We skip the technical details. (d) Defining the Euler class for an Sl bundle by using differential forms. The content of our discussions in this subsection is a special case and a preview of the general theory taken up in §6.3 and thereafter. Let #o € .^(S1) be a left invariant 1-form on the Lie group S1 = 17A) such that /s, #o = 27r; namely, #0 = d6 in terms of the polar coordinates. Let ? = (P,n,M,Sl) be a principal Sl bundle over a C°° manifold M. For each p € M, we have an identification ip : S1 ^ 7r-1(p) c P, namely, identifying S1 with the fiber over p. Thus ip is unique up to rotation. For each element z in Sl, let Rz : P -* P denote action on the right: R2{u)=uz (z€Sl,ueP). Definition 6.23. A 1-form cj on the total space P is called a connection form if it satisfies the following two conditions: (i) For each p€ M, i^u = #o; (ii) For each z € Sl, R*zu) = w.
250 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES M— Figure 6.6. Horizontal tangent vectors Given a connection form a\ we may define at each point u e P Keru;u = {X e TuP;uju{X) = 0}. We might say that it is the set of all " horizontal" tangent vectors. See Figure 6.6. PROPOSITION 6.24. Any Sl principal bundle admits a connection form. Proof. Let {Ua}a be an open covering of the base space M such that for each or there is a local trivialization <pa : ix~l{Ua) = UaxSl. If q : UQ x S1 —» S1 denotes the projection onto the second component, then (q o tpa)*9q is a. connection form on the trivial bundle 7r_1([/a). Using a partition of unity {/«}<>> w© get a connection form uj = ]P(/a ° t)(9 o <pay$0. PROPOSITION 6.25. Given a connection form in a principal Sl bundle, there is a unique 2-form Q € A2(M) on the base such that PROOF. First, some local considerations. If we identify n'1^^ with the product bundle Ua x S1 by using a local trivialization t«7a, condition (i) in Definition 6.23 says that u> is of the form
6.2 S1 BUNDLES AND EULER CLASS 251 where x = (x\,... ,xn) and 9 are coordinate functions of Ua and Sl. Next, u} is invariant by the action of Sl (condition (ii) in Definition 6.23) and therefore fi is independent of 6: fi(x,6) = ft{x). Hence du> can be written on 7r-1(?/a) in the form F.3) Y^^T.dxjKdx,. This is, however, a 2-form on the base Ua. The projection -n : P —> M is obviously a submersion. We hence conclude that n* : ,4*(M) —* A*{P) is injective (see Exercise 2.6). It follows from this that the 2-forms F.3) on Ua for each a make up a 2-form Ct and that the equation du> — 7r*Q holds on M. ¦ Proposition 6.26. The 2-form Cl formed as indicated above is a closed form. Its de Rham cohomology class [Q] € HpR{M) is independent of the choice of a connection form u). PROOF. By taking the exterior derivatives of both sides of the equation duj = -n*Q. we obtain n'dCl = 0. Since 7r* is an injection, we conclude that dQ = 0, which proves the first half of the proposition. To show the second half, suppose u/ is another connection form. We find that there is a 1-form r e A1 (M) such that <J = U) + 7T*T, as follows. As in the proof of Proposition 6.25, the connection forms u) and u/ can locally be written in the form u> = ^2fi(x)dxi + e0, J = ]T#(x)dxi + 0o, and hence J -u) = J2^(x) - ft{x))dxi, which is clearly a 1-form on Ua. By a similar argument these 1-forms determine a 1-form r on M such that u/ - u> = tt't. In this case, if dw — tt*Q, dcv' = n*Q', then 7T*(n - fi) = a\J - duj = ix*dr. Since 7r* is injective, we get Q' - Q, = dr,
252 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES and finally, [O'] = (ft] € HpR, completing the proof. ¦ Definition 6.27. The 2-form ft on the base M determined from a connection form of a principal S1 bundle is called the curvature form. Also the de Rham cohomology class is called the real Euler class of ? and is denoted cr(?). The next theorem shows that the two definitions of Euler class, namely, the one using sections and the other using differential forms, essentially coincide. Theorem 6.28. For a principal S1 bundle f, we have eR@ = e@®K€tf2(M;R). PROOF. By using a triangulation t : \K\ —> M we identify M and \K\. First, as we did at the beginning of the previous section, we construct a section of (an open neighborhood of) the 1-skeleton \Kl\ of K. Using this, we define a trivialization ?||KM = \KX\ x S1, which in turn determines a connection form on \Kl\. Extend it to a connection form u> on all of P. This can be done just as we did using a partition of unity in the proof of Proposition 6.24. Now let a be an oriented 1-simplex. Again if we choose a trivialization as at the beginning of the previous section, then we can write the section a on da in the form s<r : da B p >-> (p, sa(p)) e da x Sl. Here sa:da = S1 ^Sl. If we set cs(a) — degsa, then this is a 2-cocycle, and e(?) = [cs] € H2(M;Z), as we know. Let ip:da xSl -^da x S1 be the trivialization by s, that is, an isomorphism defined by ip(p, z) = (p, Scrip), z) (p e da, z € SY). Then by construction of u) we have F.4) ^(/^-¦(aa)) = (^r0o,
6.2 S1 BUNDLES AND EULER CLASS 253 where N(ir~l(dcr)) is a suitable open neighborhood of ir~y(da), and #o = d6 is an invariant 1-form on S1. From the definition, du = ir*Cl and d$o = 0. Hence we get ^|NGr-i(da)) = 0j that is, Q vanishes identically in a neighborhood of da. In this case we have 1 / du A 0o 47T2 JdaxS* Here in the last equality we used du) A#o = d(u A#o) and the theorem of Stokes (Theorem 3.6). On the other hand, F.4) implies W Jdaxs* 47r27aaxS1 which is equal to -degs = degsa =cs{o), where we use i/,_1(P>z) = (P> S^HpJ) (p € ^ ^ € 51). We now have eR(?) = e(f), ending the proof. ¦ Example 6.29. The Hopf map h : 53 -» S2 (§1.3, Example 1.27) has the structure of a principal 51 bundle ?, as can easily be verified. It is called the Hopf Sl bundle (see Exercise 6.1). We shall now find that e@ = -l€tf2(S2;Z) = Z. Here the orientation of S2 is one that is naturally determined by the identification S2 = CP1. See Exercise 1.3. Now let Vi = {[2i,*2];*i * 0}, U2 = {[zuz2];z2 * 0}. Then S2 = CP1 = Ux n ?/2- The trivialization y? : h-l{Ux) = Ux x S1 of the Hopf bundle over U\ can be given by F.5) h-\UlK{zuz2)~{[zuz2),^).
254 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Now we define a section s : U-i —¦ h~l(U2) by Now by setting ?> = {[1,*] ? Ui\\z\2 < 1}, we have dD S S1 = {z; |z| = 1}. For any z € Sl = 5.D we have 5(Z) = S([1)Z]) = SA,-1I]) = (^-^). Hence if </)os = (id,s), then we get $(z) = z~l. Since degs = — 1, we have completed the proof. Suppose 7r : E —> M is an oriented 2 dimensional-vector bundle. In §5.6 we defined the Euler class e(E) e H2{M\R). On the other hand, by fixing a Riemannian metric in E and setting P = {u € E\\\u\\ = 1} we see that the natural projection -n : P —> M is an oriented S1 bundle, as can easily be seen. As such, the Euler class can be defined and coincides with the first one, as expected. The verification of this fact is left as Exercise 6.5. (e) The primary obstruction class and the Euler class of the sphere bundle. In this subsection we generalize the definition of the Euler class of an S1 bundle and introduce the characteristic class called the primary obstruction class. Let F be an (? - l)-connected C°° manifold, so that ir^F = 0 for i = !,...,? - 1. We assume I > 1 for simplicity. By a well- known theorem of Hurewicz, we have i\t F = He(F\ Z). A fiber bundle ? = (E, 7r, B, F) with fiber F is called an oriented F bundle if for each 6 G B it is provided with an isomorphism H((Eb\Z) = Hi(F\Z) and they locally are concordant in the sense each point has an open neighborhood U and a trivialization v? : 7r-1([/) = U x F such that for an arbitrary point c € U the composite map He{Ec;Z) -» He(*-l(U)\Z) -» He{U x F;Z) - Ht{F;Z) coincides with the designated isomorphism. For instance, an F bundle is orientable if B is simply connected or if the structure group is connected. Now assume that we have an F bundle f = {E,rc.B,F). Start with a triangulation t : \K\ —> B and try to construct a section of ?. In the same way as in the case of an Sl bundle, we can construct a
62 S1 BUNDLES AND EULER CLASS 255 section s on the ^-skeleton Kl. Now for any (? + l)-oriented simplex <x, choose a trivialization <p : n~l{a) = a x F. We already have a section denned on the boundary da of a. Composing it with <py we get a map sa : da = Se -> a x F -* F, where the last map is a projection. Let us denote the homotopy class of this map by \sa] e ireF. Then we see that the correspondence Ci+i{K\Z) 3 a >-> [sa] e neF determines a cocycle c e Ze+1(K\7T(F), and it is known that the following theorem holds. THEOREM 6.30. Let F be an (? - I)-connected C°° manifold. For an oriented F bundle ? = (E,7r,B,F) the cohomology class [c] € Hi+l(M]ireF) of the cocycle defined as above is uniquely determined independently of the choice involved in the process. It becomes a characteristic class of the oriented F bundles. The characteristic class defined above is called the primary obstruction. In particular, the primary obstruction e(E)eHn(B\Z) of the Sn_1 bundle ix : E —> B is called the Euler class. This is a topological definition. For the characteristic classes of vector bundles such as the Pontrjagin classes and the Chern classes, we can find topological definitions as the primary obstructions of appropriate associated fiber bundles. They are meaningful with coefficients in Z. (f) Vector fields on a manifold and Hopf index theorem. We proved the Gauss-Bonnet theorem in §5.7 (a) as follows, we first constructed a vector field with a nice property by using triangu- lation of the manifold. Then we chose a good connection so that the Euler form is concentrated around the singularity of the vector field. Finally, we determined the value of the integral. In this subsection we discuss singularities of a vector field in general. Let X be a vector field on an n-dimensional C°° manifold. A singular point q € M is said to be an isolated singular point if there is a sufficiently small neighborhood of q which has no zero point of X other than q itself. In this case, we can take a closed neighborhood N of q and a diffeomorphism N = Dn so that for each
256 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES peiVwe can identify TPM with Rn. Let tt : Rn - 0 -* 5n_1 be the natural projection and consider the map ixoX-.dN^ S71'1 -> Rn - {0} - Sn~l, and, using its degree, set md{X,q) = degiroX € Z. We can easily verify that this number is independent of the choice of diffeomorphism N = Dn by using the property of the degree of mapping. This number is called the index of X at q. Now we have a result called the Hopf index theorem. It is also called the Poincare- Hopf theorem, since Poincare" already had it for the case of surfaces. Theorem 6.31 (Hopf index theorem). Let M be a closed C°° manifold and let X be a vector field on M with the singularities all isolated. Then we have ^ind(X^) = X(M), q?S where S is the set of singularities of X. Sketch of the proof. We may assume that M is connected. If M is not orient able, we can consider the double covering M -+ M as in the proof of Theorem 4.21 in §4.4. Then x(M) = 2*(M). On the other hand, the vector field X on M naturally induces a vector field X. If S = n-l{S), then obviously J2 ind(X, q) = 2 ^ ind(X, q). Therefore it is sufficient to prove the theorem for the case where M is oriented. If we introduce a Riemannian metric and set TlM = {XeTM;\\X\\ = l}, then tt : TiM -> M becomes an oriented 5n_1 bundle, which we call the unit sphere bundle. Let e{TxM) e Hn(M;Z) be its Euler class. We want to prove that (e{TlM),[M})=X(M).
6.3. CONNECTIONS 257 Now let A" be a vector field with isolated singular points. From the definition of the Euler class as the primary obstruction class, we have ((eT1M),lM]) = Y,'md(X>ti- q€S Hence in order to prove the theorem, it is sufficient to construct just one vector field X such that ?ind(X)G) = x(M). q€S For example, we may take the vector field constructed by means of triangulation in the proof of the Gauss-Bonnet theorem recalled at the beginning of this subsection. Indeed, in this case, X has a singular point at the barycenter of each i-dimensional simplex of K. But its index at the barycenter is equal to the index of the vector field V*, considered on Rn of E.27) in §5.7 at the origin. Then we obviously have ind(yi,0) = (-l)i, and hence X>d(X,g)=?(-l)dim*=X(M). q?S creK ¦ 6.3. Connections (a) Connections in general fiber bundles. Let ? = {E}ir,ByF) be a fiber bundle with a C°° manifold F as fiber. By definition, ? is locally like a direct product but globally may not be; that is, globally it may be twisted. One measure for such twisting is given by the notion of connection. Let us consider the case of a direct product E — B x F. In this case, at any point u = (b,p) € E, the tangent space TUE is written TuE = TbB®TpF, namely, the direct sum of two subspaces, one in the base direction and the other in the fiber direction. According to this direct sum decomposition, any tangent vector X € TUE is decomposed as X = Xh + Xv (Xh e TbBy Xv e TPF),
258 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES X X* B ¦ Figure 6.7. Decomposition of tangent vectors where Xh and Xv are called the horizontal and vertical components of X, respectively (Figure 6.7). For a fiber bundle in general, it is natural to call vectors tangent to a fiber "vertical vectors". To be more precise, the subspace of TUE given by Vu = TuEbcTuE F = 7r(u),Eb = 7r-1F)) is the vertical subspace at u. However, there is no natural way to distinguish vectors as "horizontal". So we go ahead and define horizontal vectors, which later will lead to the notion of curvature as adjustment and to the global invariants, i.e. characteristic classes. We state Definition 6.32. Let ? = (E,tt,B,F) be a fiber bundle with C°° manifold F as fiber. If at each point u e E we can choose a subspace Hu of TUE in such a way that u ¦-» Hu is C°° and Hu is transversal to the fiber (thus, TUE = HU(BVu is the direct sum), then we say that ? is given a connection. The notion of a general connection was introduced by Ehresmann [E] as a modernization of Elie Cartan's idea of connection. Let us recall that for any C°° manifold M, if we are given an assignment of a subspace Dp C TpM in such a way that \J Dp is a differentiate subbundle of TM, then this assignment (or the subbun- dle \JpDp) is called a distribution (§2.3 (a)) on M. In this language, a connection in a fiber bundle is nothing but a distribution on E that is transversal to every fiber.
63. CONNECTIONS 259 Proposition 6.33. An arbitrary fiber bundle admits a connection. Proof. Give the total space E a Riemannian metric and define for each u G E, the subspace Hu to be the orthogonal complement of vu. m If a connection is given, an arbitrary tangent vector X G TUE is uniquely decomposed as x = xh + xv (xheHu,xveVu). We call Xh and Xv the horizontal and vertical components of X, respectively. Vectors in Hu are said to be horizontal. Using horizontal vectors, we may relate points lying on different fibers in a certain way, as we now explain. Let c : [a,b] —» B be a smooth curve in the base B. A curve c : [0,6] -¦ E is called a lift if tt(c@) = c{t) for each t G [a, &]. If the velocity vector c(t) is horizontal for every t, we call c(t) a horizontal lift. Proposition 6.34. Let ? = {E,-n,B,F) be a fiber bundle with a compact C°° manifold F as fiber, and fix a connection in ?. Let c : [a, b] —* B be a piecewise C°° curve (that is, a continuous curve, C°° except at a finite number of points) such that &o = c(o) and b\ = cF). Then for any point uq G EbQ there is a unique horizontal lift c : [a, b] —¦ E such that c(a) = uq. PROOF. By dividing the curve c into pieces by taking a finite number of points, if necessary, we may assume that c is contained in some open subset U of B on which ? is trivial, and furthermore c has no self-intersection. In the general case, we may find a finite number of horizontal lifts that can be connected successively. Now if the image of c is denoted by C, then n~l(C) is isomorphic to [a, 6] x F, on which a connection is given. We define a vector field X on 7r-1(C) in such a way that Xu is horizontal for every u G n~l(C) and 7r*Xu = c{t), where 7r(u) = c(t). This is clearly a nonsingular vector field. Let us consider the maximal integral curve -y(t) of X through u0. By definition of X we see that 7 starts at 7@) = u0 G Ebo, proceeds transversally to each fiber, and produces a horizontal curve. Moreover, if F is compact, then we can show that 7 reaches the fiber E^ over 61. We now prove this final assertion as follows.
260 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES We assume that 7 does not reach E^ and derive a contradiction. Thus we assume that there exists to <b such that 7 is defined for all t < to but not at to. Choose a monotone increasing sequence {?„} such that limn-^oo tn = to- By composing the trivialization 7r_1(C) = (a, b] x F and the projection [a, 6] x F —> F we obtain a map q : 7r-1(C) —> F. Let pn = <jG(tn)). By assumption, F is compact. By taking a subsequence if necessary we may suppose there is a point p € F such that p = limn_o0pn. Let u € Fc(t0) De fc^e Point sucn that g(ii) = p. Then 7 passes through an arbitrary neighborhood of u but does not reach ?c(t0)- ^n tne other hand, by definition of X an integral curve of X that goes through a sufficiently small neighborhood must pass through Fc(t0)- This *s a contradiction. ¦ From Proposition 6.34 we can obtain a map hc : E6o —> Fbj by moving uo through 2?&0. This is called the parallel displacement along a piece wise smooth curve cm B. The map hc is constructed by using integral curves of a vector field X, which in turn are obtained by solutions of an ordinary differential equation. By using differentiability of solutions on the initial values we find that hc is a C°° map. The map does not depend on the choice of parametrization t for the curve c. For the curve c-1 obtained from c by reversing the direction, we have hc-i = hc~l. It follows that hc is a difFeomorphism. If two curves c and c' are such that the endpoint of c coincides with the initial point of c', then the parallel displacement along the composite curve co c' is the composite of hc> and hc. Finally, we remark that if the fiber F is not compact, parallel displacement many not be defined (up to the last point of the curve c). See Figure 6.8. One gives a conceptual illustration of parallel displacement, and the other shows an example in which parallel displacement is not defined beyond the interval Fo»^)- (b) Connections in principal bundles. A principal bundle has its structure group G acting on the entire space to the right. It is then natural to relate a connection to the action of G. It becomes possible to define parallel displacement without assuming that G is compact. We start with Definition 6.35. Let ? = (P,-n,M,G) be a principal bundle with structure group G. A connection on f is a rule to assign a subspace Hu of TUP at each point u G P in such a way that the following conditions are satisfied: (i) Hu is transversal to the fiber, that is, TUP = Hu © Vu;
6.3 CONNECTIONS FIGURE 6.8. Parallel displacement (ii) {Hu} is invariant by the right action of G, that is, if Rg : P -> P{g € G) is defined by #s(u) = ug {u e P), then #ttS = (Rg)*Hu; (iii) //u is differentiable in u. The only difference from the definition in the case of general fiber bundles is adding the second condition on invariance by the structure group G acting on the right. But it is a strong condition that makes it possible to develop the theory of characteristic classes for a principal bundle. At present, we do not know of any appropriate substitute condition for a general fiber bundle. The following proposition illustrates the power of this invariance condition. Proposition 6.36. Let ? = (P, vrt Af, G) be a principal G bundle. Given a piecewise smooth curve c : [a, 6] —> M and an arbitrary point uq € 7r-1(c(a)), there is a unique horizontal lift c : [a, b\ —» P of c such that c(a) = uo. It is thus possible to define parallel displacement hc ¦ Pc(a) -» Pc(b) o-long c. Proof. If G is compact, we already have Proposition 6.34. If G is not compact, we replace the second half of the proof of Proposition 6.34 as follows. As before, n~l(C) is isomorphic to the direct product bundle [a, 6] x G on which we considered integral curves of a nonsin- gular vector field X. We now want to derive a contradiction from the
262 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES assumption that a maximal integral curve ^(t) ceases to be defined at t = tot where to < b. In the current situation G acts on 7r_1(C) to the right, and the horizontal direction for the connection is invariant. It follows that, for each g E G, the curve i(t)g is a horizontal lift of c. Hence each integral curve j{t)g cannot reach 7r_1(c(?o))- This is, however, a contradiction to the fact that an integral curve of X through an arbitrary point of the fiber Pc(t0) over c(*o) is transversal to the fiber. ¦ Thus, given a connection in a principal bundle, arbitrary smooth curves in the base space M can be lifted to horizontal curves in the space P, by means of which parallel displacement of each point on the fiber is defined. Now what can we say about a smooth surface with boundary, or more generally, a submanifold N in M? A submanifold N in P is called a lift of N if the restriction of the projection tc to N is a diffeomorphism N = N. Furthermore we say that N is horizontal if the tangent space TUN at each point u € N is contained in Hu. With these definitions, the problem is this. Let p be a point of N and choose a point uq in the fiber n~x{p) over p. Can we then find a submanifold N in P going through p that is a horizontal lift of N7 If we try to copy the proof of Proposition 6.34, we would need an integrable distribution of a higher dimension. For higher dimensions, a distribution is integrable if and only if it is involutive (Frobenius theorem, §2.3). This condition is further equivalent to the condition of zero curvature. So much for the digression. (c) Differential form representation of a connection in a principal bundle. We shall now discuss how we represent a connection in a principal bundle ? = (P, 7r, M, G) by using a differential form. Let g be the Lie algebra of the structure group G. Consider the Maurer-Cartan form u>o, namely, a g-valued 1-form on G such that uq{A) = A for every A e g (§2.4 (b)). For each g e G, we define Lg '. G —> Gy Rg '. G —* G by L9{h) = gh, R9(h) = hg for all h 6 G. Of course, we have L*u>o = ^o- To find H*u>o, we observe that, for A e g (R9).A=:(RgULg-r).A = (tg->).A,
6.3. CONNECTIONS 263 where tg-\ is an automorphism of G given by G 3 h *-* g~lhg e G. The differential (tff-0* of Lg~l is usually denoted by Ad(p-1) g GL{q). Here Ad : G —» GL(g) is a homomorphism called the adjoint representation. Hence (jRff)»A = Ad(g~l)A. Therefore we get F.6) R*guo = Ad{g-l)uo. Since ? is locally trivial, we may take for an arbitrary fiber 7r_1 (p) (p G M) a trivialization of ? over an open neighborhood of p, and find a diffeomorphism ip:G^7r-\p) such that F.8) ip(hg) = ip(k)g (heG). This diffeomorphism is unique up to an element of G acting on the left. That is to say that the diffeomorphism determined by another trivialization is indeed given as the composite map ip o Lg : G —> 7r_1(p) for some g G G. Therefore by letting the differential of ip act on left invariant vector fields, i.e., on elements of g, we can conclude the following. For any point u G 7r_1(p) an isomorphism Tu{n~x{p)) = g is uniquely determined. But TuGr-1(p)) is nothing but the set of all vertical vectors Vu = {X G TuP;tt*X = 0}. In this way we are now led to the natural identification Vu ^ g for any u G P. This also means that an arbitrary element A G Q induces a vector field A* on P. Such vector fields are called fundamental vector fields. In other words, they are vector fields that correspond to one- parameter transformation groups RexptA- Now if ? is given a connection, then at each point u G P the horizontal subspace Hu is assigned together with a direct sum decomposition TUP = Vu 0 Hu, which induces the projection F.9) TUP ->Vu=g. We can define anwe Al{P;g), that is, a 9-valued 1-form on P, by F.9). If we know v G A1{P; g), then we can get ^ = {X6TuP;w(X) = 0}.
264 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Since Hu depends differentiably on u, it follows that u> is also differ- entiable. The form w thus defined is called a connection form. By- definition and F.7) we clearly have F.10) ipU - u0 or, equivalently, F.11) <v(A*) = A (Aeg). The action of G on P to the right, Rg : P —» P, Rg{u) = ug, and the condition Hug = (Rg),Hu for the connection imply that R*u> is 0 on an arbitrary horizontal vector. For vertical vectors, F.6) and F.8) determine the relation between R*gu and u>. In summing up, we have F.12) R*gu = Ad{g-l)u. Conversely, F.11) and F.12) together characterize the connection form; that is, Theorem 6.37. // a connection is given in a principal bundle (P,7r,M, G), then a g-valued l-form u, called the connection form, satisfying the following two conditions is determined on P: (i) uj(A*) = A for an arbitrary A € g; (ii) RJu; = Ad{g~x)u>. Conversely, for a Q-valued l-formu) on P satisfying the two conditions above, there is a unique connection whose connection form coincides with ijj. Proof. We already proved the first half. To prove the second half, we set Hu = {X € TuP;u{X) = 0} (u € P). Now we can show that the distribution u« Hu satisfies the conditions for a connection as follows. From condition (i), we see that u{X) = X for any vertical tangent vector X € Vu = g. Thus Hu is transversal to the fiber, and a direct sum decomposition TUP = Vu©//U holds. From condition (ii) we see that Hug = (i?9)»Hu. Since u is differentiate, we see that Hu depends differentiably on u. ¦ PROPOSITION 6.38. A principal G bundle admits a connection.
§6.4 CURVATURE 265 Proof. First, a product bundle M xG admits a trivial connection, as we saw in (a) of §6.3. The corresponding connection form is given by <?*u>o, where uq is the Maurer-Cartan form of G and q : M x G —>• G is the projection to the second component. Now for a principal G bundle, we take an open covering {Ua} such that 7r_1(t/0) is a trivial bundle. Let u>a be an arbitrary connection form and let {/a} be a partition of unity subordinate to {UQ}. Then w = 53(/a ° f )wQ a is a connection form. ¦ §6.4 Curvature (a) Curvature form. Suppose a principal bundle n : P —* M with structure group G is given, and let u> G .^(Pifl) be a connection form. If P is the product bundle M x G and if w coincides with the trivial connection cj = <?*cjo, then the Maurer-Cartan equation §2.4 B.46) says F.13) duj = --|w,w]. In the case where P is general, F.13) is valid when restricted to each fiber but uncertain when evaluated in horizontal vectors. Thus we simply define a g-valued 2 form Q by F.14) du) = --[u,u] + Q. The form Cl is called the curvature form and F.14) the structure equation. Let us choose a basis B\,..., Bm in g. The connection form u; is expressed by u = y^utjBj, iary 1-forms. Lil m where u>i,... ,u/m are ordinary 1-forms. Likewise, Q can be expressed by
266 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES where f2i,..., fim are ordinary 2-forms. Let c? be the structure constants (see§2.4 B.43)) of g with respect to the above basis. Then the structure equation F.14) can be written F.15) d^ =--^w,-Awit+fij 2 M by §2.4 B.45). We study the properties of the curvature form Q. PROPOSITION 6.39. Let u> be a connection form on a principal G bundle n : P —» M, and Q the curvature form. Then we have the following. (i) For an arbitrary geG, R*g?l = Ad(p-1)n. (ii) For arbitrary vectors X,Y e TUP, Sl{X,Y) = dw{Xh,Yh). Hence for any vertical vector field Z, i(Z)Cl — 0. (iii) // X, Y are horizontal vector fields on P, then ?1{X, Y) = -iw([x,y]). (iv) (Bianchi's identity) dCl = [ft,u>]. PROOF. In order to prove (i), we let R* operate on the structure equation F.14) and note that R*uj = Ad(y_1)cj. Since Ad(g_1) is a Lie algebra automorphism of g, we get [Ad(G-1)u>, Ad(<7_1)u>] = Ad(g~l)[uj,u)}. Now (i) follows from these equations. To prove (ii), let us write X — Xh + Xv and Y = Yh + Yv using horizontal and vertical components. Then Q(X, Y) = Cl{Xh, Yh) + tl(Xhy Yv) + Q(XVi Yh) + Q(XV, Yv). From the structure equation F.14) we get Q(Xh,Yh) = MXh,Yh) + ±[u>(Xh)MYh)] = MXh,Yh). Thus it is sufficient to show that ft(X, Y) = 0 if X or Y is vertical. Since ?1(Y, X) = — 0(y, X), we may assume that X is vertical and prove Q.(X, Y) = 0. We do this in two cases (a) and (b). Case (a): X and Y are vertical. There exist A,Bea such that X = ^4* and Y = ?*. In general, for two fundamental vector fields A* and B* we have [j4*,?*| = [j4,?]*. (We shall prove this in a
§6.4 CURVATURE 267 moment.) Using F.14), we get Q{A\ B") = dw{A\ B*) + huj{A*),u>{B*)} = i{^*w(B*) - B*uj(A')-l>{\A\B*]) + \A,B}} = \{A\B) - B*{A) - [A,B] + \AtB}} = 0. At this point, we prove the assertion [y4*,B*] = [.A, B]* we used in case (a) above. By virtue of local triviality, it is sufficient to show this formula for the case P = M x (?, but then a fundamental vector field is nothing but a left invariant vector field on G as the second component of P. Thus the formula is valid. We now turn to Case (b): Assume X vertical and Y horizontal. As in (ii), we can write X = A^ (A € fl). Then choose a horizontal vector field Y on P such that Y = YU. In this case, we get ?1{A\Y) = dxj{A*,Y) + hu>(A')MY)) = l{AWY)-MA*)-u>([A\Y))} = 1-{Y(A)-uj([A\Y})} = -\u}{\A\Y)Y It suffices therefore to show that |j4*,yr] is a horizontal vector field. As we remarked in §6.3 (c), the 1-parameter transformation group generated by the vector field A* is R9t, where gt = exptA Therefore the formula B.29) for Lie differentiation in §2.2 (f) says 1 J t-*o t By definition of a connection, if Y is horizontal, then so is (R9_t)*Y and hence [A*, Y) as well. This completes the proof of (ii).
268 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Now to prove (iii), let X,Y be horizontal vector fields. Then by (ii) we get Q{X,Y)=duj{X,Y) = ^{Xuj(Y)-YuJ(X)-oJ([X,Y))} = -\u([X,Y)). Finally, we prove (iv). After exterior differentiation of both sides of the structure equation F.14), and using B.41) and B,42) of §2.4, we obtain dd, =-d[u>,u;] = -([dw,cj] - {ui,du>}) = {du>,(jj\ — --[[w,u;],cj] + [Cl,u>] =[Q,u;]. (b) Weil algebra. Let 7r : P —> M be a principal G bundle. If a connection is given in P, its connection form cj and curvature form 0, are determined. They are g valued 1- and 2-forms on P. They induce a certain system of ordinary differential forms on P in the following way. First suppose we are given an element of the dual space g* of g, namely, a linear map a : g —» R. For each point u € P the composite map OLOu :TuP-+g -> R is obviously an element of T*P that is of class C°°. Hence it is a 1-form on P, which we shall denote by w(a). By allowing all a € g*, we get a linear map As in the previous section, we choose a basis Bi,...,Bm of g and write a> = ?V wtBi, where o»t are 1-forms on P. If 6\,... 0m form the dual basis of g*, then u>@,) = u^. By extending F.16) to exterior products we obtain a linear map F.17) w: AV->^*(P). Specifically, we have w@j, A--A^)=w,1A-"Awij,.
§6.4 CURVATURE 269 The meaning of the symbol uj in F.16) and F.17) should be clear For instance, A*g* denotes the set of all left invariant differential forms on G, and a given connection in P induces a linear map from A*g* to the set of all differential forms A*(P). However, in general this linear map F.17) does not commute with exterior differentiation - the deviation from being commutative is indeed expressed by the curvature. Now in the discussions above let us see what happens if we replace the role played by the connection form w by that of the curvature form Q. Then a linear map F.18) n : 0* -» A2(P) is defined by setting Q(a) =Qofi: TUP x TUP — g -» R. If we write SI = QiBi + ¦¦¦ + nmSm, then fit = Q(^). From the Maurer-Cartan equation B.45) and F.15), we get Q{ = du)i - u){ddi) = (d o uj - u> o d)[6i). Thus for any element a6g" we get F.19) CL(a) = {dw-ujd){a). This means that Q shows the difference between dou> and uj od. On the other hand, the images by F.18) are 2-forms on P and hence commutative. Here we want to consider the polynomial algebra fc=i generated by g*. Specifically, this algebra is described as follows. Although we have S1g* = g, we define its degree as an element of S1q* to be 2. To distinguish these elements, we write a for the element in 5Jg* that corresponds to a € g*. Accordingly, #i,... ,6>m form a basis of S1^. Now a map / : g —> R is called a polynomial function if it can be expressed as a polynomial of ^i,..., 6m. This definition does not depend on the choice of a basis, as we can easily see. We have S*q* = {/ : polynomial function g -+ R}.
270 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES If we choose a basis as above, then S*g" can be identified with the ring of polynomials R[9\,..., 9m\. In this case, if we set tofa .. jik) = Sl^ A---Aftu, F.18) can be extended to a linear map F.20) H:SV -»-4*(P). We now define which is called the Weil algebra. We combine the two linear maps F.17) and F.20). Then w = u <g> Vt is a linear map F.21) w:W{q)->A*(P) that preserves degree and product. The image of w is the subalge- bra generated by the connection forms w, and curvature forms f^ in A*(P). (c) Exterior differentiation of the Weil algebra. The Weil algebra W(q) defined in the preceding subsection can be regarded as a universal model for a differential system generated by the connection and curvature forms of a principal G bundle. If a connection is given on a concrete principal G bundle, it induces by F.21) a linear map from W(q) to A*{P). Now it is natural to define a linear map 6 : W(g) —» W(q) that raises degree by 1 and also makes the following diagram commutative: F.22) W(9) -^ A*{P) 4 1- Wfo) > A*(P). w To do this, we shall define the following three actions that exist in^'(P): (i) Interior product by a fundamental vector field i( A*) : A*(P) -* A*(P) (Ae&); (ii) right action R*g : A*(P) -* A*{P) {g e G)\ (iii) Lie differentiation LA> : A*{P) —* A*(P) by a fundamental vector field A* (A e g),
§6.4 CURVATURE 271 on the algebra W(g) in such a way that they are preserved by the maptu : W(g) -> A*{P). (i) Interior product i(A) : W(g) —» W(g). For any element a € g* C W(g), a property of connection form requires i(A)a = ct(A). Also for the corresponding a € Slg C W(g) properties of the curvature form (Proposition 6.39 (ii)) require i(A)a = 0. Now i(A) is to be an extension to the whole W(g) as antiderivation of degree -1. Lemma 6.40. For any A € g the following diagram commutes: W(g) -?-» A*(P) W(g) —- A*{P). PROOF. Since w is a linear map that preserves product and since i(A) and i(A') are antiderivations of degree -1, it is sufficient to prove that w(i(A)a) = i(A*)w(a), w(i(A)a) = i(A*)w(a) for any a G g. The first identity follows from w(i(A)a) = w(a(A)) = a{A) and i(A*)w(a) = i{A*) = a(A). The second identity follows from i{A)a = 0 and i{A*)w(a)Q = 0. ¦ (ii) Defining the action g* : VV(g) —> W(g) by an element g G G. Let Ad(o-1)* : g* —» g* be the dual map of Ad{g~l) : g —> g. Also, we denote the corresponding automorphism of S*g* by the same notation Ad((?-1)*. Now set g*a = Ad(g~1)*ay g*a = Ad(p-1)*d. Lemma 6.41. For an arbitrary element g ? G, the following diagram is commutative: W(fl) —^ ^(P) W(g) > .4*(P).
272 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Proof. Since w is a linear map that preserves product and since both g* and Rg also preserve product, it is sufficient to show that for any a € g we have w{g*a) = R*gw{a), w{g*a) = R*gw{a). The first identity immediately follows from the property R*goj = Ad{g~l)u of the connection. The second identity follows from R-gn = Ad(g-l)?l, which is obtained as follows. Let Rg operate on the structure equation F.14) and use the property above for the connection form u as well as the fact that Ad(p_1) is an automorphism of the Lie algebra g. ¦ (iii) Defining Lie differentiation LA : VV(g) —> W(g). From the definition of Lie differentiation §2.2 (d) we set, for any a G g*, LAa{B) = -«{[A,B}) (Beg). Also, for the corresponding d € 5!fl* we set LAa = LAa. We then extend these to all of W(g) as derivation of degree 0. LEMMA 6.42. For any element a € g, the following diagram is commutative: W(fl) -^ A*{P) LA[ [LA. W{q) -—- A*{P). PROOF. Since Lie differentiation by an element of g is an infinitesimal version of the action of G, that is, LA.n = hm t '- (n e A*{P))t commutativity of the diagram follows from Lemma 6.41. ¦
§6.4 CURVATURE 273 With these preparations we proceed to define exterior differentiation 6 : W{q) -* W{g) so that F.22) becomes commutative. First of all, for a 6 9*, F.19) implies dw{a) = w(da) + H(a) = w{da + a). By defining F.23) 6a = da + a we get d o w(a) — w o <5(a). As for $i € g*, we have Next we define 6a for each a € 51g*. Letting 6 operate on both sides of F.23), we realize that we should define F.24) 6a = -6{da). Therefore we should have, for 6i, F.25) 60i = -5{dBi) = -6{-\Y,<?i*9i A *0 1 i.k j.k j,k Here we used d(d$i) = 0. Also we skipped the sign <g> in denoting elements of W(q). Since elements of A*g* and S*g* commute with each other, we changed the order of their products in an arbitrary fashion. For instance, the last term of F.25) can be handled as follows. From LB,0i{Bk) = -0i{[Bj,Bk]) = -c)k we obtain F.26) LBi6i = -Y,J3k0k- k Comparing F.25) and F.26), we get 6§i = ^20j<S>LB^9i. j Because of linearity of this equation in 6i we find that we should define F.27) <5d = ]T 0* <g> L^~a {a € SV).
274 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES What this means is that we extend the definition of <5 on g* and Slg* given by F.23) and F.27) onto the whole of W{q) as antiderivation of degree +1; that is, we require two homogeneous elements x,y € W(q) to satisfy F.28) 6(xy) = {6x)y + (-l)degxx6y, where degx denotes the degree of x. It is easily shown that 5 is now defined without contradiction. PROPOSITION 6.43. Exterior differentiation 6 of the Weil algebra W{g) -» W(g) satisfies 6oS = 0. PROOF. Let S operate on both sides of F.28); we get 62(xy) = F2x)y + (-l)d'**xF2y). Thus it is sufficient to show S2a = 62a = 0. The first result follows from F.23) and F.24). To show the second result, we let 6 operate on both sides of F.27), and get F.29) 62a = ^(Mi ®LBla-0i® 5LBia) = ^{(e#i + 0i)®LBia -J2e*Aej® LB}LBia} » i = ]T0xLBta + J2d6l®LBia-,^20iA Bi <g> LB}LBia. i i i,j We already have F.26), in which i,j can be interchanged. Therefore we get F.30) X^iW* = -^24k^k = 0. t i,k Here we have used commutativity of 0* and §j as well as c\k — -c?ki- By linearity of F.30) in §j, we get F.31) 53^LBia = 0.
6.5. CHARACTERISTIC CLASSES 275 Next we obtain F.32) ]T dOi ® LBla = - Y, r Y] c)kd} A $k <g> LB.a = -J2Y,cU9JA0><®LBi«- i j<k Finally, since we have LBiLBj ~ LBjLBi = L[Bi,Bj) = z2cijBk, k we obtain F.33) 53 ^ A 9i ® LBiLBxot = - 53h A ei ® 53c?L^d- From F.29), F.31), F.32), and F.33), we finally get 62a = 0, concluding the proof. ¦ 6.5. Characteristic classes (a) Weil homomorphism. Given a connection in a principal G bundle n : P —¦ M, there is determined a homomorphism F.34) w:W{q)-*A*{P) from the Weil algebra W(q) of the Lie algebra g into the de Rham complex A*{P), and it is commutative with exterior differentiation. In general, when a fiber bundle 7r : E -* B is given, the map 7r* : .4*(B) -+ «4*(?) turns out to be injective (see Exercise 6.3). If a differential form n e A* (E) is the pull-back of a differential form on the base, that is, if 77 = tt'tj, where f) € A*{B), we say that n is basic. Lemma 6.44. Let n : P —» M be a principal G bundle. A form n € A*(P) is basic if and only if the following two conditions hold: (i) i(A*)n = 0 for an arbitrary A € 0. (ii) R*n = 77 for an arbitrary g G G.
276 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES If G is connected, condition (ii) can be replaced by (iV): La'V = 0 for an arbitrary A e g. Proof. It is clear that the two conditions above are satisfied if t) e Ak{P) is basic. We prove the converse. Let xi,...,xn and V\, ¦ ¦ •, Vm be local coordinates for M and G. We may express v ? Ak(P) locally in the form F.35) 77 = 2ja/j(x,y)dXi1 A • • • A dxlr A dyn ¦ ¦ ¦ A dyjs, where / = (ii,..., ir) and J = (j\}..., js) run through all combinations with r + s = k and ajj(x,y) are functions of Xj, Vj- If 77 satisfies condition (i), then all s in F.35) are 0 and hence V = ^ a/(z,2/)<&tj A • • • A dxik. / Next, condition (ii) implies that each coefficient a/(x,y) is a function of Xi's only and is independent of j/j's. We thus conclude that 7? = Y^j a[(x)dxil A • • • dxik, which is basic. When G is connected, we show how condition (ii) follows from (ii'). From Exercise 2.8 in Chapter 2, we see that La*V = 0 implies ^expt>4?7 = 77. Since G is connected, we also know that an arbitrary element g G G can be written as a product expAi ¦ • expAk, where A\,...Ak eg. Hence R*n = 77. ¦ In view of Lemma 6.44 it is natural to make the following definition. Definition 6.45. Let G be a Lie group and q its Lie algebra. An element x of the Weil algebra W(g) is said to be basic if it satisfies the following two conditions: (i) i(A)x = 0 for all elements A € g. (ii) g*x = x for all elements g G G. We denote by 1(G) the set of all basic elements in W(g). We note that if G is connected, condition (ii) is equivalent to (ii): Lax — 0 for all A e g. By condition (i) above, no element of 1(G) contains a component of A*g*. Therefore 1(G) is a subalgebra of the set S*g of all polynomial functions on g. From the definition of the action of G on S*q*, we may rewrite 1(G) as follows. First, it is the set of all / € 5*g* such that g* f — f ¦ Second, it is the set of all polynomial maps / € 5*g
6.5 CHARACTERISTIC CLASSES 277 such that f(AdgA) - f(A) for all g e G and for all A G g. So we call elements of 1(G) invariant polynomials and I{G) the algebra of invariant polynomials. Denoting by Ik(G) the set of all homogeneous elements of degree 2k we have J(C) = fV(G), (Ik(G) c SV). fc=0 PROPOSITION 6.46. For an arbitrary element f e 1(G) we have <5/ = 0. PROOF. First for an arbitrary element a € S1^* we recall we defined Set = J^0t <8> ?#<<* (see F.27)). Now set ?> = ^ 0i ® Lb, • Think of it as an operator and write 8a = Da. Since 8 is an- tiderivation, we have 8(xy) = (8x)y+(-\)d^xxEy) for any elements x,y € S*g*. Since degx is even, we get 8(xy) = (Sx)y + x(8y). On the other hand, L^ is differentiation, we get D(xy) = f^Oi ® {(LBtx)y + x(LB<y)} = (?>x)y + x(?>y). i=l Since E and D coincide on S1g* and obey the same rules, we conclude that <5 = D on the entire space S*g*. Now for / e /(G), condition (ii') in Definition 6.45 implies that LBl/ = 0. Hence which finishes the proof. ¦ Now observe that the homomorphism w : W(g) —> A*(P), when restricted to 1(G), induces a homomorphism w : 1(G) —> .4*(M), because w takes basic elements to basic elements by virtue of Lemma
278 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES 6.40 and 6.41. Furthermore, since any / e 1(G) is closed relative to <5, we can take cohomology and obtain a homomorphism F.36) w : 1(G) -* H"DR{M) = H*(M : R), which is called the Weil homomorphism. Theorem 6.47. (Principal theorem of Chern-Weil theory) Let ? = (P, 7r, M, G) be a principal G bundle. If we choose a connection in ?, then the Weil homomorphism w : 1(G) —¦ i/*(M;R) is determined. This is independent of the choice of connection. If we set f@ = w(f)eM2k(M;R) for an arbitrary element f € Ik(G), then we get a characteristic class of the principal G bundle. PROOF. We first prove that the Weil homomorphism does not depend on the choice of connection. Let uq and u>\ be two connections on ?. Set J = [0,1] and write ? x / for the pull-back of f by the natural projection p : M x I —* M. Then f x / is a principal G bundle over M x I. Let Co be a connection on f x / such that cj|m><{o} = ^o and oj\mx{i} = wi- To see that such a connection Q in f exists, we may, for example, apply the construction of connections in the proof of Proposition 6.38, in the open covering M x [0,1/2], M x A/3,2/3), M x A/2,1]. More concretely, we set ut = A - t)cj0 + tu)\ (t € /) and see that each ut is a connection on ?. Let H^ C TUP (u G P) be the horizontal subspace relative to the connection u>t. Now we set at each point (u, ?) € P x I Hiult)=Htu®TtIcT{Utt)(PxI), and get a connection we wanted. In this case, the following diagram is obviously commutative: ¦* #*(Mx{0};R) - //*(Mx/;R) - ff*(Mx{l});R). 1(G) II 1(G) II 1(G)
6.5. CHARACTERISTIC CLASSES 279 Here wo>wi,w axe the Weil homomorphisms defined by the connections u>o,uji,u>, respectively. The maps i0 and ii are natural inclusions. If we naturally identify M x {0} and M x {1} with M, then the composite map i\ o ig ] is the identity map and hence vjq = t^j, showing that the Weil homomorphism is independent of the choice of a connection. Next we show that w(f) (/ € I{G)) is a characteristic class of the principal G bundle. Let Pi —*— P2 I I Mi » M2 h be a bundle map of principal G bundles. If u> is any connection on P2, we see that h*u> is a connection on Pi. Thus w\(f) = h*(w2{f)), which concludes the proof. ¦ (b) Invariant polynomials for Lie groups. We are giving a concrete description of the Weil homomorphism F.37) w:I(G)-^H*{M;R) of a principal G bundle ? = (P, 7r, M,G). For a given connection on ?, let u> and ft be the connection form and curvature form. For a fixed basis B\,..., Bm of g we write w = u>iPi + h u}mBm, Q = Q1B1 H 1- fim5m. On the other hand, we know that if 0i,... ,0m is the dual basis of 9*, then S*g* is nothing but the polynomial ring R[$i,... ,$m) generated by the corresponding elements 9\,...,0m € S1q*. Therefore it follows that any element / G Ik{G) can be written in the form In this case, C.67) is given by F.38) w{f) = (%2 ajili, A • • • A Oik] e H2k(M- R). 1 On the right-hand side of the last equation, the differential form between the braces is a basic closed 2k form on P. We think of it as a closed form on the base space M and take its deRham cohomology class.
280 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES At this point we want to introduce a somewhat different view point of I{G). Let us denote by Skg* the set of all multi-linear maps / : g x ••• x g -> R, that are symmetric, that is, f(Aa{l),--- ,Aa{k)) = f{Au-- ,Ak) for any A\y..., Ak G g and any permutation a of A,..., k). Given two elements / G Skg* and g € S'g*, we want to define a product fg e Sk+eg* by F.39) fg(Au...,Ak+e) = 77—TTf Yl f(A^ih---tAaik))g{A<r{k+i)l...tAa{k+e)). In this way, S*g* = ®kSkg* becomes a graded algebra. Proposition 6.48. For f e Skg* we define f e Skg* by f(A) = f(A,...,A) (A eg). Then this correspondence: Skg* —* Skg* gives an isomorphism S*g* * sv between graded algebras. Proof. We can easily see that Skg* 3 / ¦-»€ fSkg* preserves product, namely, fg = fg. Next for any monomial / = 9{l ... >0lk € SkQ* we define / € Skg* by setting F.40) f(Au...,Ak) = ± ? A,(V))-.^Ma(fc)) * aeek This linear map Skg* —* Skg* turns out to be the inverse of the first map. ¦ Henceforth we identify Skg* and 5*g* by the isomorphism in Proposition 6.48. In particular, an invariant polynomial / € Ik{G) on G is characterized as a symmetric multi-linear map / : g x ••• x g -> R
6.5. CHARACTERISTIC CLASSES 281 that is invaxiant by Ad G: f(AdgAi,..., AdgA^) = /(A1(..., Ak) for all g e G and Ai e g. We shall have another look at Weil homomor- phisms F.37) from our new point of view. For each / € /fc(G), as we know, its image w(f) € H2k(M;R) is represented by the closed form on M between the brackets on the right-hand side of F.38). On the other hand, the exterior power of the curvature form ft 6 A2k{P\$) is written ft* = ft A • • • A ft € A2k{P; 0 ® • • • <g> g) = A2k(P; g®*). If / is a symmetric, AdG invariant multi-linear map: g® —> R, the composite map / o (ft10), written also /(ftfc), is a 2k form on P. To be specific, we have for X\,..., X2k € TUP F.41) /(ftfc)(X1,...,X2Jt) = —— ]T SgnG/(fi(^a(l).^<7B)).---»^(^<7Bfc-l),^aBfc)))- ^ '' a€S2fc As a matter of fact, we shall see that /(ftfc) coincides with the differential form (that should be written /(ft) in the notation of §5.4) within the brackets on the right-hand side of F.38) and hence it is a closed form on M expressing w{f). Lemma 6.49. The two 2k forms on P defined above coincide, that is, mk) = w(f). PROOF. More generally we shall prove that the diagram below is commutative: SV -^ A*(P) 5*g* -JL- A'{P). Here the homomorphism w is defined by F.41). By linearity it is sufficient to prove the assertion for the elements of the form f = $il ¦¦¦§'* eSkg\ By computation using F.40) and F.41) we obtain for Xi,..., X2k € TUP
6. FIBER BUNDLES AND CHARACTERISTIC CLASSES = 7^! E Sgn^/(«(^tl),^B)),.--,n(^Bfc-l),A:aBJfc,) <76©2k a662fc * TG©»e t2>(/)(x1)...,x2fc) = /(nfc)(x1,...)x2,) 1 B*)! _ _1_ ~ Bky = T2k)\ ^ sgna'^ii^'7A)'^?7B))---fiifc^B;t-1)'^Bfc)) = (nllA---Anlk)(xl,...,x2k) = w(f)(Xu...yX2k), which finishes the proof. In the above equations, T°,..., T% are shorthand for n(xffA)lxaB)),... ,n(^Bfc-i),^Bfc))- ¦ (c) Connections for vector bundles and principal bundles. We now compare the bundle connections for vector bundles in Chapter 5 and the connections for general principal bundles. The conclusion: they are exactly equivalent. For simplicity, we deal with real vector bundles and the principal bundles that are associated. The case of a complex vector bundle differs slightly. THEOREM 6.50. Let n : E —» M be an n-dimensional vector bundle and n : P —+ M the associated principal GL(n;R) bundle. Then there is a natural one-to-one correspondence between the set of connections on E and the set of connections on P. Proof. For simplicity we write G and g for GL(n\ R) and gl{n; R), respectively. We prove the assertion in the case where E is trivial. Choose a trivialization <p : E = M x Rn and let si,... ,sn be the corresponding frame field. Now for a given connection V on ?, we write Vsj = y^} ® 5»> as in §5.3. In this way, a g valued 1-form, namely, the connection form u = u>* is determined so as to contain all information on the connection. Conversely, given an arbitrary g valued 1-form u> can be chosen as the connection form of a certain connection. In this way,
6.5. CHARACTERISTIC CLASSES 283 we may say that the totality of connections on E is identical with Next we shall determine a connection in the principal bundle P by using a connection form iv € A1{M;q). A trivialization <p : E = M x Rn induces a section s : M —> P, and furthermore a trivialization ofP is determined by the correspondence M x G 3 {p,q) *-* s{p)g e P. Set . u> = v*(Ad(g~l)Lj + uJo) e A^P.g) {g € G). Here u>0 is the Maurer-Cartan form of G, and <I> turns out to be a connection of P. In fact, condition (i) in Theorem 6.37 is clearly satisfied because an equivalent condition F.10) is satisfied by <I>, and (ii) is easily verified. In this way, a map i is defined from the set of all connections on E into the set of all connections on P. The map is obviously injective, and it is surjective, because an arbitrary connection on P is completely determined by the 1-form on M that is the pull-back by the section s and it is also the image by i. We show next that l does not depend on the trivializations </? : E = M x Rn. This will complete the proof of the theorem. Indeed, each of the connections on E and P can be thought of as successively pasting connections on open subsets ?/, over which the bundle is trivial, so that on each U i is bijective. Suppose i> : E = M x Rn is another trivialization with corresponding connection form u'. We follow the same construction as above and let ij' = ^*(Ad(<7-1)u/ +u>o) be the resulting connection on P. We want to prove that Cj — u>. Let g^ : M -+ G be the transition functions between trivializations V and rp (§5.1 (b) and §6.1 (b)). Then by Proposition 5.22, §5.3, we get F.42) J = g'l^n, + g'^dg^ Now we consider the following commutative diagram:
284 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES P —?—> M xG II TidxL, P —^— MxG. Here id x Lg(p,g) = {p,g^{p)g) (p € M,g e G). It suffices to prove F.43) (id x Lg)*{Ad{g-l)uj + u;0) = Ad(g~l)u>' + u0. Clearly, we have F.44) (id x LpJ'Adfo-1)* = Ad(g-l)(g-lujg^). On the other hand, the Maurer-Cartan form can concretely be written cj0 = 9~ldg {g e G). (Verify this in Exercise 6.8.) Therefore we get <* ac\ (id x Lfl)*w<> = {9^9)~^d{g^g) F.45) , , = Ad(p l)g^,dg^+uo. Comparing F.42), F.44) and F.45), we see that F.43) follows and the proof is over. ¦ We have just finished showing that the characteristic classes for vector bundles in Chapter 5 and the characteristic classes of associated principal bundles in this chapter are totally equivalent. (d) Characteristic classes. Our discussions so far show that, for a given Lie group G, any element of the algebra 1(G) of invariant polynomials plays the role of a characteristic class of the principal G bundle. As we proved in §5.4 (Theorem 5.26), we have I{GL(n;R)) = R[au...,an\. /(GL(n;C))=R[c1,...,cn]. We mention several more examples without proof. Example 6.51. I{0{n)) = R[pi,p2,• • • ,P[n/2\\- Example 6.52. I{SOBn)) = R[pll...,pn_llett].
§6.6 A COUPLE OF ITEMS 285 Example 6.53. I(SO{2n+ 1)) = R[pi,...,pn]. Example 6.54. I(U(n)) = R(ci,...,c„]. These four examples are compact Lie groups. In general, for a compact group, it has been proved by H. Cartan that the set 1(G) of all invariant polynomials and the set of real-coefficients characteristic classes of the principal G bundle can naturally be identified. §6.6 A couple of items (a) Triviality of the cohomology of the Weil algebra. Let g be the Lie algebra of a Lie group G. When the principal G bundle is given a connection, the Weil algebra W(g) is to serve as a model of the subcomplex of the de Rham complex of the total space generated by the connection form and the curvature form. Therefore we could say that its cohomology group H*(W(q)\$) represents the cohomology of the entire space of the principal G bundle. Our next theorem asserts that it is indeed trivial. In the language of the classifying space BG for G, this corresponds to the fact that the total space of the universal principal G bundle is contractible to 0. THEOREM 6.55. The cohomology group of the Weil algrebra W(q) relative to exterior differentiation 5 is trivial, that is, (R (fc = 0) "•™ = {o (*>0). Proof. Choose a basis B\,..., Bm of g and let 6i,..., 6m be the dual basis in g*. Recall that we have W(g) = AV ® SV = ?@i, ¦ ¦ ¦, *m) ® R[0i, • • •, 0m}. Here E denotes the exterior algebra. Now for each homogeneous element x we define its 'weight' f(x) as the largest integer A; such that x has a non-zero component in A*g* ® Skg*. For example, iFi) = 0J{6iej) = l,l{et6j) = 2, etc. We also set ?@) = -1. We also define a linear map k : W(g) - W(g) by setting k(qc) = 0 for any a in g*, /c(a) = a for any a G Slg*, and extending it to the whole space W(q) as antiderivation of degree -1. If we set ?> = «S + <5k : W(g) - W{g),
286 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES then this is a derivation of degree 0, because for any x, y € W(fl) we have D{xy) = k{6x ¦ y + (-l)degIx • 6y} + 6{kx ¦ y + (-l)degIx • «y} = (/c<5x)y + x{K6y) + {Snx)y + x(<5/cy) = (Z>x)y + x(Dy). Now for any a 6 9* we show F.46) Da = a, Da = a - da. As for a, we have Da = (k6 + 5n)a = n(da 4- a) = a. As for d, we get Da = *(]?^0i <S> Laid) + 5a = - ^ 9i ® /c(L #. d) + da + a = — >J #t A Lj5,a + da + d = a - da, completing the proof of F.46). We have used J2i ^t A ^BcC* = 2da. Suppose x is a homogeneous element of degree m > 0. Then Dx is also a homogeneous element of the same degree. Obviously, ?{x) < m. Set x' = x -—-tDx. m - ?{x) We can show F.47) ?{x') < ?{x). as follows. We know that x is a linear combination of elements of the forms 0ir--9lkeh--9jt {k + 2? = m). If we introduce the notation ej=0i,•¦¦***. §./= **¦¦¦**. then we can write D(eIeJ) = D(eI)eJ + eID(Qj) = {k + ?)QiQj + terms with weight less than ?,
§6.6 A COUPLE OF ITEMS 287 by using the properties F.46) of the derivation D. If we set ? = ?{x), then k + ? = m - ?(x). Hence ?{x') < ?(x), which proves F.47). Let us furthermore assume Sx = 0. Then X' = X rrr-rDx = X - 6 { TT-rKX J m-?{x) \m-?{x) ) so that Sx' = 0 and x' is cohomologous to x. If we do the same argument starting with x', we get an element x" cohomologous to x' such that ?{x") < ?{x'). By repeating this procedure a finite number of times, we arrive at an element with ? = — 1, that is, 0. This shows that the original element x is cohomologous to 0, thus proving the theorem. ¦ Remark 5.56.. The proof above is constructive in the following sense. Namely, for any cocycle x G W(q) with positive degree, we can actually construct an element y €.W(g) such that Sy = x. (b) Chern-Simons forms. As we have seen by now, given an invariant polynomial / G Ik{G) for a Lie group G, any principal G bundle f = (P,n,M) gets a characteristic class /(?) G H2k(M;R). Furthermore, this characteristic class is represented by a basic, closed form f{?lk) induced by the curvature form Cl of an arbitrary connection in P. Now what happens if we take the de Rham cohomology not on the base manifold M but on the total space P? The answer: we get 0. This can be seen as follows. The pull-back n*? of f itself by the projection tc : P —¦ M is a trivial principal bundle by Example 6.8, thus f(n*?) = 0. On the other hand, relative to the pull-back connection from ? onto 7r*f we see that f{n*?) is represented precisely by f{?lk). Therefore the cohomology is 0, that means that it is an exact form on P. This fact also follows from Theorem 6.55. Because Sf = 0, Theorem 6.55 says that there is an element Tf G W{q) such that STf = /. Therefore the image w{Tf) G A2k'l{P) satisfies dw(Tf) = /(Hfc). It was Chern and Simons that gave a concrete expression for such a differential form Tf. Nowadays they are called Chern-Simons forms. They are indispensable tool in research of low-dimensional topology and gauge theory. Unfortunately, we have no more space to go into any of these topics. Interested readers are referred to the original paper Ann. of Math. 99A974), pp.48-69. (c) Flat bundles and holonomy homomorphisms. We start with flat bundles.
288 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES Definition 6.57. Let ? = (P,7r,M,G) be a principal G bundle over a C°° manifold M. A connection w on ? is said to be a flat connection if its curvature form fi is identically 0 (H = 0). A principal G bundle is called a flat G bundle if it has a flat connection. Two flat G bundles ?i and ?2 over the same base space M are to be isomorphic if there exists a bundle isomorphism / : P\ —> P2 such that /*u>2 = u)\. Here, of course, Pi is the total space of ?u, and u^ is a flat connection in &, with i = 1,2. As an example of flat bundle, the product bundle M x G has a trivial connection, which is uninteresting. However, depending on the base M and the structure group G, there can be many non-obvious flat G bundles, whose study is often an important problem. The geometric meaning for a connection u; to be flat can be explained by using the theorem of Frobenius (§2.3, Theorems 2.17, 2.21) as follows. For each point u € P, we define Hu = {XeTuP;uj{X)=0}. Then H is a ditribution consisting of all horizontal vectors relative to u>. We have Proposition 6.58. For a connection u on a principal G bundle ? to be flat it is necessary and sufficient for the distribution H to be completely integrable. PROOF. By Theorem 2.17, H is completely integrable if and only if it is involutive. Hence it suffices to show that u> is flat if and only if H is involutive. Denote by T(H) the set of sections of H, that is, the set of all horizontal vector fields on P. If X and Y are arbitrary horizontal vector fields on P and X/, and Yh their horizontal components, then Proposition 6.39 (ii) implies 2Q.{X,Y) = 2du{Xh,Yh) = Xh{u{Yh)) - YhMXh)) - u([Xh, Yh}) =-u(lXh,Yh)). It follows that Q = 0 if and only if [Xh,Yh] € T{H), that is, H is involutive. ¦ We study the structure of flat bundles by using Proposition 6.58 and its terminology. Suppose ? is a flat connection on P. Choose a
5.6 A COUPLE OF ITEMS Figure 6.9 point po 6 M as reference point and pick a point uo € 7r_1(Po)- Let LU0 be the maximal integral manifold of H through u0. We denote the restriction of ir to LUo by ttq : LUoM. For any point on LUo, 7r0 is obviously a diffeomorphism on a certain neighborhood. This means that 7r0 is a covering map. Hence for any closed curve a in M with starting point po, there is a unique lift d starting at uq and lying in the integral manifold Luo. (See Figure 6.9.) The end point of a lies in the same fiber 7r_1(po) as uo- Therefore there exists an element g € G such that the end point of d can be expressed as uog. Furthermore it is easy to show that the end point of d does not vary if we continously move a while fixing po- Therefore the element g above depends only on the element (written, say a for simplicity) of the fundamental group ttiM relative to the reference point po determined by the curve a. Thus by setting p(a) = g~r we can define a map p : n\M —* G. We now show that p is a homomorphism, which we call a holo- nomy homomorphism. PROPOSITION 6.59. The map p : tk\M —> G is a homomorphism. PROOF. Let a,P € tt\M and let a,C be their lifts starting at u0 into LUo. By definition of p, we see .that the end point of a is uop(a) ~l and the end point of/? is uop(fi)~l. Since H is invariant by the action of G, we see that 0p(a)~l is the lift of 0 with initial point uop(or)-1 into LUo — LuoP(a)-i. Hence we may take d • @p(a)~l) as the lift of a/3. Now the end point of 0p{a)~l is equal to uop{0)~lp(a)~l.
290 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES uop@)-yp(a)-1 uop(a)'1 uop(/3)~l Figure 6.10 Therefore that is, p is a homomorphism. ¦ Next, we shall see how the holonomy p will change if we replace the point uo € tt-1(Po) by another point in the same fiber u'0 € 7r_1(po)- Then we can write u'0 = uoh. Let p' be the holonomy corresponding to u'0. Let a (resp. a') be the lift of a (resp. a') into LUo (resp. L^) with initial point uo (resp.u0). Then, obviously, a' = ah. Hence u'op(ot)-1 = u0p(a)_1/i. On the other hand, we have u'op'ict)'1 =«oV(a)_1- We conclude p'(a) = h~lp{a)-lh, namely, p'(a) = /i"V(a)/i. That is, p' and p are conjugate homomorphisms. FVom the discussions above we see that if two flat G bundles over the same base space M are isomorphic, then their holonomy homomorphisms are conjugate to each other. Hence for each isomorphism
class of a flat G bundle over M, the homomorphism, called the ho- lonomy, p : ttiM —> G, is defined up to conjugacy class. In fact, the structure of a flat bundle is completely defined by its holonomy as follows. Theorem 6.60. Let M be a C°° manifold and G a Lie group. Then the map that associates to any flat G bundle over M its holonomy is a one-to-one correspondence between the set of isomorphism classes of flat G bundles and the set of conjugacy classes of homo- morphisms p : 7TiM —» G. For the proof of this theorem and its applications, see the sequel to this volume, S. Morita, Geometry of Characteristic Classes, AMS, 2001 Summary 6.1 A fiber bundle is made up by manifolds, called fibers, one to each point of a manifold, called the base space, in such a way that lcally it is a product. The way the fibers are arranged is described by a transformation group, called the structure group. 6.2 A fiber bundle is called a principal bundle if the structure group is a Lie group together with its natural action on the fibers to the left. 6.3 An oriented Sl bundle is completely classified by its Euler class. 6.4 For a vector field with a finite number of singular points on a closed C°° manifold, the sum of the indices at singular points is equal to the Euler number. This is called the Hopf Index Theorem. 6.5 A connection in a principal bundle is a distribution of horizontal subspaces at each point of the total space that is invariant by the action of the structure group to the right. 6.6 The curvature is defined as the derivative of a connection. It describes the way a principal bundle is curved. 6.7 For a Lie group G with Lie algebra g, W(q) = A*g* <g> S*q* is called the Weil algebra of g. W(g) plays the role of the de Rham complex of the total space of the principal G bundle. 6.8 For a Lie group G the subalgebra I{G) of all basic elements of the Weil algebra is called the algebra of invariant polynomials of G. For any principal G bundle, there is a homomorphism
292 6. FIBER BUNDLES AND CHARACTERISTIC CLASSES from 1(G) into the cohomology algebra of the base space. It is called the Weil homomorphism. Elements that are images of the Weil homomorphism are characteristic classes of the principal G bundle. 6.9 The notion of a connection of a vector bundle and that of a connection in an associated principal GL(n;R) bundleare equivalent. 6.10 A principal bundle with a connection is called a flat bundle if the curvature is 0. Exercises 6.1 We define a generalization of the Hopf map h : S2n+l —* CPn of the Hopf map h : 53 -> S2 (Example 1.27) as follows. For 52n + 1 = {(zlf..., *n+1); ?. |z,|2 = 1}, set h(zu..., zn+1) = [z\,..., ?n+i]. Prove that this is a principal 51 bundle. It is called the Hopf 51 bundle. 6.2 Let ? = (?, n, ?, F, G) be a fiber bundle. Let M be a C°° manifold and / : M -* B a C°° map. Set f*E = {(p, u) e M x E : f(p) = 7r(u)}. Prove that f"E is naturally a C°° manifold. Define f'rc{p,u) = p, and show that /*@ = (/*?\/*7r, M, F, G) is a fiber bundle. 6.3 Let 7r : E -> S be a fiber bundle. Show that tt* : .4*(?) — A*(E) is injective. 6.4 Prove Lemma 6.19. 6.5 Prove that the Euler classes of an oriented 2-dimensional vector bundle and those of an oriented S1 bundle coincide. 6.6 Directly from the definition by means of differential forms, prove that the Euler class of the Hopf S1 bundle is equal to -1 € H2(S2,Z). (See Example 6.29 and Exercise 6.1.) 6.7 Give a concrete description of the Weil algebra W(soB))) of SO{2). 6.8 Show that the Maurer-Cartan form u> of GL(n; R) can be written in the form ui = g~ldg. 6.9 In the Weil algebra W(su{2)) of 51/B), find the Chern-Simons form of C2 € I(SU{2)), namely, an element Tc2 such that STc2 = c2.
EXERCISES 293 6.10 Construct a non-trivial flat connection for the product bundle 7r : S1 x S1 —> Sl (where n is the projection onto the first component).
Perspectives We shall briefly summarize the roles played by the theories of differential forms and characteristic classes in geometry today. The starting point was the introduction of characteristic classes such as Stiefel-Whitney classes and Pontrjagin classes that describe, in the language of cohomology, the way a differentiate manifold is globally bent. Prom the point of view of differential forms, we may say that the de Rham theorem and the Chern-Weil theory are two pillars. That is, we grasp the cohomology of the manifold by using differential forms and then express the way the manifold is bent in terms of the curvature of a connection we introduce in the tangent bundle. Finally, by integrating the curvature form over various cycles we obtain characteristic classes and characteristic numbers as global invariants. We touched slightly upon the cobordism theory initiated by Pontrjagin and Rohlin and completed by Thorn. The idea of cobordism is quite natural, and has appeared time and again in various forms in topology. Thorn had the epoch-making idea of classifying all closed manifolds by an equivalence class called cobordism, reduced it to the computation of the homotopy groups of a space called the Thorn complex, and finally solved the problem by the techniques of algebraic topology. The final answer, as stated in Theorem 5.54 of §5.7, is that the Pontrjagin numbers and the Stiefel-Whitney numbers are complete invariants. This work of Thorn was a pioneering achievement and a model of progress in the field called differential topology that flourished in the 1950s and 1960s. First, Milnor constructed a non-standard differen- tiable structure on S7. Shortly after, Smale achieved the /i-cobordism theorem and the solution of the general Poincare conjecture. Around the same time, Milnor and Kervaire built the classification theory for homotopy spheres. Novikov and Browder started the theory of 295
296 PERSPECTIVES surgery, by which they concretely built and classified diffentiable manifolds. The theory was completed by Sullivan and Wall. Pontrjagin classes played important roles in these flows. In fact, the basic technique was to freely modify manifolds by process called surgery while keeping global structures in mind. And the work by Kirby and Sieben- mann in 1969 reached the pinnacle in differential topology: the final achievement on triangulation of topological manifolds. As stated in the text, Hirzebruch used the work by Thorn almost immediately to prove the signature theorem that bears his name. It says that the signature, which is a cohomological invariant of a manifold, can be expressed by the Pontrjagin numbers. This result itself belongs to differential topology, but its idea turned out to be far-reaching. That is, right after the signature theorem, Hirzebruch himself proved the Riemann-Roch theorem for algebraic varieties, and Atiyah and Hirzebruch generalized it to differentiable manifolds. The famous Atiyah-Singer theorem was published in 1963. It says that an invariant called the analytic index can be defined for a certain system of elliptic differential operator acting on various vector bundles, and equals the topological index that can be described by the characteristic class. It is a truly beautiful result - a model of a theorem in mathematics. Hirzebruch's Riemann-Roch theorem has since been deeply generalized by Grothendieck so as to include maps for algebraic varieties. We might say that it is one of the deepest theorems in mathematics of the 20th century. In its formulation as well as in various index theorems, /(-theory, also created by Grothendieck, plays an important role. Very briefly, /(-theory is a theory dealing with an abelian group structure in the set of certain equivalence classes of vector bundles over a manifold. With this theory we can clearly explain Bott's periodicity theorem and a deep theorem of Adams on topology of spheres. What we mentioned as Grothendieck's work is only a portion of his huge amount of achievement, most of which belongs to algebraic geometry. For these results, see K. Ueno, Algebraic Geometry 1,2,3 in this series. In any case, Grothendieck's work is unparallelled in its depth and magnitude. Let us go back to the relationship between the geometry of manifolds and characteristic classes. As we already said, Pontrjagin classes play an important role, but it is as a cohomology class and not as a differential form itself. So we cannot quite say that Chern-Weil theory
PERSPECTIVES 297 itself plays a direct role. However, the appearance of Bott's vanishing theorem and Gel'fand-Fuks cohomology theory in around 1969 changed all that. By using connection theory in vector bundles, Bott proved the existence of a new topological obstruction for a distribution on a manifod to be completely integable. On the other hand, Gel'fand and Fuks constructed the cohomology theory for an infinite- dimensional Lie algebra formed by all vector fields on a manifold. Both of these are applied to geometry by a certain design formed by a foliation structure. By a concrete refinement of the classical Chern-Simons work, the theory of secondary characteristic classes for foliated structures was born. The theory of general secondary characteristic classes and the geometry of flat bundles are among the themes currently being pursued actively. See my book Geometry of Characteristic Classes, in this series. As we entered the 1980s, the geometry of manifolds saw a radical change, which was triggered by Donaldson's famous work. He considered principal bundles over 4-dimensional manifolds, and by studying the space of all connections he drew a startling conclusion about the topology of smooth 4-manifolds. Then by combining it with Freedman's solution of the 4-dimensional Poincare conjecture, it was proved that on R4 there exist many differentiable structures different from the usual one. This was followed by the work of Floer, Jones, and Witten. At this time of writing, 1997, the research goes on as vigorously as ever. As a common feature of all these works, we see attention to infinite-dimensional objects and global analysis of concrete partial differential equations on manifolds. Beyond these, substantial roles are played by Riemannian metrics, connections, curvature, and differential forms. The Atiyah-Singer index theorem, and in particular, the index theorem for families of elliptic operators on fiber bundles, is being seriously applied. Where are we going with such vigorous research? It is hard to tell. When you try to get a finite number from an infinite-dimensional object, what seems to work most effectively is physical ideas, as in Witten's work. The attempt to take out combinatorial sides and reconstruct mathematical objects has been fairly successful and has produced a large amount of topological invariants for 3-dimensional manifolds. However, the geometric meaning of such invariants remains
298 PERSPECTIVES unclear. If we think back and realize that it took 150 years from Eu- ler numbers to Euler-Poincare characteristic number, we might have to say that the story has just begun. In any case, there is no doubt that the true value of geometry, and of mathematics, is being seriously sought out. It is possible, just as Riemannian geometry in the early 1920s provided the theory of relativity a good foundation, that there may be a serious move under way to find a concept of space that can resolve various difficulties now encountered in physics. If so, what changes will manifolds and differential forms undergo?
Solutions to Exercises Chapter 1. 1.1 The function fm(z) can be written as fm{z) = gm{z)+ihm{z), where gm{z), hm(z) are the real part and the imaginary part of it respectively. On the other hand, since — (x + iy)m = m{x + iy)m~ \ q~(x + **/)"* = im(x + iv)™' l> we have ?5m = mReC^), ?-gm = -mlm^-1). Similarly for the partial derivatives of hm. Therefore the required Jacobian matrix is /mRe(zm-1) -mlm(zm-1)\ \mlm(zm-1) mRe^'1) )' 1.2 The set MB; R) of all real matrices of order 2 can be naturally identified with R4. Then 0B) is defined by the equation lAA = E {A € MB;R)). Here E is the identity matrix. If we let ¦C % the equation becomes a2 + b2 - 1 = 0, ac + bd = 0, c2 + d2 - 1 = 0. That is, 0B) is defined by the above 3 equations in R4. Then, the corresponding Jacobian matrix is Ba 26 0 0> dab v0 0 2c 2d J If we compute its four minors of order 3, they are Aa(ad-bc), Ab(ad-bc), 4c(ad-6c), 4d{ad-bc).
300 SOLUTIONS TO EXERCISES Since these do not vanish simultaneously on 0B), the rank of the Jacobian matrix on 0B) is constantly 3. Therefore by Example 1.13, we see that 0B) is a 1-dimensional C°° manifold. More precisely, we can prove that 0B) is the disjoint union of two circles 51. 1.3 CP1 is obtained from two copies of C by glueing each subset C - {0}, where we remove the origin, by the correspondence z *-> j. Now we define two maps f± : C —* 52 C R3 by /±B)" VTTW TTW ±TTffl) (with the double signs the same order). Then it is easy to verify that f+(z) = /_( i ) for an arbitrary non-zero complex number z € C. Therefore, a C°° map from CP1 to 52 is obtained from this. We leave it to the reader to verify that this map is in fact one to one and onto and its inverse map is also of class C°°. 1.4 An arbitrary element of 50C) which is not the identity is a rotation around a line through the origin in R3 by an appropriate angle, as is well known. Using this fact, we try to assign an element in 50C) to an arbitrary element in 53 that is a unit 4-dimensional vector (a, 6, c, d). If d = ±1, we let it correspond to the identity of 50C). If d ^ ±1, since (a, 6, c) is a non-zero vector of R3, a line of R3 through the origin in the direction of it is determined. Let the rotation angle be it if d = 0, and as d approaches ±1, let it approach 0 or 27r. If we specify an orientation of R3, the rotation angle can be determined for example by the right-hand system. In this way, a map from 53 to 50C) will be determined, and furthermore it is easy to see that ±(a, b, c, d) are mapped to the same element by this map, so that we would finally obtain a map RP3 —> 50C). Here we used the fact that RP3 is obtained by identifying every pair ±(a,b, c,d) in 53. To verify that this conjecture is actually right, we argue as follows. First of all, consider the case of (a,0,0,d) (that is, the case where it can be written (sin 9,0,0, cos 9)). In this case we let it rotate around the x-axis by the angle 29. Then it goes well, because, if a = 0,1, 29 = 0,7r. The corresponding matrix is 1 0 0 0 d2 - a2 -2ad 0 2ad d2 - a2 by a special case of the addition formula in trigonometry. If we argue in the same way for the cases of a = b — 0 and a = c = 0 and compare the corresponding matrices, we see finally that a general
SOLUTIONS TO EXERCISES element (a, 6, c, d) should correspond to the matrix (a2 - 62 - c2 + d2 2ab - 2cd 2ac + 2bd 2ab + 2cd -a2 + b2 - c2 + d2 -2ad + 26c \ 2ac - 2bd 2ad + 26c -a2 - 62 + c2 + d2) belonging to 50C). We leave it to the reader to check that the map thus obtained becomes a diffeomorphism. 1.5 If we define a map / : M —> M x N by f(p) = (p, /(p)) (p g M), obviously Tj = Im/. Let n : M x N —» M be the projection to the first factor. Then we have n o / = id^. Therefore the differential of / is an injection and we see that / becomes an immersion. Furthermore, it should be easy to check that / is an embedding. 1.6 We use the fact that each component of the map L is expressed by a linear function in the coordinates xi, • • • ,xm of Rm. 1.7 If one uses the local expression A.10) of the bracket, the proof of (i), (ii) may be easy. The proof of (iv) is not so difficult. Here we prove only the Jacobi identity (iii). By the linearity of (i), it is enough to prove the case where the local expressions of X,Y, Z are X = f— Y = g— Z = h d dxi dxj dxk respectively. Then we have [[X,YIZ] =(ffrhj ~ 9fA)-^~k ~ Wk9i + hf9ik)-^ + (hgkfj + h9fjk)-^-- Here, for example, & stands for -—g. If we sum up all those expres- OXi sions, each of which is obtained by applying a cyclic permutation on f,g, h and i, j, fc, we see that the given formula becomes zero. 1.8 Let xi, • • • , xm and yi, • ¦ • , yn be coordinate functions around the points p and f(p) respectively. By the linearity of the problem, we may assume that v = -—. Then we have OXi dh Mv)hssym.™ On the other hand, by the formula for the differential of a composition, we see that -z—(h o f) is also equal to the r OXi above formula, so that the given formula holds.
302 SOLUTIONS TO EXERCISES 1.9 Let X\, • • • , xn and y\, • • • , yn be two positive local coordinate systems defined near the point p € M. Then the Jacobian detf —- J is positive. Here if p € dM, then xn = yn = 0, so that the Jacobian matrix is / M ... fr/i flvi \ / 0u ai„_i ain \ V o - o f?/ Since obviously -^— > 0, we have detf ^—) > 0. There- OXn \OXj' l<t,j<n-l fore, all such xi, •• • ,xn_i give an orientation on dM. 1.10 First of all, we shall see that RPn is a manifold obtained by identifying each point p 6 Sn with its antipodal point -p in an n- dimensional sphere Sn. Next, if we define / : Sn —» 5n by /(p) = -p, we see that this is an orientation reversing or preserving diffeomor- phism according to n being even or odd respectively. This is because / can be extended to a diffeomorphism / of the whole Rn+1 by the same formula, and since /, ( -—) = — -—, / preserves the orienta- V dxi / oxi tion if n is odd, and reverses the orientation if it is even. On the other hand, / obviously maps the outward normal vector of Sn to the outward normal vector. From this fact, the above property of / follows easily, and we leave it to the reader to deduce the claim from it. Chapter 2. 2.1 By the linearity, it is enough to prove the assertion for u> = fdxil A ¦ • • A dxH., 77 = gdxji A • • • A dxj,. If we use the equation dxj/\dx{ = —dxiAdxj repeatedly, A) is proved. On the other hand, from u> A 1) = fgdXi} A • • • A dxik. A dxjl A • • • A dXj, we obtain d{u) A 77) = (dfg + fdg)dXii A • • • A dxik. A dxjx A • • • A dxj,. B) follows from this. 2.2 Since <p*{b}/\-q) = y>*ujA<p*r), it is enough to prove the assertion in the cases where a; is a function / and dx{. First in the case of a
SOLUTIONS TO EXERCISES 303 function, since <p*{f) = / o <p, we have On the other hand, since df = V^ -—dxi, we have . OXi *>•(*) = Eg g%, and the claim is shown. Next we let w = dx^. Then obviously dw = 0, while we have ^•((fal)> - <e g*,) = E(E ggfe-*) a *, = o, and the claim is shown. Here we used the facts that the partial differential does not depend on the order of differentiation and that dyk Adt/j = -dy3 Adyk. 2.3 By the Cartan formula, Lx{u) A r?) = {i(X)d + di{X)){u> A 77). Here, using the fact that d and i(X) are anti-derivations of degree 1 and -1 respectively, we decompose the right-hand side of the above formula and arrange the result. Then (i) is proved, (ii) follows from Lxdoj = {i(X)d + di{X))dw = di{X)dw = d{i{X)d + di{X))w = dLxuj. 2.4 If n = 2, a direct computation shows that J1 = 2dx\ A dx2 A dx3 A dx^. Also, in the case of general n, a brief consideration tells us that un = n\dx\ A • • • A dx2n- 2.5 It is enough to show that an arbitrary fc-form on N can be extended to a fc-form on M. If we put Uo = M \ TV, this is an open set by the assumption. Also, by the definition of a submanifold, if we let the dimensions of M, TV be m, n respectively, then there exists a family {Ui}i>\ of open sets of M such that each Uj is diffeomorphic to Rm, UiHN = R" C R™, and {t/, D N}i>{ is an open covering of N. In this case we may assume further that {t/t},>o is a locally finite open covering of M. Let {/,} be a partition of unity subordinate to
304 SOLUTIONS TO EXERCISES this open covering. Now we put u>o = 0, and for i > 1 we choose eAk{Ut) such that It is clear from the form of Ui that this is possible. If we put t=0 this is the required extension of u to M. 2.6 It is enough to show that if /*w = 0, then u> = 0. By the definition of submersion, for an arbitrary point q € N and a point p e M such that /(p) = qf the map /* : TPM —> T^AT is a surjection. Therefore, the map /* : A*T*N -* A*T*M induced from it is an injection. Since f*v{p) = 0 by the assumption, we have u)(q) = 0. As q can be taken arbitrarily, we conclude that ut = 0. 2.7 Since ||x|| = y/x\ H hi^, we have <*IMI = TnT(xid:ri + • • • + xndxn). 11*11 Therefore, ^ = ~n l ii nWJ.o n^i A • • • A dxn + -r—r-ndxi A • • • A din = 0. ||x||n+2 j|x||n 2.8 By Proposition 2.13, A) LxW = iim*L^. Therefore if <p*tu) — u for all t, obviously we have Lxw = 0. Conversely, suppose Lxu = 0. Since the problem is local, we may assume that ip^uj is expressed as (p\u) = V^ fi(t, x)dxil A • • • A dx{k ti<--<tfc in a coordinate neighborhood. Here / = {i\, — - ,ik}- Then by A), we have Lxu = Y, ^¦@,x)dxil/\--'Adxik. ti<-<u Again by A), we see that Lx<p*su = <p*sLxw = 0 for an arbitrary s. Therefore, replacing ui by </?*u> in the above discussion, we see that f(...)-o
SOLUTIONS TO EXERCISES 305 for an arbitrary J. Since s was arbitrary, //(t,x) does not depend on t and thus is a function only in x, so that <p^u) = u>. 2.9 By the definition of the polar coordinates, we can write r = \Jx2 -f y2, 6 = arctan -. By a direct computation we have x dr = / o, o(xdx + vdy)' de = 12TT2(xdy ~»<&)¦ Observe further that rdr Ad9 = dx Ady. 2.10 We choose as a basis of the Lie algebra of SU{2). Then [BUB2] = 2B3, [B2, B3] = 2BU and [B3, BJ = 2B2. Therefore, if we let wi,w2,W3 be its dual basis, the required Maurer-Cartan equations are du>i = —a;2 A u;3, du>2 — —<^3 A u/i, du>^ = —c^i A u^- Chapter 3. 3.1 For B), if we pay attention to the boundary, the proof of Theorem 3.4 can be used without much change. Furthermore, if we refine the discussion there a little, we see that the non-triviality of Hn(M;Z) is equivalent to the orientability of M. A) follows from this. C) is easy. 3.2 A bounded closed interval [a, 6] is a 1-dimensional differentiate manifold with boundary, and f(x) is a 0-form on it. Since df = f'dx and d[a,b] = {6} - {a}, by the Stokes theorem / f'(x)dx = f J[a,b) Jd\a f{x) = f(b) - /(a). ¦*1 3.3 If we recall the definition of the integral of n-forms on oriented n-dimensional C°° manifolds, A) and B) are easy. 3.4 Since ujAt] does not vanish everywhere on M, M is orientable. We specify an orientation. Then fMu> A 77 ^ 0. Now if [u>] = 0, there exists an element 6 e Ak~l(M) such that dd = u. On the other hand, since u; A 77 = d(8 A 77), by the Stokes theorem, / u;A77= / Jm Jm d{6 A r?) = 0. This is a contradiction.
SOLUTIONS TO EXERCISES 3.5 It is easy to see that, in general, if a C°° manifold M is connected, HpR(M) = R. Next, while an arbitrary 1-form f(x)dx on R is a closed form, if we put f{x) = r Jo f(x)dx we have dF = fdx, so HlDR{R) = 0. It is clear that H^,R(R) = 0 for fc>l. 3.6 We shall prove only that H^^S1) = R. If we define a map / : Hl,R(Sl) -* R by I(u>) — /sl u>, it is easy to see that this is a surjection. This is because an arbitrary 1-form u> on Sl can be expressed asw = fdx by a function / : R —¦ R such that f(x 4- 1) = f(x) for an arbitrary x G R, and then we have /([o»]) = / f(x)dx. Jo Now let us assume that I{[u>]) = 0; that is, / f(x)dx = 0. Then, if ./o we put F(x)= f*f(x)dxt Jo F(x) becomes a function on Sl, because clearly F(x 4- 1) = F(x). Also, since it is obvious that dF = fdx = u>, we have [u>] = 0, and the claim is proved. 3.7 By using the polar coordinates, we see that R2 - {0} is dif- feomorphic to Sl x R. Therefore, by the homotopy invariance of the de Rham cohomology, we have HbR(R2-{0}) = HhR(Sl), and Hp^S1) is determined by the previous problem. Also, for example, jr-^(xdy-ydx) is a closed 1-form on R- {0}, and we see that its de Rham cohomology class is not 0 (see Exercise 2.9 of Chapter 2). 3.8 If we denote the unit disk in R3 by D3, then 3D3 = S2. Therefore, by the Stokes theorem, / <jj = / du> = 3 / dx\ A dx2 A dx$ = 4tt. Js2 Jd3 Jd* 4 Here we used the fact that the volume of D3 is -it.
SOLUTIONS TO EXERCISES m 3.9 If d = 0, we can take a constant map. Assume that d ^ 0 Then we take \d\ distinct points pt (i = 1, • • • , |d|) on M, and let U{ be mutually disjoint small coordinate neighborhoods around them. Then we can construct a C°° map / : M -* Sn such that it maps each point Pi to the north pole of Sn and each ?/* to the northern hemisphere preserving or reversing orientation according to whether d is positive or negative respectively, and, furthermore, all the remaining part of M to the southern hemisphere. Practically, we can use an appropriate finite open covering of M and a partition of unity subordinate to it. Now let a; be an n-form on Sn such that supp u> is in the northern hemisphere and furthermore /s„ u> = 1. Then obviously JM fu> = d. Therefore by Proposition 3.29, the mapping degree of / is exactly d. 3.10 Let 7r : M —> M/G be the natural projection. If w is a closed fc-form on M/G, then tx'uj is a closed fc-form on M invariant under G. The correspondence u> ¦-» ir*u) induces a natural linear map B) tt* : H*Dr{M/G) - H*DR(Mf. It is enough to show that this map is in fact a bisection. Since it obviously is for k = 0, we assume that k > 0. First we shall see that it is an injection. Assume that 7r*([u>]) = 0. Then there exists an element 77 € Ak~l{M) such that n*u; = drj. Now we put '¦AS'* Here |G| stands for the order of the group G. Then we have drf = tt*u>. On the other hand, since it is easy to see that 7/ is invariant under the action of G, there exists an element f) e Ak~1(M/G) such that 7r*77 = rf. Therefore n*(dfj - a;) = 0. Now, since n* : A*{M/G) —» A*{M) is obviously an injection, we have u> = dfj, and so [u>] = 0. Next we shall see that the map B) is a surjection. Assume that a de Rham cohomology class [a;] represented by a closed fc-form u on M is invariant under the action of G. Then if we put this is also a closed form and we see that [u/] = [u>]. On the other hand, since u/ is invariant under the action of G, a similar argument as above implies that there exists an element u> € Ak{M/G) such that 7r*u> = u)'. Since <I> is obviously a closed form, we can write n*([Q}) = [u/] = [u;], and the proof is finished.
308 SOLUTIONS TO EXERCISES Chapter 4. 4.1 Let V be a vector space and fx{ : V x V —> R (i = 1,2) two positive-definite inner products. It is simple to see that, for any t € @,1], A - ?)pq 4- tfi) is a positive inner product. Now the proof should be easy. We can similarly show that the set of all Riemannian metrics is contractible. 4.2 The inverse of the correspondence given in the problem is .1+w „ \—w dz dw 3t - ^ \d~\ "A-ti;)" '^l~ \dz\ 2\dw\ y ~ l-H2 *-??*• 2|dn;l ll-wp- In this case, we have From these we get 4.3 We have On other hand, in the isomorphism T*Rn = TzRn induced by the Eucliean metric it is clear that dxi corresponds to -^. 4.4 If the local representation of g is hij relative to another positive local coordinate system (V; y\,..., yn), then we have detfoy) = [det(^)]2det(^), dxi dx\ A • • • A dxn = det(-5—-)dy\ A • • • A dyn. oyj Prom these two equations we have Jdet{gij)dxi A • • • A dxn = Jdet{hij)dyi A • • • A dyn. If for a point p we choose yt such that det(/ii;) = 1, then we get % = dj/i A • • • A dyn at p, proving the first half. It also follows that the volume element of H2 can be written in the form dx^ffi. 4.5 By definition, we have divXi>M = (*d * ujx)vm = *2d * ujx = d*uix-
SOLUTIONS TO EXERCISES 309 Hence by the Stokes theorem we have / divXVA, = / *ojx. JM JdM Now at an arbitrary point p e dM we choose a positive orthonormal basis ei,..., en in TPM such that e\ = n, and we let 9\,..., 0n be the dual basis. In this case, (X,n)vdM = (X} ei)#2 A • • • A 6n. On the other hand, X = Yli(x>ei)ei implies ux = ^(X.e,)^. Therefore if i : dM cMis the inclusion map, we get i*(*u>x) = {X, eiH2A- • -A0n. It now follows that / *"x = / {X,n)vdM, JdM JdM completing the proof. 4.6 By definition, we get A/ = {d6 + $d)f = 6df = -*d*df = -divgrad/. 4.7 Consider, for example, Massey products arising from the relation xy = yz of three cohomology classes. If a, 0,7 are harmonic forms representing x, y, 2, then by assumption a A 0 and 0 A 7 are both harmonic forms. By the Hodge theorem, they coincide with each other and the corresponding Massey product is 0. The proof in the general case is similar. 4.8 Denote by m,n, *m,*n the dimensions of M,N and the Hodge operators for MtN. For uj e Al(M),rj € Aj(N) we get *{tt*1uj A 7^77) = (-l)(m_l);7rf (*Mu;) A ir^NV)- If 6uj = 6t) = 0, it follows that 6(n[u A 7r*77) = 0. If u>, 77 are both harmonic forms, then, of course, we get d{it{uj A 7^77) = 0. Hence in any case, we have AGrju/ A tt^) = 0. The second half follows by applying the Hodge theorem to what we just proved. 4.9 Get two copies of M and let DM be the manifold obtained by pasting them along the common boundary dM. Then DM is an odd-dimensional closed manifold, and \(DM) = 0 by Theorem 4.21. On the other hand, think of the naturally induced triangulation of DM induced by that of M. Then we get x{DM) = 2\{M) - x(dM), and hence x(M) = ^x(dM). 4.10 The matrices representing the two intersection forms are A) and {(? !)}¦
310 SOLUTIONS TO EXERCISES Chapter 5. 5.1 For a trivialization <p : E\u = U x Rn over an open subset U of M, let f*E\f-HU) ^ (p,u) - (?,</>(")) € /_1(^) x Rn- This correspondence is a trivialization of f*'E over ef~l{U). 5.2 Let n and r (n > r) be the dimensions of E and F. Choose an open subset U so that E\y and F|t/ are both trivial. Then we can take a frame Si,..., sn of E over [/ such that its subframe Si,..., sr forms a frame of F\u- Now the n - r sections U 9 p »-+ [sj(p)] G Ep/Fp (i = r -f 1,..., n) form a trivialization of E/F over [/. 5.3 It consists of two line bundles, one trivial line bundle and one non-trivial line bundle. The second is obtained by pasting the two ends of [0,1] x R so as to identify x *-+ -x. 5.4 The normal bundle of Sn in Rn+1 is obviously a trivial line bundle. Hence we can combine the facts that TSn @e = rRn+1|5n and TRn+1 are trivial. 5.5 It suffices to verify directly that Yli ^*Vj satisfies the two conditions for connections. 5.6 For u) e A1(M) and s e T{E) we have D{u <g> s) = du> <8> s - u <g> Vs. Hence for X,Y € ?(M) we have D{u ® s){X, Y) = )-{Xu}{Y) - Yu{X) - u([X, Y))}s -±{u,(X)VYs-u,(Y)Vxs} =\{Vx(u>(Y)s) - VyMX)s) - »{IX,Y])8}. Since Vs can be written as a linear combination of elements in the form above, we have D(Vs)(X, Y) = i{Vx(Vy5) - Vy(VXs) - V(x,y]s}. Therefore we get D(Vs)(X,Y) = R(X,Y)s, completing the proof. 5.7 If we set h(t) = A + txi){l + tx2) • ¦ • A + txn), then we get h(t) = 1 + to\ + t2a2 + • • • + tnan. On the other hand, from ft{\ogh{t)) = jfch'{t), we obtain h{t)±{\ogh{t)) = h'{t). Formally we have jt(\ogh(t)) = x,(l - ix, + t2x\ -...) + ... + xn(l - txn + t2xl ).
SOLUTIONS TO EXERCISES 3 Finally we end up with — (log/i@) = si - ts2 + t2s3 and A + tax + t2a2 + ¦ ¦ • + tnan){si - ts2 + t2s3 ) = Gi + 2ta2 + ¦¦• + ntn~1(Trx. By comparing the coefficients of t{ we get Newton's formula. 5.8 Use a known property of Pfaff polynomials, namely, the fact that for two alternating matrices X and Y pf{o y) = pfmpf(yy 5.9 For s 6 T(E') and t e T{E), there is a function (s,t) on M. We define a connection V* such that X(s,t) = (V*xstt) + {syVxt). If si,..., sn is a loal frame field on E and 01,..., 0n the corresponding dual frame. If we write Vsj = J2iuj ® 5»> tnen tne condition above implies V*#* = ]T^ —u>*-0J'. This shows the meaning of the problem. 5.10 A) p2 = 9,p? = 18; B) c3 = 6,C!C2 = 24,c? = 54. Chapter 6. 6.1 Over an open subset Ui = {[z\,... ,zn'?l]\Zi ^ 0} a triv- ialization can be given by the correspondence /i_1(C/i) 3 {zj) *-* ([zj),Zi/\zi\) eUiX Sl. The action of Sl on the total space 52n+1 is given by (z,) -» (Zjz) (z € S1). 6.2 Let U C ? be a coordinate neighborhood and let y>: n 1(U) = G x F be a trivialization over U. In this case, the map that takes (p,u) € (/-ir)-1^1^)) to ?(p,ti) = (/(p),v>(«)) € 17 x F is a bijection. By postulating such <p to be diffeomorphisms, we can define a C°° structure on f*E. Furthermore, (p gives a trivialization of f*E over f'l{U). 6.3 By definition of a fiber bundle, the projection n : E —> B is clearly a submersion. By Exercise 2.6 in Chapter 2, -n* : A*(B) —> «4*(?) is an injection. 6.4 Pick a 1-cochain d e C^/f.Z) such that c - cs = 6d. Next take a section s' with s' = s at each vertex and satisfying the following condition. For any 1-simplex ac, there is a map from the oriented path
312 SOLUTIONS TO EXERCISES s'(k) -s(k)~1 into 51, by the same argument as in the proof of Lemma 6.18. Now we determine s'(k) so that its degree coincides with s'(k). In this case, we get cs> = cs + 5d = c. 6.5 For an oriented 2-dimensional vector bundle 7r : E —* M, introduce a Riemannian metric and choose a connection V compatible with the metric. If u — (o>j) and ft = (ft*) are the connection and curvature forms, we can write --(.J "»')¦ -U ?)¦ Moreover, we have u\ = -wj, ft? = -ftj, ftj = dwj. Now if P(E) denotes the principal GLB; R) bundle associated to E, then V determines a connection u) by virtue of Theoorem 6.50. On the other hand, if we set S(E) = {u G E; ||u|| = 1}, then the projection 5(E) —> E is an oriented 51 bundle. We can also consider S(E) as a submanifold of P{E). That is, for u e S(E) let u' be the vector obtained by -n/2 rotation of u in the positive direction. We associate the frame [u,u'\ to u. In this case, the B,1) component of the 1- form that is the restriction of u> to 5(E), namely, the portion u)\, is a connection form of the principal 51 bundle S(E), because exp((? "M-f"' """A \1 0 / \sin< cost J Now by definition the Euler class of E as an oriented 2-dimensional vector bundle is represented by the closed 2-form j^ty- On the other hand, the Euler class of the 51 bundle S(E) is represented by -^rftf- This concludes the proof. 6.6 On 53 = {z = {zi,z2)\\z\\2 + \z2\2 = 1}, consider a 1-form u) = Xidyi — yidx\ + x2dy2 - y2dx2. For each point z = [z\, z2) € 53, define a map /, : 51 -» 53 by /z(e*) = (zicw,z2eifl). A direct computation shows that /*a> = dO. Also, w is invariant by the action of 51 on 53 (cf. Exercise 6.1). It follows that uj is a connection form on the 51 bundle h : 53 -» CP1. Since cL> = 2(<fai Ady! +ctr2 Ady2), the Euler class can be determined by computing — ^ /cpl do;. Now we
SOLUTIONS TO EXERCISES 313 identify U = {[reie, 1]} C CP1 with C and define a section s : C — S3 by s{rei0) = (-7=L=reie, , * J. VT+75 n/TTT2^ Then 5* (da;) = i^^drdd. Now we can finish the proof by using: 2irJCP, 27rJc(l + r2J 6.7 A*(so)B)* is an exterior algebra E{6) . Hence W(so)B) as E@)®R[0]. 6.8 First, that g~ldg is a left invariant 1-form with values in gl(n;R) follows from (9o9)~'ld(gog) = g^g^godg = g~ldg for any g0 e GL(n;R). Second, for any left invariant vector field A = (a*) € fll(n;R) the value g~ldg(A) at e is equal to (g~ldg)A = A, since {g~ldg)e = {dg)) and A = Eaj^r. 6.9 Let Bi, B2, B3 be an arbitrary basis in g = suB) and 0i, 02,03 the dual basis. Set 0 = ]T 9{ ® Bi € W(g) <8> g and 1 2 Tc2 = s-^Tr@ A E0 + -0 A 0 A 0). Verify that STC2 = c2 by direct computation. 6.10 For example, the pull-back /j.*(d9) of dO by the map fj,: S1 x S1 —» S1 that defines multiplication is a connection form satisfying the condition in the problem.
References Some of the books and research articles listed here are original, historical, or highly recommended. Others are quoted in the text as useful or convenient references on the subjects. [A] Allendoerfer, C.B., The Euler number of a Riemannian manifold, Amer. J. Math. 62 A940), 243-248 [AW] Allendoerfer, C. B. and Weil, A., The Gauss-Bonnet theorem for Riemannian polyhedra, Trans. Amer. Math. Soci. 53A943), 101-129 [BT] Bott, R. and Tu, W., Differential Forms in Algebraic Topology, Springer, 1982 (C) Chern, S.S., A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemannian manifolds, Ann.of Math. 45A944), 747-752 [Ca] Cartan, H., Notion d'algebre differentielle; application aux groupes de Lie et aux varietis ou opere un groupe de Lie, 15- 27; La transgression dans un groupe de Lie et dans un espace fibre principal, 57-71; Colloque de Topologie, Bruxelles, Mas- son, Paris, 1951 (deR] de Rham, G., Varietes Differentiables, Hermann, 1955 [DFN] B.A. Dubrovin, B.A., Fomenko, A.T., and Novikov, S.P., Modern Geometry-Methods and Applications, Part I, The geometry of surfaces, transformation groups, and fields, Part II. The geometry and topology of manifolds, Part III. Introduction to homology theory, Springer, 1984, 1985, 1990 [E] Ehresmann, C, Les connexions infinitesimales dans un espace fibri differentiable, Colloque de Topologie, Bruxelles, Masson, Paris, 1951 A950), 29-55 [Fj Fenchel, W., On total curvatures of Riemannian manifolds I,. London Math. Soc. 62 A940), 243-248 315
316 REFERENCES [Fl] Flanders, H., Differential Forms with Applications to the Physical Sciences, Academic Press, 1963 [GP] Guillemin,V. and Polack, A., Differential Topology, Prentice Halll, 1974 [H] Helgason, S.,Differential Geometry, Lie Groups and Symmetric Spaces, Academic Press, 1978 [KN] Kobayashi, S. and Nomizu, K., Foundations of Differential Geometry, I, II (Interscience), John Wiley, 1963, 1969 [M] Milnor, J.,On the cobordism ring Cl* and a complex analogue, Amer.J. Math. 82A960), 505-521 [MS] Milnor, J.W. and Stasheff, J.D., Characteristic Classes, Princeton University Press, 1976 [Mu] Munkres, J. R. Elementary Differential Topology, Revised Edition, Princeton University Press, 1966 [NS] Nash, C. and Sen, S., Topology and Geometry for Physicists, Academic Press, 1983 [S] Steenrod, N., The Topology of Fiber Bundles, Princeton University Press, 1951 [T] Thorn, R., Quelques proprietes globales des varietes differentiables, Comm.Math.Helv. 28A954), 17-86 (W] Whitney, H.,Geometric Integration Theory, Princeton University Press, 1951 [Wal] Wall, C.T.C., Determination of the cobordism ring, Ann. of Math. 72 A960), 292-311 (War] Warner, F.W., Foundations of Differentiate Manifolds and Lie Groups, Springer, 1983 [We] Wells, R.O. Jr., Differential Geometry on Complex Manifolds, Prentice Hall, 1973; Springer, 1979
Index C , 21 CPn , 22 Cr function , 5 Cr map , 5 C°° atlas , 15 C°° diffeomorphism , 24 C°° differentiate homeomorphism , 5 C°° differentiable manifold , 15 C°° function , 5 , 23 C°° manifold , 15 C°° map , 5 , 24 C°° singular k chain , 103 C°° singular ^-simplex , 103 C°° singular cochain complex , 104 C°° structure , 15 C°° triangulation , 101 C°° vector field , 9 Diff M , 43 ExptX , 43 G-structure , 234 GL(n;C) , 22 GL(n; R) , 20 Hn , 44 k cochain , 120 fc-form , 58 /-chain , 97 /-simplex , 96 n-dimensional numerical space , 2 n-dimensional sphere , 17 n-dimensional torus , 17 n-dimensional vector space , 6 n-sphere , 17 0{n) , 22 Pn , 21 R , R2 , R3 , 2 Rn , 2 , 6 RPn , 21 SO{n) , 23 TxRn , 6 c-neighborhood , 3 abstract simplicial complex , 97 action of a group , 50 adjoint operator , 154 admissible , 234 Alexander-Whitney map , 133 algebra , 24 , 57 alternating , 63 alternating form , 63 anti-derivation , 73 associated bundle , 236 atlas , 14 automorphism group , 50 base space , 171 basic element (in a Weil algebra) , 276 Betti number , 116 Bianchi's identity , 196 boundary , 45 boundary cycle , 98 boundary operator , 98 bracket , 39 bundle map , 232 , 235 Cartan formula , 74 Cartan-Eilenberg theorem , 138 Cech cohomology , 119 cell , 96 chain complex , 98 characteristic class , 198 , 238 characteristic number , 226
318 INDEX Chem class , 206 Chern number , 225 Chern-Simons form , 287 classes Cr , C°°t 5 classifying space , 239 closed form , 60 , 111 closed manifold , 46 cobordant , 226 coboundary , 99 cochain complex , 98 cocycle , 99 cocycle condition , 171 , 233 coherent orientation , 46 cohomologous , 99 cohomology , 98 commutative vector fields , 82 compact , 27 compatible (with a metric) , 199 complement of a knot , 20 complete , 43 completely integrable , 80 complex Lie group , 22 complex manifold , 21 complex projective space , 22 complex vector bundle , 171 complexification , 175 conjugacy , 290 conjugate bundle , 209 connection , 185 for a complex vector bundle , 205 in a general bundle , 258 in a principal bundle , 260 connection form , 185 , 264 contractible , 119 contractible open covering , 121 coordinate change , 15 coordinate functions , 12 coordinate neighborhood , 12 cotangent bundle , 67 , 177 cotangent space , 67 covariant derivative , 181 , 185 covariant exterior differential , 193 covering , 27 covering manifold , 51 covering map , 51 curvature form , 186 , 188 , 252 , 264 cycle , 98 de Rham cohomology , 111 algebra , 113 group , 112 de Rham complex , 112 de Rham theorem , 114 concerning the product , 131 for triangulated manifolds , 115 derivation , 38 diffeomorphism , 5 , 24 diffeomorphism group , 43 differentiate manifolds , 1 differential , 33 differential form , 58 coordinate-independent definition , 63 differential ideal , 87 directional derivative , 8 discrete group , 50 distance , 3 distribution , 80 , 258 divergence , 152 double complex , 123 dual bundle , 176 dual space , 63 elliptic PDE , 161 embedding , 34 Euclidean simplicial complex , 96 Euclidean space , 147 Euler characteristic , 164 Euler class , 212 , 246 , 254 Euler form , 213 Euler number , 164 Euler-Poincare characteristic , 164 exact form , 60 , 111 existence and uniqueness of the solution of ODEs , 41 existence of partitions of unity , 29 exterior algebra , 58 , 63 exterior differentiation , 59 , 70 exterior power bundle , 177 exterior product , 59 , 69 face , 96 fiber , 171 , 232 fiber bundle , 232 fiat connection , 288 flat G bundle , 288 frame field , 172
INDEX 319 free , 50 Frobenius theorem , 81 , 88 fundamental class , 103 fundamental vector field 119 Gauss-Bonnet theorem , 216 Gaussian plane , 22 general linear group , 20 general position , 96 geodesic , 180 gradient , 148 graph , 55 Grassmann algebra , 63 Green's operator , 161 harmonic form , 155 harmonic function , 155 Hausdorff separation axiom , 11 Hausdorff space , 11 Hirzebruch signature theorem , 227 Hodge decomposition , 160 Hodge operator * , 150 Hodge theorem , 159 holomorphic mapping , 22 holonomy homomorphism , 289 homeomorphism , 4 homogeneous coordinate , 21 homologous , 98 homology group , 97 homology theory of cell complexes , 96 of simplicial complexes , 96 homotopy invariance of de Rham cohomology , 119 homotopy type , 119 Hopf invariant , 134 Hopf index theorem , 256 Hopf line bundle , 174 Hopf map , 25 , 34 horizontal lift , 259 horizontal vector , 259 hyperbolic space , 147 immersion , 34 index , 256 induced bundle , 236 induced connection , 198 integrability condition , 88 integral curve , 39 integral manifold , 80 interior product , 73 intersection form , 166 intersection number , 165 invariant polynomial function , 194 inverse function theorem , 5 , 6 involutive , 81 isomorphic bundles , 172 , 235 isolated singular point , 255 Jacobi identity , 39 Jacobian , 5 Jacobian matrix , 5 knot , 20 Kronecker product , 99 Kiinneth formula , 168 Laplace-Bertrami operator , 155 Laplacian , 155 left-hand system , 47 lens space , 53 Levi-Civita connection , 201 Lie algebra , 39 , 90 Lie derivative , 77 Lie group , 22 lift , 259 line bundle , 171 link with two components , 140 linking number , 141 local chart , 12 local coordinate system , 1 , 12 positive , 49 locally finite , 27 manifold with boundary , 45 mapping degree , 139 Massey products , 136 triple , 136 Maurer-Cartan equation , 92 Maurer-Cartan form , 91 maximal atlas , 15 maximal integral curve , 41 metric connection , 199 metric space , 3 multilinear , 63 nerve , 119 Newton's formula , 195 nonzero section , 172
320 INDEX normal bundle , 174 null cobordant , 226 one parameter group of local transformations , 42 one parameter group of transformations , 43 open covering , 27 open neighborhood , 3 open set , 3 open simplex , 119 open star , 119 open submanifold , 20 orbit , 50 orbit space , 50 ordered basis , 47 orientable , 46 , 48 , 211 orientation , 47 , 97 orientation preserving , 49 oriented manifold , 48 orthogonal group , 22 paracompact , 27 parallel along a curve , 183 parallel displacement , 183 partition of unity , 29 Pfaffian , 212 Poincare disk , 167 Poincare' duality , 163 Poincare' lemma , 118 polar coordinates , 16 polyhedron , 97 Pontrjagin class , 200 Pontrjagin form , 201 Pontrjagin number , 225 primary obstruction , 255 principal bundle , 236 principal G-bundle , 236 product bundle , 171 product manifold , 16 projection , 171 proof of the de Rham theorem , 126 properly discontinuous , 50 pullback , 72 , 236 quotient bundle , 174 quotient space , 50 real projective space , 21 reducible , 235 refinement , 27 regular submanifold , 21 restriction of a bundle , 173 Riemannian manifold , 146 Riemannian metric , 146 in a vector bundle , 175 right-hand system , 47 second countability axiom , 12 section , 172 , 233 self-adjoint , 155 signature , 166 simplicial complex , 97 singular fc-chain , 100 singular fc-simplex , 100 singular chain complex , 100 singular homology group , 100 singular homology theory , 96 singular point of the vector field , 42 special orthogonal group , 23 stabilizer , 50 standard /c-simplex , 99 Stiefel-Whitney class , 227 Stokes theorem , 107 on chains , 109 structure constant , 91 structure equation , 188 , 265 structure group , 234 subbundle , 174 submanifold , 20 submersion , 34 support , 29 , 106 symbol , 162 symplectic form , 93 system of Pfaffian equations , 88 tangent bundle , 170 tangent frame bundle , 240 tangent space , 6 , 30 tangent vectors , 7 , 30 topological manifold , 13 topological space , 3 topologically invariant , 100 torsion tensor , 203 total Chern class , 206 total differential , 59 total Pontrjagin class , 201 total space , 171 , 232
INDEX 321 transition function , 171 triangle inequality , 3 triangulation , 97 trivial bundle , 232 trivial connection , 185 trivialization , 171 unit sphere bundle , 256 unitary group , 23 universal covering manifold , 51 universal G-bundle , 239 upper half space , 44 vector bundle , 171 vector field , 9 , 37 vector space , 6 velocity vector , 8 vertical vector , 259 volume element , 151 volume form , 139 , 151 Weil algebra , 268 Weil homomorphism , 275 , 278 Whitney formula , 207 -8 Whitney sum , 176 Whitney's embedding theorem , 10 , 36 zero section , 172