Translations of
MATHEMATICAL
MONOGRAPHS Volume 205
Variational Problems in Geometry Se'' Nishikawa
k
American Mathematical society
Variational Problems in Geometry
Translations of
MATHEMATICAL MONOGRAPHS Volume 205
Variational Problems in Geometry Seiki Nishikawa Translated by Kinetsu Abe
Amerl=n Mathematical 8ocisty Providence. Rhode Island
Editorial Board Shoshichi Kobayashi (Chair) Masamichi Takesaki
6 fol 0: FL16 M** 9 3 KIKAGAKUTEKI HENBUN MONDAI by Seiki Nishikawa Originally published in Japanese by Iwanami Shoten, Publishers, Tokyo, 1998 T anslated from the Japanese by Kinetsu Abe
2000 Mathematics Subject Ciasaication. Primary 53-01, 53C21, 53C43, 58E20, 58J25.
Library of Congress Cataloging-In-Publication Data Nishikawa, Seiki. [Kikigakuteki henbun mondai. English]
Variational problems in geometry / Seild Nishikawa ; translated by Kinetsu Abe
p. cm. - (Translations of mathematical monographs, ISSN 0065-9282 ; v. 205) (Iwanami series in modem mathematics) Includes bibliographical references and index. ISBN 0-8218-1356-0 (acid-free paper) 1. Harmonic maps. 2. Variational inequalities (Mathematics). 3. Riemannian manifolds. I. Title. It. Series. III. Series: Iwanami series in modern mathematics QA614.73.N5713 2001 514'.74-dc21 2001046350
© 2002 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.
® 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://vvv.ams.org/
10987654321
070605040302
Contents Preface to the English Edition
ix
Preface
Outlines and Objectives of the Theory Chapter 1. Are Length of Curves and Geodesics 1.1. Are length and energy of curves 1.2. Euler's equation 1.3. Connections and covariant differentiation
1
1
9 16
Geodesics Minimal length property of geodesics Summary Exercises
28 38 43 44
Chapter 2. First and Second Variation Formulas 2.1. The first variation formula
47 47 54 65 69 77
1.4. 1.5.
2.2.
Curvature tensor
The second variation formula Existence of minimal geodesics Applications to Riemannian geometry Summary Exercises 2.3. 2.4. 2.5.
Chapter 3. Energy of Maps and Harmonic Maps 3.1. Energy of maps 3.2. 3.3. 3.4. 3.5.
Tension fields
The first variation formula Harmonic maps The second variation formula Summary Exercise
82 82 85
85 90 99 103
110 114 114
v
CONTENTS
vi
Chapter 4. Existence of Harmonic Maps 4.1. The heat flow method 4.2. Existence of local time-dependent solutions 4.3. Existence of global time-dependent solutions Existence and uniqueness of harmonic maps 4.4. 4.5. Applications to R.iemannian geometry Summary Exercises Fundamentals of the Theory of Manifolds and Functional Analysis A. I. Fundamentals of manifolds A_2. Fundamentals of functional analysis
Appendix A.
Prospects for Contemporary Mathematics Solutions to Exercise Problems Bibliography Index
Preface to the English Edition This book, published originally in Japanese, is an outgrowth of lectures given at Tohoku University and at the 1995 Summer Graduate Program of the Institute for Mathematics and Its Applications, University of Minnesota. In these lectures, through a discussion on variational problems of the length and energy of curves and the energy of maps, I intended to guide the audience to the threshold of the field of geometric variational problems, that is, the study of nonlinear problems arising in geometry and topology from the point of view of global analysis. It is my pleasure and privilege to express my deepest gratitude to Professor Kinetsu Abe who generously devoted considerable time and
effort to the translation. I would also like to take this opportunity to express my deep appreciation to Professor Phillipe Tondeur who invited me to join the 1995 Summer Graduate Program, and to my friend Andrej Treibergs for making his notes [261 available to the organization of the last chapter. Seiki Nishikawa April 2001
vH
Preface It is said that techniques for surveying were developed from the need to restore lands after frequent floods of the Nile River in ancient Egypt. Geometry is the area of mathematics whose name originates from this method of surveying; namely, "to measure lands" (geo = lands, metry = measure). As such, it is an ancient practice to study figures from the view of practical applications. It is also said that the ancient Greeks already knew of the method of indirect surveying using the congruence conditions of triangles. A minimal length curve joining two points in a surface is called a geodesic. One may trace the origin of the problem of finding geodesics back to the birth of calculus. Many contemporary mathematical problems, as in the case of geodesics, may be formulated as variationl problems in surfaces or in the more generalized form of manifolds. One may characterize the geometric variational problems as a field of mathematics that studies the global aspects of variational problems relevant in the geometry and toplogy of manifolds. For example, the problem of finding a surface of minimal area spanning a given frame of wire originally appeared as a mathematical model for soap films. It has also been actively investigated as a geometric variational problem. With recent developments in computer graphics, totally new aspects of the study on the subject have begun to emerge. This book is intended to be an introduction to some of the fundamental questions and results in geometric variational problems, studying the variational problems on the length of curves and the energy of maps. The first two chapters approach variational problems of length and energy of curves in Riemannian manifolds with an in-depth discussion of the existence and properties of geodesics viewed as the solution to variational problems. In addition, a special emphasis is ix
PREFACE
x
placed on the fact that the concepts of connection and covariant differentiation are naturally induced from the first variation formula of this variational problem, and that the notion of curvature is obtained from the second variational formula. The last two chapters treat the variational problem on the energy of maps between two Riemannian manifolds and its solutions, namely harmonic maps. The concept of harmonic maps includes geodesics and minimal submanifolds as examples. Its existence and properties have successfully been applied to various problems in geometry and topology. This book takes up the existence theorem of Eells-Sampson, which is considered to be the most fundamental among existence theorems for harmonic maps. The proof uses the inverse function theorem for Banach spaces. It is presented to be as self-contained as possible for easy reading. Each chapter of this book may be read independently with minimal preparation for covariant differentiation and curvature on manifolds. The first two chapters, through the discussion of connections and covariant differentiation, are designed to provide the reader with a basic knowledge of Riemannian manifolds. As prerequisites for reading this book, the author assumes a few elementary facts in the theory of manifolds and functional analysis. They are included in the form of appendices at the end of the book. Details in functional analysis may
be skipped. The reader, however, is encouraged to do the exercise problems at the end of each chapter by himself or herself first. The solutions may be consulted if necessary, since many of the exercise problems complement the contents of the book. This book is an outgrowth of lectures delivered at Tohoku Univer-
sity and the 1995 Summer Graduate Programs held at The Institute for Mathematics and Its Applications, University of Minnesota. The first half of the book aims at a junior and senior level, and the last half at a first and second year graduate level. Each half roughly consists of the amount of topics that may be covered in one semester. In the actual lectures, the author also discusses the harmonic maps between Riemann surfaces. This portion is not included in this book due to the limited space. The reader who is interested in the study of harmonic maps is advised to first study the harmonic maps between Riemann surfaces.
It would be this author's wish as well as pleasure if this book could interest many readers in variational problems in geometry.
PREFACE
xi
Last but not least, the author expresses his sincere gratitude to the editorial staff of Iwanarni Shoten for their valuable help in the publication of this book. Seiki Nishikawa
December 1997
Outlines and Objectives of the Theory Among geometric variational problems, the extreme value problem regarding the length of curves is as old as those in calculus. Chapter 1 of this book is devoted to discussions of variational problems of curves in manifolds. As is well known, the length of a curve joining two points in a plane is given by integrating the magnitude of tangent vectors. Similarly, one can define the length and energy for curves in more general Riemannian manifolds by measuring the magnitude of the tangent vectors using Riemannian metrics. In Chapter 1, Euler's equation is calculated. It characterizes the critical points of the length and energy of curves when they are considered as functionals defined in the space of curves. Consequently, the equation of geodesics is obtained. The concepts of connections and covariant differentiation are naturally induced from the equation of geodesics in a manifold. Covariant differentiation, an essential tool for studying variational problems in manifolds, is an operation that defines the derivative of a vector field by a vector field in a manifold. The most fundamental connection, called the Levi-Civita connection, is uniquely determined in a manifold equipped with a Riemannian metric, i.e., a Riemannian manifold. The notion of parallel transport is induced from this connection. The discovery of the notion of parallel transport in Riemannian manifolds (1917) and Einstein's use of geometry based on a four-dimensional indefinite metric for his general relativity (1916) greatly advanced the study of Riemannian geometry.
Geodesics in Riemannian manifolds correspond to straight lines in the plane and they are locally characterized as the curves of minimal length between points. One can construct a special local coordinate system, called a normal coordinate system, using these minimal geodesics about each point in a Riemannian manifold. Parallel transport and normal coordinate systems are the most basic tools in Xi;i
xiv
OUTLINES AND OBJECTIVES OF THE THEORY
comparing the geometry of a Riemannian manifold with the geometry of a model space (for example, Euclidean space).
In Chapter 2, using covariant differentiation, the first variation formula (Euler's equation) for the variational problem regarding the energy of curves in Riemannian manifolds is computed in the general case where the image of a curve is not always contained in a local coordinate neighborhood. The second variation formula is subsequently computed. Just as the notion of connections is derived from the first variation formula, it is seen that the second variation formula possesses an intimate relationship to the notion of curvature in R.iemannian manifolds. In other words, the notions of curvature tensor
and the curvature of a Riemannian manifold are naturally induced from the second variation formula for the energy of curves. Given two points in a Riemannian manifold, the distance between these two points is given by the least upper bound of the lengths of piecewise smooth curves connecting them. Whether a Riemannian manifold becomes a complete metric space with respect to this distance is an important question. It was relatively recently (1931) that Hopf-Rinow gave necessary and sufficient conditions for the question. The results by Hopf and Rinow are significant not only in making the notion of completeness succinct, but also in showing that this completeness is the condition that guarantees the existence of a minimal geodesic joining two given points. As stated above, the second variation formula for the energy of curves is closely related to the curvature of Riemannian manifolds. Using this, one can study the effects of the curvature of a Riemann ian manifold on its topological structure. Myers' theorem and Synge's theorem are discussed as typical examples of such applications. The
former states that the fundamental group of a compact and connected, Riemannian manifold of positive curvature is a finite group, and the latter states that an even-dimensional compact, connected and orientable Riemannian manifold of positive curvature is simply connected. Research on Riemannian manifolds using existence and properties of geodesics is being actively pursued. In Chapter 3, harmonic maps and the energy of maps are discussed. They generalize the variational problem of the energy of curves in Riemannian manifolds. Namely, a functional called the energy of maps is defined in the mapping space consisting of smooth maps between Riemannian manifolds, and harmonic maps given as its
OUTLINES AND OBJECTIVES OF THE THEORY
xv
critical points are investigated. The energy of maps is a natural generalization of the energy of curves. Examples of harmonic maps appear in various aspects of differential geometry. Harmonic functions, geodesics, minimal submanifolds, isometric maps, and holomorphic maps are a few typical examples. The first variation formula, which characterizes the critical points
of the energy functional, can be obtained by essentially the same approach as in the case of geodesics. However, the computations become unnecessarily complicated and only yield results of a local nature without use of the covariant differentiation that is naturally induced from the Levi-Civita connection of Riemannian manifolds. To alleviate these difficulties, it is designed in this chapter to derive, through discoveries in the process, the computational rules for the covariant differentiation that is induced from the Levi-Civita connection in tangent bundles and their tensor products over Riemannian manifolds. This route may not be the most direct one, but the author believes that it is more effective in familiarizing the reader with the definition and the rules of computations for covariant differentiation than the axiomatic approach. At first, the reader may feel uneasy, especially about the portion of the induced connections. Nonetheless, actual computations help promote understanding of the notion. The fastest way to grasp the rules of computation involving covariant differentiation is actually to engage in the computations. The computations of the first variation formula for the energy functional of maps yield a vector field called the tension field. It is given as the trace of the second fundamental form of the maps. A harmonic map is then characterized as a map whose tension field is identically 0. Chapter 4 is devoted to the existence problem of harmonic maps
between compact Riemannian manifolds. Whether or not a given map is homotopically deformable to a harmonic map is one of the most fundamental questions among geometric variational problems. It may be regarded as a generalization of the existence problem of closed geodesics. To this end, the "heat flow method" is first introduced. This is an effective technique for deforming a given map to a harmonic map. Then, using this technique, it is proved that any map from a compact Riemannian manifold M into a compact Riemannian manifold N of nonpositive curvature is free homotopically deformable to a harmonic map. This theorem was first proved by Eells-Sampson in 1964.
xvi
OUTLINES AND OBJECTIVES OF THE THEORY
The proof of this theorem using the heat flow method first requires the existence of a time-dependent solution to an initial value problem with any initial map of the parabolic equation for harmonic maps. The original proof uses successive approximations to construct a solution after converting the problem to a problem of integral equations via the fundamental solution of the heat equation. In this book, the solution is constructed through use of the inverse function theorem in Banach spaces in an effort to minimize the amount of preparation. The existence of time-dependent local solutions is always guar-
anteed, but the existence of global time-dependent solutions is not self-evident, since the parabolic equation for harmonic maps is nonlinear. In fact, proving the existence of global time-dependent solutions
entails some estimates of the growth rate of solutions in time. The curvature of the Rieinannian manifold N plays a crucial role in est.imating the influence of nonlinear terms. An estimation formula that guarantees the existence and convergence of time-dependent global solutions is obtained using the Weitzenbock formula for the heat operator under the condition that N is of nonpositive curvature. The Weizenbock formula, in general, gives the relationship between second order partial differential operators naturally acting on tensor fields on Riemannian manifolds and the Laplace or heat operator acting on functions. It is revealed that the Riemann curvature and its Ricci identity play essential roles for existence of solutions to those differential operators. In this chapter, an a priori estimate regarding the growth rate of solutions is obtained using the Weizenbock formula for the energy density of solutions to the parabolic equation for harmonic maps and the heat operator. This idea is originally due
to Bochner. It has become an effective and fundamental technique for the proofs of theorems such as the Kodaira vanishing theorem and more recently in gauge theory.
As in the case of geodesics, one can also investigate the structures of Riemannian manifolds using the existence and properties of harmonic maps. The theorem of Preissman, one of the typical applications of harmonic maps, is discussed. The theorem states that a nontrivial, Abelian subgroup of the fundamental group of a compact manifold of negative curvature is infinitely cyclic. The research of Riemannian manifolds using the existence and properties of barmonic maps seems to possess a promising future. For example, new proofs from a more analytical point of view for the topological sphere theorem and the Frankel conjecture were recently given by exploiting
OUTLINES AND OBJECTIVES OF THE THEORY
xvii
the existence theorem of harmonic spheres due to Sacks and Uhlenbeck. A strong rigidity theorem regarding complex structures in Kahler manifolds of negative curvature was also obtained using the existence theorem of Eells and Sampson.
CHAPTER 1
Arc Length of Curves and Geodesics "Given two points in a surface, find a curve joining them of the minimum arc length." A solution to this question is called a geodesic. Finding geodesics is a typical problem in the calculus of variation. Its origin could be traced back to the birth of calculus.
In this chapter, the variational problem of arc length and the energy of curves in Riemannian manifolds is discussed as an introduction to geometric variational problems. The critical points in this variational problem satisfy a differential equation called the Euler
equation. The concept of covariant differentiation is naturally induced from this equation. The first variational formula of the energy of curves is obtained. Geodesics are characterized as the critical points of this variational problem.
1.1. Are length and energy of curves The reader, who has already learned the theory of surfaces, knows how the arc length of a curve drawn on a surface is defined. We begin by reviewing definitions. Let M be a surface in the Euclidean space E3 of dimension threeLet c be a smooth curve in M. Using the coordinates in E3, we denote the parametric representation of the curve c = c(t) by c(t) = (x(t), pi(t), z(t)),
a < t < b.
The tangent vector c'(t) of the curve c at the point c(t) is given by
c(t) = (x'(t), y'(t), z'(t)), a < t < b. If we consider the curve c as the trace of a moving particle c(t), the
tangent vector c/(t) is nothing but the velocity vector of the particle at time t. Since c(t) is a vector in E3, its magnitude Ic'(t)I may be measured in terms of the Euclidean inner product (, ). In fact, the
i
2
1. ARC LENGTH OF CURVES AND GEODESICS
magnitude lc'(t)I is given by XI(t)2
(c'(t) _ (C(t),c(t))1'2 =
+ y'(t)2 + x'(t)2,
and it represents the speed of the particle at time t. Then length L(c) of the curve c is given by integrating the magnitude of the tangent vector c(t) in t as b
L (c) =
j
Ic'(t) Idt.
The arc length L(c) is the distance traveled by the moving particle along the curve c(t) from time a to time b. Similarly, we can define the are length of a smooth curve in a more general manifold by measuring the magnitude of the tangent vectors
and by integrating it along the curve. In the case of a curve in a surface, the magnitude of the tangent vectors is measured by using the inner product of the Euclidean space E3 that contains the surface. In a general c°O manifold, a tensor field called a Riemannian metric plays the role of the inner product in E. This entails the introduction of Riemannian metrics. Let M be a c°° manifold of dimension m. Given a point x in M, denote by (U, 0) a local coordinate neighborhood system around x. Here, U is an open subset of M and 0 is a homeomorphism from U onto an open subset i(U) in the mrdimensional Cartesian space Rm. Using the coordinates of Rm, O(x) can be expressed as '0(x) = (x1(x), ... , xm(x))
E Rtm.
We call (x1(x), x2 (x), ... , xm (x)) the local coordinates of x and the m-tuple (x1, x2, ... , xm) of functions the local coordinate system
with respect to (U, i). Since 0 is a homeomorphism from U onto ,O(U), we identify a point x E U and 0(x) E R' via ¢. Hereafter, we denote the local coordinates of x by
x = (x1, x2, ... , xm) = (xs),
1 < i < M.
As far as the local discussion of manifolds is concerned, identifying U and O(U) is more convenient than mapping points of U into Rm under 46 each time; consequently, it helps the reader reach the essence of the arguments more promptly. With this convention understood, for example, given a cO° function f : M - R, we may simply write its local coordinate representation f o 0-1 : 0(U) -+ R by F
1
2
m
1.1. ARC LENGTH AND ENERGY OF CURVES
3
instead of
f (x) = f o O-' (x' (x), x2(x), ... , xm(x)),
x E U.
Let TxM be the tangent space of M at x. Given a c°O function f and local coordinate functions x' near x, the directional derivative of f in the direction x' is expressed as
()(f)= Lf
(x),
15 i<m.
x
/ (
From the definition of the tangent space (see §A.1(b), Appendix),
ft6
is a tangent vector to M at x.
/z
l1
(1.1)
, ... ,
8x2
,
10
x
&m 01 )XI
forms a basis for the tangent space TxM; hence, any vector v E TxM is given uniquely as a linear combination of the basis elements in
(1.2)
V= i-1 DO (ax=
x
(C ', C2'. .. , gym) are called the components of C with respect to
the local coordinate system (x', x2,... , x"'). By applying v to each coordinate function x', we see that C' = v(x`),
1 < i < M.
On the other hand, applying the differential (dxi)x of each coordinate function x', we get
(r / lx/l ( ( axi x )( ldx)x\\a _
x)=
1<$j<m.
Therefore, we see that (1.3)
1(dz')x, (d c2)x, ... (dx'n)x)
gives rise to a basis for the dual space ,,x* of TxM. (1.3) is called the dual basis of the basis (1.1) for T.M. Now assume that there is given an inner product gx in the tangent
space TxM at each point x of the rn-dimensional c°° manifold M. Namely, gx is a bilinear form
9x:T,,MxT=M-.R
1. ARC LENGTH OF CURVES AND GEODESICS
4
on the tangent space TM, satisfying (i) (symmetric property) 9,(V, w) = gx(w, v) for any v, w E TxM and (ii) (positive definite property) g(v, v) > 0 for any v E ;.t. The inner product 9x is not arbitrarily given at each tangent space T.M. In particular, if gz is smoothly related to x E M in the following sense, gx is said to define a Riemannian metric in M. DEFnvITION 1.1. Suppose that an inner product gx is given in the tangent space TxM at each point x of a cO° manifold M. Furthermore,
we assume that given an arbitrary coordinate neighborhood U in M and its local coordinate system (2I, x2, ... , x'), a function defined by
(1.4)
9i; (x) = 9x t
(ai)
, i axi
1 < i, j < m
is c°°O in U for all i, j. Then we call the family of the inner products g = {gx}ZEM a Riemannian metric on M. When a Riemannian metric g is given to a c°0 manifold M, we call the pair (M, g) of M and g a Riemannian manifold The function gs f is called the (i, j) component of g with respect to the local coordinate system (xi, x2, ... , x"'). The m x m matrix (ge.1) is a positive definite symmetric matrix at each point x E M. In fact, the symmetry of gx implies that
()) = (()
/l
=Fur
9x being positive definite is nothing but the matrix (9ij (x)) being positive definite.
As is readily seen, the set of all bilinear forms gives rise to a vector space over R under naturally induced addition and the scalar product. We denote this vector space by TxM` 0 TxM', that is, TYM a TxM' = { f :TxM 0 T.M -+ IR J f is a bilinear map}. TxM` OTxM` is called the tensor product of TxM` and TxM". Given the dual base (1.3) for TM*, we define a bilinear form (dx' )x g (dam )x by
(dx`)x ® (dx')x(v, w) = (d r')x(v)(dXj)x('w),
Then we can easily verify that
{(dx')x0(dxj)xI I
v, w E T. M.
I.I. ARC LENGTH AND ENERGY OF CURVES
S
forms a base for TZM' 0 TZM*. Hence, we see that TZM' ® TZM' has dimension ma .
Since the inner product gx is a symmetric bilinear form, using the above base, we can express m
9i,j(x)(dx=)z 0 (dxj)z
9x =
Noting the symmetric property of gx, if we write (dx=)x (dx' )T = 2 { (dxi)x ® (dx' )x + (dxj )x ®(dxi)Z },
we can also express m
9x =
9ij(x)(dxi)x
i,j=l
(d)Z
Hence, a Riemannian metric was often traditionally denoted by in
9 > gijdx=dx'. t,?=1
EXAMPLE 1.2. Let M be a submanifold of the
Eu-
clidean space. By definition, there exists a c°° immersion cp : M E". Since cp is an immersion, the derivative dcp. of cp is an injective linear map from TM into T,(z) at each x E M. As is for a surface in IE3, using the Euclidean inner product (, ), we define 9Z(v, w) = (d(PZ(v), dws(w)),
v, w E T,,M.
Since d'px is injective, gz gives rise to an inner product in T. The family 9 = {9x )zE M gives a Riemannian metric on M. This Riemannian metric g is called the induced metric from the Euclidean space E". More generally, let cp : M --> N be a c°° immersion of a cO° man-
ifold M into another cO° manifold N, where (N, h) is a Riemannian manifold. At each x E M, we set gx(v.w) = h9,(z)(dco (v), dPo (w)),
v, w E TIM.
Then the family g = {9z}xEM defines a Riemannian metric on M. This g is called the induced metric from h through 'p, and it is denoted by g = W*h. As for a general c°° manifold M, not necessarily a submanifold, we can define the Euclidean inner product in each coordinate neighborhood by identifying it with an open subset of the
1. ARC LENGTH OF CURVES AND GEODESICS
6
Euclidean space. Using the partition of unity, we can glue them together to obtain at least one (in fact, infinitely many) Riemannian metric on M. Consequently, we can regard a c°O manifold with the second countability axiom as a Riemannian manifold. When there is no fear of confusion, we simply denote by M a Riemannian manifold (M, g).
We now consider curves in M. Let (a, b) (-oc < a < b < oo) be an open interval in real line R. We also regard R and (a, b) as a one-dimensional c°° manifold and its open submanifold, respectively. A cO° map c : (a, b) -+ M from (a, b) into M is called a c°° curve or a smooth curve in M defined over (a, b). The coordinate function t in (a, b) is called the parameter of the curve c. If c(t) = x, we say "the
curve c passes through point x in M at t." Since we have defined a curve c as a map from (a, b) into M, two such curves are considered distinct if they are distinct as maps, even if their images are identically
the same as sets in M. Let [a, b] (-oo < a < b < oo) be a closed interval, and let (a - e, b + E) ( > O) be an open interval containing [a, b]. We call a curve c : [a, b] -- M a c"O curve or a smooth curve in M defined in [a, b], if it is the restriction of a c°O curve c : (a - e, b + e) --+ M to [a, b]. Also given a continuous curve c : [a, bJ -- M, if there is a partition a = ao < a1 < . < a,. = b of the interval [a, b] such that c is a C0° curve in each interval [aa_1, a,] (1 < i < r), we call c a piecewise smooth curve in M defined over [a, b]. Let c : (a, b) -+ M be a c0O curve defined in an open interval (a, b)
and let to be a point in (a, b). Regarding (a, b) as a c°O manifold, the directional derivative with respect to the coordinate function t determines a tangent vector
l (! ET.(a,b) 1 to
to (a, b) at to. The image of this vector under the derivative map d,
: Tto (a, b) - TO(E) M is a tangent vector
(1.5)
dcen ((f)) E T4to)M
to M at c(to). We call this vector the tangent vector of c at t = to, and denote it by c' (to). Let (xl, x2, ... , x"') be a local coordinate system defined near c(to) and let c' = x` o c (1 < i < m). Then the curve c is expressed
I.I. ARC LENGTH AND ENERGY OF CURVES
7
about c(to) as c(t) = (cl(t),... ,cm(t))-
The tangent vector c'(to) is nothing but the vector given as c'(to) =
( )
)
\ / (to) l a1 ) C(w
If c is a c' curve defined in a closed interval (a, b], noting that c is the restriction to [a, b) of a cd0 curve defined in an open interval containing [a, b], the tangent vector c'(to) of c at each point to E [a, b] can be defined as above. If c' (to) 0 0 at every to E [a, b], we call c a regular c'°0 curve.
Since M is a Riemannian manifold, we can measure the magnitude (the norm) Ic'(t)I of the velocity vector c(t) using the inner product g,,(t) defined in each tangent space Tc(t)M. In fact, ]c'(t)I is given by
Ic'(t)I =
ge(t) (c'(t), c'(t)), t E [a, b],
and is a continuous function of t. DEFINITION 1.3. Given a c°O curve c : [a, b] - M defined in a closed interval [a, b] (-oo < a < b < oo), the integral of Ic'(t) I
L(c) =
rb
Ja
Ic'(t)Idt
is called the length of the curve c.
For a piecewise smooth curve, since it is composed of smooth curves defined in closed intervals, we can define its are length as the sum of the arc lengths of those smooth curves. We have defined a curve as a map; hence, even though the image is fixed, we could arbitrarily choose the domain and the parameter. We see here that the arc length of a curve is a "geometric quantity" which is independent of domains and parameters. LEMMA 1.4. Let c : [a, b] M be a curve in M defined in the closed interval [a, b]. Let 9 : [a, ,(3] - [a, b] be a diffeomorphi mn. Then the curves c and c o 9 : [a, fl] -+ M have the same arc length; namely, the following holds: L(c) = L(c o 0).
I. ARC LENGTH OF CURVES AND GEODESICS
8
PRooF. By (1.5), (c o 0)'(t) = c'(0(t)) (d0/dt)(t) holds; hence, 13
L(c o 0)
J«
(c o 0)'(t) I dt = J«
) (t) dt.
1 c'(0(t)) I I
On the other hand, since 0 is a diffeomorphism, we may alway assume dO
dO >Oor
>0.
In the former case, noting 0(a) = a and 0(f3) = b, we have dO
L(c o O) _
lc'(0(t)) I
(t)dt =
6
IC'(t) I dt = L(c).
Ja In the latter, since 0(a) = b and 0(0) = a, we get L(c o 0)
dO
Ja
Ic'(O(t)) I
b
(t)dt JQ
1c'(t) ldt = L(c).
0 Given a c°° curve c : [a, b] - M defined in the closed interval [a, b], we set t
s(t) = j Ic (u)jdu. The function s; [a, b] - [0, L(c)] is called the are length of the curve c. In particular, if c is a regular curve, c' (u) 0 0 always holds. Then s is a monotone increasing function of t; therefore, the inverse function s-I : [0, L(c)] -, [a, b] exists. We define a c°° function by c o s-1 : [0, L(c)] -t M and call it the arcwise parametrization of c. When c is parametrized by are length, we have 1(co s-')'(s)J = (c'(t)IIG (t)I-1 = I. This implies that the magnitude of the tangent vector is always 1. DEFINITION 1.5. Given a C°Q curve c : [a, b] -, M defined in a closed interval [a,6] (-oo < a < b < oo), the integral of le,(t)12/2 Ic'(t)I2dt
E(c) = 1 a
is defined to be the energy or the action integral of the curve c.
For a piecewise smooth curve c, as in the case of the length, we first decompose c as a union of C°° curves, and then define the energy E(c) of the curve c as the sum of the energies of the COQ components.
1.2. EULER'S EQUATION
9
Although the length of a curve is independent of the parametriza-
tion, the same does not hold true for the energy of a curve. In fact, under the same conditions as in Lemma 1.4, it can readily be seen from the definition that E(c) j4 E(c o 0) in general. Nonetheless, as we shall see in the following sections, it is often more convenient to investigate the functional E(c) rather than to deal directly with the functional L(c) when studying variational problems (critical point problems) concerning length of curves. We will take up the relationship of L(c) and E(c) in §1.2.
1.2. Euler's equation Let (M, g) be an m-dimensional Riemannian manifold, and let c : [a, b] --; M be a C°° curve in M defined in a closed interval [a, b]. For the velocity vector c'(t) of the curve c, we defined, using the Riemannian metric g, its magnitude as Ic'(t)I =
9c(t) (c'(t), c'(t))
t E [a, b].
The length L(c) and energy E(c) were, respectively, given by b
L(c) =
E(c) = 2
I c'(t)I dt,
Jn
f
(c'(t) 12dt.
Let (x) = (xi, ... , xm) be a local coordinate system of M defined in a neighborhood of c(t). By setting ct = X' o c, c(t) is expressed as
c(t) = (C' M' ... , cm(t))As was seen in (1.6), c(t) is then given by i
(1.7)
c'(t) _ a-1
(s-)
(t)
(_) (t) E T°it)M. C
Denoting the components of g by gt.j, c' (t) local coordinate system (x), given by I
m I
c'' (t)
I = , 11:
I
is, in terms of the
(t) d d (t)
Now let (x) = (x',... , xm) and (Y) = (fi, ... , Y) be two local coordinate systems of M around c(t). The base of the tangent space
1. ARC LENGTH OF CURVES AND GEODESICS
10
T.,M
isfies the transformation formula
a
m
i
-x
axe (x) j,l
a axi
x
Hence, from (1.4), the components gig and gii of g relative to (x`) and (2i), respectively, satisfy the transformation formula m
gij(x) = F fti )I k,t=1
(1.8)
(x)
j
(x)gkt(x).
00
On other hand, from (1.7), we see that the components of the velocity vector c'(t) undergo the transformation: i
}(t)=
(1.9)
(c(t))(d(t).
1=1
From (1.8) and (1.9) it follows that the local coordinate representations of g (c' (t), &(t)) relative to the coordinates (x') and (xi) satisfy the equation ( 175' -&q (X)
t,,y=1
(t)
r ddt0
(t) =
gkt(x)
(1) (t)
(t).
k,t.1
As is seen from the definition, the above equation tells that the local coordinate representations of the inner product g,, (c' (t), c'(t)) given
by the Rican metric does not depend on the choice of local. coordinate systems. Therefore, we often define the length L(c) and the energy E(c) of a curve c, using a local coordinate system as b
L(c) a
m
dcdcJ
E gij dt dt 1j=1
,
1 lb ",
E(c) _
-2
> g#jdedt-dtdc)-dt.
Now we consider the critical value problem of L(c) and E(c) regarding the CIO curves in a Riemannian manifold M. Namely, we invesigate which c gives rise to a critical point among all the C°°
curves joining two points p = c(a) and q = c(b) in M. For the sake of simplicity, we assume that a curve c; [a, b] , M is a regular Coo curve whose image lies in a local coordinate neighborhood
1.2. EULER'S EQUATION
11
U. Let (x1, ... , xm) denote the local coordinate system in U. We define 2m independent variables x1,... , x"`, C1, ... , Cm as follows. Let
x= (x11 ...,xm, l,... 'Cm) E U,
ER. Weset for
L(c) Jr
(x, rr )
= .f (x1, ...
,xm ,1 , .. .
,
gijsiv,
Sm) _ s,j =1
and for E(c)
f(z,
gij s J.
)=J(21,...,x
i,j=1
Under these conventions, we can express both L(c) and E(c) as (1.10)
F(c) =
b f (c'(t)... Ja
,
cm(t),
11 (t)'...
dem ,
dt
(t} dt.
We find conditions for the integral F(c) in (1.10) in general to assume
a critical value. To that end, we arbitrarily choose a C°° function t defined in an open interval that contains [a, b] and which satisfies ry(a) = i7(b) = 0.
(1.11)
Using this 77, we set, for each i = 1,... , m and any sufficiently small real number c, c1(t; E)
= c (t) + Eij(t),
cj(t; E)
=
c? (t)
(i 54j)-
Then c(i; E)(t) = (C' (t; E), ... , Cm(t; E)),
t E [a, b],
gives a family {c(i; E)} of C°O curves in U. From the definition, for sufficiently small c, each c(i; E) gives rise to a regular curve joining p = c(a) and q = c(b). c(i; 0) gives c itself. For these curves, F(c(i; E)) is given by
F(c(i; E)) =
f
b
. , c+ e (t), ... , Cm (t), dc'
(t), ... ,
_
(t} + E dg (t),
... ,
d
(t) dt.
1. ARC LENGTH OF CURVES AND GEODESICS
12
This is a differentiable function of c in a neighborhood of 0. Hence, a necessary and sufficient condition for F(c(i; e)) to assume a critical
value at =0 is
d
(1.12)
0;
namely, {
axi
(c(t),c'(t))t7(t) + (j,)
(c(t)>
c'(t))
(t) dt = 0.
FIGURE 1.1. Family of cA0 curves {c(i; E)}
Here, if we apply integration by parts to the second term on the left hand side of the equation, noting (1.11), we get
Lb
(L' ) (c(t), c'(t))
I
b
L') (c(t), c(t))i7(t) In
(t)dt =
jab d a
d
--
l of (c(t), c'(t))q(t)dt (c(t), c'(t))j(t)dt.
a
a
(}
(c(t), (1(t)) } 9(t)dt = 0.
The above equation is equivalent to (1.13)
Jb 1 Of )
`
(c(t), fi(t))
-
d
1.2. EULER'S EQUATION
13
Noting that (1.13) holds for any q satisfying (1.11), we see that the following equation
(/)
(1.14)
e
(c(t), c'(t)) -
l
\
dj) (c(t), c (t)) = II
must hold. In fact, suppose that the left side of the equation (1.14) does not equal 0, say > 0, at some point to E [a, b]. Then, by continuity, the left side is positive in a neighborhood to C [a, b] of to. If we choose a C°° function 77 which takes positive values in 1o but 0 identically outside Io, the left side of (1.13) becomes positive, contradicting
the assumption (Exercise 1.3 at the end of Chapter 1).
From the above discussion, we have learned that if a curve c assumes a critical value of F(c) among the C°° curves joining p = c(a) and q = c(b), then (1.14) holds for i = 1, ... , m. We call the equation (1.14) Euler's equation for (1.11). We may consider F(c) as a function defined in the family of curve {c(i; e)}. More generally, a function
defined in a space consisting of functions or mappings is called a functional, indicating a function of functions. When F(c) is regarded as a functional, (1.12) is often called the first variation Jomtda of the variational problem concerning the functional F(c). Solutions to the first variational formula or to Euler's equation give extreme values of the corresponding functional; however, they may necessarily be neither relative maximal nor minimal in general. Therefore we call them critical points of the functional, and the values of the functional at those points are called critical values.
Now we actually compute Euler's equation (1.14). In order to distinguish the functionals L(c) and E(c), we will denote f for L(c) and E(c) by f 1(x, ) and f2 (x, ), respectively. By the assumption, the curve c is regular and C°°. Hence, the arc length parametrization of c gives, as was seen in § 1. 1,
Ic`(t)I = fi(c(t),c'(t)) - 1.
(1.15)
By the definition of f, and f2, we have f,(x,
) =
2f2(x, C)
Therefore, equation (1.14) for L(c) is given as
%22
(c(t),c'(t)) _
I(
-a
2
2) (c(t), c'(t)) = 0. 2
1. ARC LENGTH OF CURVES AND GEODESICS
14
From (1.15), this is equivalent to 12
ai } (c(t), e(t)) = 0. C
(c(t), c'(t))
(1.16)
This implies that under the assumption of (1.15), Euler's equations for L(c) and E(c) are both reduced to computing (1.16) for 21
f2 (X,
M
xm
I
,
m) = 2
E 9ij(xxiti.
i.j=1
We compute (1.16) for each i = 1,... , m. From the symmetric property of (g13), we get m
M
9%j We + 2 >9js(xW _
2 9=1
a9jk t
j,k=1
9ij(x) 'j, j=1
3=1
m
2
m
(X)Vk.
We now see that (1.16) is equivalent to m
d
dt
7=1
m
dcj 9ij (c(t)) dt (t)
_
d(t) = 0.
dc
(c(t)) ) dt (t) dt 2J.,k(Igjk ax= =1 1
This equation can be rewritten as in
m
j=2
9ij (C(t)) d2 2 (t) + d
E j,k=1
dck
ax } (CM) dt (t) at (t)
M
-2
dc
9ij
'9
j,k=1
ax' l) (C(t))
dcj (
dd-`tk
(t) = 0.
Hence, noting the second term being symmetric, we obtain (CM) (1.17) 1
m
da
13 (t)
a9,j
19ik
+2 j,k=1 (axk+axj
I
dck axi)(c(t)) d(t)d(t)=0.
19j k
1.2. EULER'S EQUATION
15
The positive definite matrix (gq., (x)) has its inverse matrix at each point x in U. Namely, we define the inverse matrix (gU(x)) by m 9:k
gik (x)9k1(x) = 8e ,
(x)9kj (x) = S9,
1 < i, j < m.
k=1
k=1
Then we get an m x m symmetric matrix (g'') whose components
tare C' functions in U. Using g' J, we define in U a family
jk
(1 < i, j, k < m) of C°° functions by
{i}_1gu(o91i+a91ko9ik
(1.18)
{
2
a-1
8XI
jk 's are called Christo f'el symbols, and they occupy an important
position in Riemannian geometry. We point out here that
jk }'s are
determined solely by the components gt, of the Riemannian metric
g in the local coordinate system (x'), and that they are independent of the curve c. Using the Christoffel symbols, we can rewrite (1.17) in a simpler form. In fact, if we replace the script i by 1, and add all the equations obtained by cross-multiplying g`l to (1.17)
over I=1,...,m,weget i
(1.19)
m
J-(t)+ E {;k}(C(t))(t)(t)=o1
1
i
m.
This is what we have obtained by actually computing Euler's equation for L(c) and E(c). In general, the curves satisfying (1.19) are called geodesics in a Riemannian manifold. DEFINITION 1.6. Let C : [a, b] --; M be a C°° curve in a Riemann-
ian manifold M. In a coordinate neighborhood U of each point in c, we denote by ` C(t) = (C1(t), ... , Cm(t))
the coordinate representation of c near the point. Then the curve c is said to be a geodesic if each c' (t) (1 < i < m) satisfies equation (1.19).
From the above observations, we have realized that among the
C°O curves joining two points p = c(a) and q = c(b) in U, the
1. ARC LENGTH OF CURVES AND GEODESICS
16
geodesics, when parametrized by arc length, are the extreme points of L(c) and E(c). We point out here the following fact. Let c : [a, b] - M be a C°O curve in M. Suppose there are two local coordinate systems
(xi) = (.T',. - , x') and (t) = (ct1, ... , x") in an open subset U, through which c passes. Then the Christoffel symbols { 3' } and 1JJ
jk
satisfy the following relationship (see Exercise 1.4 at the end
of this chapter): M
7k
5P=I
ati
m
v-xp
J
p
k
+ 9E1
07x9 01xT Q7'
' 52k
On the other hand, when c is represented as c(t) = (c, (t),
.
en (t))
{e1(t), . , . , c""(t)), in these coordinate systems, respectively, (1.9)
holds between the components of the tangent vector ca(t). If we pay attention to these transformation formulas and the coordinate transformation in U 0xi dxk axk 1
_
°
a simple computation yields the following relationship: m
dt2
(t) +
tc1
k
(OXIdt
(t)
(t)
(t) + j,k=1
{}(c(t))(t)(t)). jk dt dt
From this relationship follows that if the curve c in U satisfies (1.19) in the local coordinate system (x1), (1.19) also holds in (if). This fact suggests that the geodesics in Riemannian manifolds can be defined geometrically, independent of local coordinate systems.
In the next section, we will take up this problem from a more general point of view.
1.3. Connections and covariant differentiation Let (M, g) be an m-dimensional Riemannian manifold and let [a, b] --y M be a C°° curve in M defined in a closed interval [a, b]. In the last section, we derived the equation of geodesics from
c:
1.3. CONNECTIONS AND COVARIANT DIFFERENTIATION
17
the first variation formula of length and energy, when the image of c is contained in a coordinate neighborhood of M. In this section, as part of preparation for a general first variation formula, we discuss differentiation for vector fields. For the time being, we assume M to be simply a C°O manifold and U an open subset of M. A vector field X = {X (x)}iEU in U is defined to be a correspondence that assigns to each x E U, a tangent vector
X (x) E T=M. In particular, if U = M, X is simply called a vector field in M. Let (xe) = (x1, ... , xm) be the local coordinate system of an arbitrary local coordinate neighborhood V which intersects U. At each point x E U n V, X (x) is uniquely expressed as T (1.20)
X(x) _
(x)
(axi)Them
functions l;' (1 < i < m) are called the components of X. When the components are C°° functions, X is said to be a C°° vector field in U.
For example, assuming U to be a local coordinate neighborhood, set
ax=) (x) - a_i)
x E U.
Then we obtain a C°° vector field in U :
l
x(JJJ ZEU
1
The family of these vector fields {
ural frame in U.
ll
i
I 1 < i < M } is called the natJJ
Given an open subset U in M, denote by 1(U) the set of all C°° vector fields in U. Also denote by C°° (U) the set of all C°° realvalued functions defined in U. C°°(U) is a commutative ring under the ordinary sum and product of functions. On the other hand, given any X, Y E 1(U) and f E C°° (U), we define X + Y and f X, respectively, by
(X +Y)(x) = X(x) +Y(x),
(fX)(x) = f(x)X(x), x E U.
As can be seen readily, then X + Y, f X E X(M). Fbr any X, Y E 1(M) and for any f, h E C°°(U), we have f (X + Y) = f X + f Y, (f + h)X = f X + hX, (f h)X = f (hX ). Hence, we see that 1(U) is a C°° (U) module.
1. ARC LENGTH OF CURVES AND GEODESICS
18
Now given X E X(U) and f E C°° (U), we can define a function
Xf in U by
(Xf)(x)=X(x)f, xEU. By (1.20), we have, in the local coordinates (xi), m
(X f)(x) = >Ei(x)af (x),
(1.21)
x E u n V.
i=1
Hence, X f E C°°(U). We call X f the derivative of f by a vector field X or in the direction of a vector field X. From (1.21), we see that if X(x) = 0, then (Xf)(x) = 0. We see dearly that the following properties of the differentiation hold: (i)
(ii)
(iii)
X(f + h) = X f + Xh, (X + Y)f = X f +Yf, X (f h) _ (X f )h + f (X h).
Given X, Y E X(U) and f E COO (U), we set
[X, YJ(x)(f) = X(x)(Yf)
- Y(x)(X f ),
x E U.
Then we can readily see that
[X,Y](x)(fh) = ([X, Y](x)f)h(x) + f(x)([X, Y](x)(h)) This implies [X, YJ E T,,M; hence, [X, YJ = { [X, YJ (x) }1EU defines a vector field in U. Let V and 77i denote the components of X and Y
in the local coordinate system (xi), respectively. Then noting
a
i' axe
(x)f = axaa_ -
i)I (x) = 0,
1 < 2,. <_
,
a simple computation verifies that the components of [X, Y] are given by m m J [X, Y1(x)
ai
(x) (i). x
Hence, [X, YJ E 3E(U)
The vector field [X, Y] is called the commutator product or the bracket product of X and Y, and it has the following properties (see Exercise 1.5 at the end of this chapter). Namely, given X, Y, Z E X(U)
1.3. CONNECTIONS AND COVARIANT DIFFERENTIATION
19
and f, g E Coo (U), the following hold: (i)
[X, Y + Z] = [X, Y] + [X, Z], [X + Y, Z] _ (X, Z] + [Y, Z],
(ii) (iii)
[X, YI = --[l', X], [f X, hY] = fh[X, Y] + f (X h)Y - h(Y f )X,
(iv)
[X, [Y, Z]] + [Y, [Z, XI] + [Z, [X, Y]] = 0, Jacobi identity.
These properties imply that X(U) forms a Lie algebra over R. Corresponding to the differentiation of functions by vector fields, the differentiation of vector fields by vector fields is an operation called covariant differentiation, and it is defined as follows.
DEFINITION 1.7. Let M be a C°° differentiable manifold, and let X(M) denote the CI (M) module of all C°° vector fields in M. Then a bilinear map
V : X(M) X X(M) -> X(M) satisfying the following conditions (i),(ii) is called a linear connection or an affine connection in M :
V(fX,Y) _ fo(X,Y), V(X, fY) (Xf)Y + fV(X,Y).
(i)
(ii)
Here, X,Y E X(M), f E C°°(M). Given a linear connection V, we express V (X, Y) as VxY, and call this quantity the covariant derivative of Y by X. By the definition, it should be noted that the covariant derivative Vx Y is characterized by the following rules: (i)
Vx(Y+Z)=VxY+VxZ,
V x(f Y) _ (X f )Y + fVxY, (iii) Vx+YZ = VxZ + VYZ, (ii)
(iv) vfXY = fvxY. For example, if M is an m-dimensional Euclidean space E, a C°° vector field Y in M, by definition, can be regarded as a Rm-valued
C°° function defined in M. Accordingly, Y = (f 1, ... , r) . Using the notion of the differentiation of functions by a vector field, we set v x Y = (X f 1, ... , X f m) . Then V x Y clearly satisfies the above
conditions; hence, it defines a covariant differentiation of Y by X. This linear connection V is called the standard connection of E. The properties (ii), (iv) imply
1. ARC LENGTH OF CURVES AND GEODESICS
20
LE mA 1.8. Given X, Y E 1(M) and an open subset U, either XJU - O orYJU - O implies VxYlU -= 0. PROOF. Assume Y I U = 0. For any x E U, choose an f E C(M )
such that f (x) = 0 and f
1 outside U. Then f Y - Y; hence, at x
we have
(V xY)(x) = (VxfY)(x) = (Xf)(x)Y(x) + f (x)(V xY)(x) Since x is arbitrary, we get V YIU - 0. A proof for the case where X 1U - 0 can be obtained in a similar manner.
0
COROLLARY 1.9. The linear connection V an M naturally in-
duces a linear connection VU :1(U) x 1(U) -1(U). PROOF. All we have to do is determine the covariant derivative (Vu )xY for any X, Y E X (U). Given x E U, goose an open neighborhood V of x such that the closure V of V is contained in U. Now c h o o s e a f u n c t i o n f E C°O (M) such that f - l and f 0 outside U.
Define X,YEX(M)to be X = fX, Y= fY.Then ,X'=X, Y-=Y in V. Define, at any x E V,
((Vu)xl')(x) = VX1'(x) Lemma 1.8 and the linearity of V imply that the right hand side is determined independently of the the extensions k, Y of X, Y. (V )xY being a covariant derivative is clear from V Y(x) being a
0
covariant derivative. Prom now on, we also denote VU by V for simplicity.
Let U be a coordinate neighborhood of M and let (z') denote the loco) coordinate system in U. Using the natural frame {a/&xi}, 1 < i, j, k < m} in U we can define a family of C'°° functions (1 I
as follows: (1.22)
V
j=EL13 k=1
These { r, 11 < it j, k < m} are called the amnection coefficients of the linear connection V relative to (zi) . Given X, Y E X(M), denote by
X= :1
s,
Y=En`ax; ;=1
1.3. CONNECTIONS AND COVARIANT DIFFERENTIATION
21
the component representations of X, Y in the coordinates (x{), respectively. Then the covariant derivative VXY is given by
pvT
VxY =
j is=1
,a
'r'0X(I j
m
)
l
=
axi axj
8 axk'
=,J,k=1
namely, m
k
VXY =
(1.23)
in
k=1
j=1
i=1
Hence, if X(x) = 0 at each x E U, (VXY)(x) = 0. Given v E TIM and Y E X(M), if we choose an X E X(M) and set Vq,Y = (VxY)(x) E TIM, we see, as in Corollary 1.9, that the right hand side is well defined, independently of the choice of an extension X of v. We call this quantity the covariant derivative of Y E X(M) in the direction V E TIM.
Furthermore, a stronger statement can be made with regard to the vector field Y, which is covariantly differentiated. Let X, Y E
X(M) and let c : [0, c) -+ M be a C°° curve such that c(0) = x and c`(0) = X (x). Then we see that (VxY)(x) is well determined depending only on X (x) E TxM and {Y t) 0 < t < E}. In fact, let (x1) be the local coordinate system about x. Since c'(0) = X(x) implies ` (x) = (dc' /dt) (0), the first term on the right hand side of (1.23) becomes in
oe )
in
(x)
dei
(0)
(c(0)) =
d{>1k
c) (0).
;=1
=1
Given a C°° curve c
:
(a, b) -+ M, a correspondence X =
{X(t)} fE(a,b) which assigns to each point t E (a, b) a tangent vector X (t) E TT(t)M is called a vector field along the curve c. When we express it, using the local coordinate system (x`) about c(t), as m
X(t)= i=1 fa(t)axi
c(t)
1. ARC LENGTH OF CURVES AND GEODESICS
22
at each point to E (a, b), we see that e (1 < i < m) is a function defined in a neighborhood of to. If they are C' functions, we define X to be a C°° vector field along the curve c. Let [a, bJ be a closed interval and let (a - e, b + E) (e > 0) be an open interval that contains [a, b]. The restriction of a CO° vector field X = {X (t)}tE(a.. ,..f) along a C°° curve c : (a - e, b + e) --+ M is called a CO° vector field along the C°° curve c : [a, b] - M. For example, the tangent vectors of a C°° curve c give rise to a C°° vector field d = {c' (t) } t et,,,bl along the C°° curve c. This is called the tangent vector field of c. Also given a C°° vector field X in M, a vector field along c defined by X o c = {X (c(t)) } is a C00 vector field along c. This is called the restriction of a C°° vector field X to c. The set of all C°'° vector fields along c will be denoted by X(c). Given X, Y E X (c) and f E C°° ([a, b) ), we define vector fields X + Y and f X, respectively, by
(X + Y)(t) = X(t) + Y(t),
(fX)(t) = f(t)X(t), t E [a,bJ.
Then they are C°° vector fields along c. Under these operations X(c) becomes a C°'° ([a, bJ) module.
PROPOSITION 1.10. Let V be a linear connection on a CO° manifald M. Then there is a uniquely determined linear map
D : X(c) -+ X(c) satisfying the following: (i) given X, Y E X(e) and f E C°° ([a, bJ),
D (ii)
(X+Y)= aX +
dDt(fX)=
x+f DX
;
if X E X(c) is the restriction of X E X(M) to c,
DX = Ve(t)X, att
tE
[a, b].
PROOF. First, we verify uniqueness. As before, let to E (a, b] and let (a{) denote a local coordinate system of M about c(to). We respectively represent c in (x*) by c(t) = (c1(t), ... , cm(t)) and X E X (c) by in
X (t) _ i=1
fi(t) 49 s
) c(t)
1.3. CONNECTIONS AND COVARIANT DIFFERENTIATION
23
From (i), (ii) and (1.7), we get
DX
dt - j=1
(1.24)
= k=I
a
a
dt axj +
axj
(c+
Mr 13
a
(c)d
axk
#,j=1
DX
(to) is uniquely determined. Conversely, if we define DX /dt by (1.24) in a coordinate neighborhood intersecting with c([a, b]), we can readily verify that the expression satisfies (i), (ii). It follows, therefore, that if UnV flc([a, b]) # 0, DX /dt' defined in U and V as above coincides in U n V due to the Hence,
uniqueness. This implies that Dt : X (c) - X (c) is well defined.
0
Given X E X(c), D/dt E X(c) determined in Proposition 1.10 is called the covariant derivative along c of X. In particular, when D/dt = 0, X E X (c) is called a parallel vector field along c. The following holds.
PROPOSITION 1.11. Denote by V a linear connection in a C°° manifold M, and let c : [a, b] -p M be a Cl* curve in M. Given an arbitrary tangent vector v E Tc(a), there is a unique vector field X E X(c) parallel along c with X (a) = v.
PROOF. First, we prove the proposition when c([a, b]) is contained in a coordinate neighborhood U. Denote by (xi) the coordinate system in U. In terms of (x'), we express c and v by c(t) _ (c1(t), ... , cm(t)),
a
v= s=1
axe
l
) c(t)
respectively. We set the desired vector field X E X(c) as X (t) El n, V (t)((a/(9xi),(t)). Then, from (1.24),
DX
m
(t) _
m
k
(t) + i=1
i,?_I
i
F z.(c(t)) 4dt (t)tj(t)
axa
x
()()
_
1. ARC LENGTH OF CURVES AND GEODESICS
24
Hence, in order for DX/dt = 0 to hold, it suffices that each Ek(t) satisfies the following system of differential equations: to
(1.25)
£(t) + 1: I ;j (c(t))
o,
1 < k < M.
This is a system of first order linear ordinary differential equations. Hence, given an initial condition (v1, ... , v), there is a unique soluv' (1 < i < m) tion (41, ... , ) to (1.25) defined in [a, b] with due to the fundamental theorem regarding the existence and uniqueness of the systems of linear ordinary differential equations (see Rikigaku to Jyobibun Hoteishiki, Gendai Sugaku e no Nyumon, Iwanami Koza). Next, in the general case, noting that c([a, b]) is a compact subset, it can be covered by a finite number of coordinate neighborhoods, in which the proposition holds. By uniqueness, the solutions coincide in
the intersections of the neighborhoods to produce a solution defined 0 in the entire [a, b]. This completes the proof.
In Proposition 1.11, we call the tangent vector X(b) to M the vector obtained by parallelly displacing v = X(a) along c. By oorresponding X (b) to X (a), we get a map P,, : TT(Q) 4 Ta(b) between the tangent spaces. We call Pc the parallel displacement along c.
It is easily seen from the uniqueness and existence of solutions to the system of linear ordinary differential equations (1.25) that the parallel displacement P, : TT(0) -> T,(b) is a linear isomorphism form Te(a) onto TT(b).
Now we go back to the Riemannian manifold (M.9). The following
result, often called the fundamental lemma in Rieannian geometry, is a theorem, which is to be the starting point of Riemannian geometry.
THEOREM 1.12 (Levi-Civita). Let (M, g) be a Riernannian manifold.
Then among the linear connections on M, there is a unique
linear connection V such that for any X, Y, Z E X (M), (i) (ii)
X g(Y, Z) = g(V XY, Z) + 9(Y, V X Z),
VXY - VYX = [X, Y].
1.3. CONNECTIONS AND COVARIANT DIFFERENTIATION
25
PROOF. We first show uniqueness. Assuming that such a V exists, from (i) follows
X9(Y, Z) = 9(VxY, Z) +9(Y: VxZ), Yg(X, Z) = g(VYX, Z) + g(X, VYZ), -Zg(X, Y) = -g(V ZX, Y) + g(X, VZY). Adding the left hand sides together and the right hand sides together, respectively, and using (ii) yields
Xg(Y, Z) + Yg(X, Z) - Zg(X, Y) =g(V xY, Z) + g(V YX, Z)
+g(X,VYZ - VZY) +g(Y,VXZ - VZX) =2g(VxY, Z) +g([Y,X], Z) + g(X, [Y, Z]) + g(Y, [X, ZI). Namely, we get (126)
2g(VxY, Z) = Xg(Y, Z) + Yg(X, Z) - Zg(X, Y)
- g(X, [Y, Z}) - g(Y, [X, ZJ) + g(Z, [Y, XD.
Noting that the Riemannian metric g defines a nondegenerate inner product gx in each tangent T. ,M, we see from (1.26) that V XY can be uniquely expressed in terms of g and the commutators of vector fields. Hence, if V exists, it is unique. Next, we show existence. Given X, Y E X(M), define V xY by (1.26). It is easily verified that VxY defines a covariant derivative from its definition and the properties of the commutator product. We immediately see from (1.26) that VxY satisfies the conditions (i),(ii) 0 in the theorem. DEFINITION 1.13. Given a Riemannian manifold (M, g), the linear connection V in Theorem 1.12 is called the Levi-Civita connection or the Riemannian connection of (M, g). In general, a linear connec-
tion V is said to be compatible with the Riemannian metric g if it satisfies the condition (i), and symmetric if it satisfies the condition (ii), respectively.
From now on, we will consider the Levi-Civita connection in a Riemannian manifold (M, g) unless otherwise mentioned. Let (x) be a local coordinate system in M and let I' denote the 9 connection coefficients in (x_) of the Levi-Civita connection. From
1. ARC LENGTH OF CURVES AND GEODESICS
20
(1.22) and (1.26) follows
V.
1 (89j,
k
rji Stet - 2 k-= I
OX'
499;1
+ ax.
_ a9s;) ax!
/
If we change the index k to h, multiply both sides by the components (gk1) of the inverse matrix (g$) of (g j ), and add them over 1, we get (1.27)
1
t=i
a a9-a, ag;t gk1\Ox +Ox - Oz1 !II
Hence, it follows that the connection coefficients 1' 's of the LeviCivita connection are nothing but the Christoffel symbols
g3
's.
ExAMPLE 1.14. If M is Euclidean space Em, we have
1'ij = d, 1 < i, 3, k < m, since gE, = &,f in the natural frame relative to the coordinate functions in R'. Hence, from (1.23), we see that the Levi-Civita connection of
E"' is precisely the standard connection. If X E X(c) is a parallel vector field along a curve c, (1.24) implies that each component of the vector becomes a constant; hence, X E X(c) is parallel in the sense of Euclidean geometry.
As is readily seen from (1.22) and (ii) of Theorem 1.12, a linear connection that is symmetric is equivalent to the connection coefficients 14jj's that are symmetric in the indices i, j; namely,
I
=F
,
1<_i,j,k<m.
On the other hand, that V is compatible with the Riemannian metric g implies the following: PROPOSIT ON 1.15. Let V be a linear connection compatible with
the Riemannian metric g in M and let e, (a, bJ -+ M be a CIO curve in M. "en the following hold: (1) Given any X, Y E X(c), // d . 'Y) +g `X, dt clatg(X'Y) 4t (2) The prat displacement Pea) : T4,(a)1bl TG(b)M along C is an isometry; namely, PP preserves the inner product induced from
-9
9.
l
1.3_ CONNECTIONS AND COVARIANT DIFFERENTIATION
27
PROOF. (1) is nothing but rewriting (i) in Theorem 1.12. (2) For any v, to E T ,(,,)M, choose X, Y E 1(c) so that X(a) = v,Y(a) = w hold. By (1), we have
(x(t),
dt 9c(t) (X (t), Y(t)) = 9e(t)
Ddt
(t}
= 0.
This implies that ge(t) (X (t), Y(t)) is a constant; hence, we have 9e(t) (Pc(v), PP(w)) = 9c(t) (v, w)-
0 Now we calculate the covariant derivative Dc"/dt along c of the
tangent vector field c' E 1(c) of a Coo curve c : [a, b] -* M. In a local coordinate neighborhood U which c passes through, we express c(t) = (ci (t), ... , c"'(t). Then i
'(t) From (1.24) it is readily seen that Dc'/dt is given by Dcf
dt
"` k=1
deck dt2
m
+
(c
dc' dc)
8
dt dt
09xk
Hence, the fact that c is a geodesic; that is, c satisfies, in each local coordinate neighborhood that c passes through, the equations of geodesics dt2
+
l'k(c)dc;dc-
tJ
dt dt
=0, 1
is equivalent to the fact that the equation Dc'/dt = 0 holds; that is, the tangent vector c' of c is a parallel vector field. We may redefine a geodesic as follows.
DEFINITION 1.16. A C° curve c : [a, b] -* M in a Riemannian manifold (M, g) is a geodesic if the tangent vector field c'(t) E 1(c) of c is parallel along c; namely, Dc' dt
=0,
t E [a, bJ,
holds.
The following is evident from Proposition 1.15.
1. ARC LENGTH OF CURVES AND GEODESICS
28
PROPOSITION 1.17. The magnitude of the tangent vector field of
a geodesic c in a Riemannian manifold (M,g) is constant, namely, g(c'(t), cf (t)) = a constant.
1.4. Geodesics Let (M, g) be a Riemannian manifold of dimension m and let c : [a, b] -- M be a C' geodesic in M. As was seen in the last section, the tangent vector field c' E X(c) of the geodesic c is parallel along c. It implies that
! (t) = Ve(tyc'(t) = 0,
(1.2$)
t E [a, bl;
hence, the magnitude Ic'(t) I of c' is constant. In what follows, denote the constant by Ic'(t)I = co ,t 0. Namely, we do not consider a geodesic whose image is a point. Since the are length of c is given by t 8 (t)
=
fa
Ic'(a)Idv = co(t - a),
the parameter t of the geodesic is proportional to the arc length s. In general, t is called an afne parameter. In particular, if co = 1, that is, Ic'(t)I = 1, c is called a unit speed geodesic. Let U be a local coordinate neighborhood that c passes through, and let c be expressed as c(t) = (c'(t), ... , c(t)) in the local coordinates. In U, (1.28) is equivalent to
+
(1.29)
W_
fJ(cl, sj=1
-
en)
dt dt
=0,
1 < k < m.
This is a system of second order nonlinear ordinary differential equa-
tions. By introducing a new variable e = c /dt (1 < i < m), the system can readily be transformed into a system of first order normal linear ordinary differential equations: dek
(1.30) -et-
Applying the fundamental theorem on existence and uniqueness regarding the systems of first order ordinary differential equations (see Riligaku to Jyobibun Hoteishiki, Gendal Sugaku e no Nyumon, Iwanatni Kola) to (1.30), we get the following proposition.
1.4. GEODESICS
29
PROposrriON 1.18. At each point p in M, there exist a neighborhood V of p and positive numbers fl, E2 > 0 so that the following holds. Given a point q E V and a tangent vector v E TqM at q with jvj < E1i there exists a unique geodesic c (-E2, E2) -+ M defined in Itl < E2 and satisfying an initial condition C'(0) = v.
cv(0) = q,
From the uniqueness of geodesics in this proposition, we see that there is a certain homogeneity property, in the following sense, regarding the initial vectors and the intervals of existence of geodesics. LEMMA 1.19. Let c : (-E2, E2) - M be a geodesic with v E TqM
as its initial vector. Then given any positive number a > 0, there exists a geodesic cs,, : (-E2/a, E2/a) - M such that c(at) = Gav (t),
t E (-E2/a, E2/a)
PROOF. In fact, if we define a C°° curve c : (-E2/a, E2/a) -+ M by
c(t) = c(at), t E (-e2/a, E2/a), c' (t) = ac,(at) implies c(0) = q,
cr(0) = av,
Dc'
= V'c' =
a2V,,c/
= 0;
hence, c is a geodesic with an initial condition av. By the uniqueness
in Proposition 1.18, we get c = cr,,,. Namely, cn(at) = can(t), t E D
(-E2/a, t2 /a) holds.
From Lemma 1.19, we see that we may uniformly choose the interval of existence for geodesics about a given point p in M in Proposition 1.18; namely, THEOREM 1.20. At each point p in M, there exist a neighborhood
V of p and a positive number E > 0 such that for any point q E V and any tangent vector v E TqM with lvj < E, there exists a unique geodesic
c,,:(-2,2)-*M
satisfying the initial conditions
4(0)=q,
r.`(0)=v.
1. ARC LENGTH OF CURVES AND GEODESICS
30
PROOF. By Proposition 1.18, there exists the geodesic cg(t) for v E T,,M with (vI < el and Itf < E2. By Lemma 1.19, we can define a geodesic ;.,v/2(t). Hence, if we choose c > 0 satisfying f < E1e2/2, we can define a geodesic c (t) in Itl < 2 for any v E TqM. 0 We point out here that it was crucial, in the proofs of this theorem
and Proposition 1.18, to transform the equation of geodesics (1.29) to the system of first order ordinary differential equations (1.30). In other words, although equation (1.29) is a system of second order ordinary differential equations with m unknowns c1(t) in a coordinate neighborhood U in M, by introducing new variables J` = c'', it is reduced to a system of first order ordinary differential equations (1.30) with 2m unknowns c' (t), *. This implies that the equation (1.29) is
a second order equation with unknown c(t) = (c'(t)), but we can consider it as a first order equation (1.30) with the pair of unknowns
(CM, c'(t)) = (c(t), c'(t)). This suggests that we should regard the equation of geodesics as a first order equation defined in the tangent
bundle TM over M rather than a second order equation in M. We now look into this situation a little more closely.
In general, given a C°° manifold M, the set of all the tangent spaces in M
TM= U TpM={(p,v)JpEM, vETpM} pEM
is called the tangent bundle of M. The map it : TM -s M defined by ir(p, v) = p,
(p, v) E TM
is called the projection of TM onto M. sr- I (p) = T,,M is called the fiber of the tangent bundle TM over p. Let (e , ... , z") be the local coordinate system in a local coordinate neighborhood U about p. A tangent vector v E TqM at q E U is uniquely written as m (1.31) i=1
.
ax4 Q
(1.31) determines a Conversely, given m real numbers tangent vector v E T,QM. Therefore, each element (q, v) E TMIU = LJPEU TpM is in one-to-one correspondence with a 2m tuple of real numbers EUx (x'(q),... ,xm(q), Rm.
1.4. GEODESICS
31
Namely, the set of all 2rn-tuples (1.32) (x1, ... , xm, 1, ... , C,n) E TMIU = 7r-1(U)
gives rise to a bijection of TMJU onto the direct product U x Rm. In this context, if we introduce by (1.32) a local coordinate system of TM in TMJU = it-1(U) for each local coordinate neighborhood U in M, TM naturally becomes a C°° manifold of dimension 2m. It is readily verified that the projection it is a C°° map from TM onto M (see §A.1(e), Appendix). In other words, we can naturally introduce
a C°° manifold structure in the tangent bundle TM of M so that each T MI U as an open submanifold becomes C°° diffeomorphic to
the product manifold U x R'n. From now on, we assume that TM is equipped with this COQ manifold structure. In particular, if (Mg) is a Riemannian manifold, it should be noted that each fiber 7r-1(p) carries an inner product gp. Now let c : [a, b] --+ M be a C°° curve in M. For each t E [a, b], since (c(t), c'(t)) E TM, c(t) _ (c(t), c'(t)) gives rise to a COO curve in TM c : [a, b] - TM. This means that the tangent vector field c' E X(c) of c defines a C°° curve C = (c, c') in TM. In particular, let c be a geodesic. In a coordinate neighborhood U through which c passes, express the curve c by c(t) = (ci(t)) and set e = cs'. Then it is readily verified that c(t)
is a solution to (1.30) in T M U. Namely, (1.30) is nothing but the system of equations that define the tangent vector field c' E X(c) of the C°O curve c = (c, c') in TM determined by the geodesic c. This is the geometric meaning of (1.30) interpreted in terms of the tangent bundle (Exercise 1.7, at the end of this chapter). Going back to our original setting, for the neighborhood V C U of a point p E M, as stated in Theorem 1.20, we define a subset U of the tangent bundle TM by U={(q,v)IgEV, vETgM, lvJ
t E (-2,2), (q, v) E U,
we obtain a map
p : (-2,2) x U -+ M.
1. ARC LENGTH OF CURVES AND GEODESICS
32
Noting that the solutions to the system (1.30) of ordinary differential equations of order 2 are C°° dependent of initial conditions (see Rikigaku to Jobibun Hoteishiki, Gendai Sugaku e no Nyumon, Iwanami Koza), we easily see that,0 is a C°O map from (-2,2) x U into M. In other words, c=,(t) is C° dependent of the parameter t and the initial condition (q, v). Based upon the above discussions, we define the exponential Wrap
exp:U --M by (1.33)
exp(q, v) _ 4(1J q, v),
(q, v) E U.
Since 0 is a C°O map, exp is also a C°° map. In particular, given each q E V, the restriction of exp to the tangent space TqM of M; namely,
the map expq : BE(0) C TqM - M defined by expq(v) = exp(q,v) is called the exponential map at q.
Here, A(0) denotes the c neighborhood of the origin 0 of TqM, namely,
B,.(0) = {v I V E TqM,
Ivy < e}.
From the definition (1.33), expq(v) is nothing but the point at the unit time 1 along the geodesic emanating from q with the initial vector v. Clearly, expq is a Cc' map and expq (0) = q. EXAMPLE 1.21. Let M be the it-dimensional Euclidean space El. As was seen in Example 1.14,
r aj = 0,
1 < i, j, k < era,
with respect to the natural frame determined by the coordinate functions (xf) of ]Rm. Here, the equations of geodesics are given by deck
dt2
= 0.
This implies that the geodesics in Em are nothing but straight lines. Hence, if we identify the tangent space of Em with Em via a parallel Em is the identity translation, the exponential map expp : ]L' map. EXAMPLE 1.22. Let M be the unit sphere in Em+1 : .Sm =
{(1 , ... ,
m+l)
>Z=1
(x)2 = JJJ
1.4. GEODESICS
33
As a Riemannian metric in S, we consider the induced metric. If we set
U={(x',...,xm+1)ES"xm+1>01 and define a map 0: U 0(X1
lRm by m+1
(
m)
1
(U, 0) gives rise to a coordinate neighborhood of S. Noting that
¢-1(x1,...,x"`)= xi,...,x"=,
1-
m
(xi)s
gives in U the embedding of S' into Eit+i, the components gij of g, gl) and the connection coefficients n-'s are, through simple computations, given, respectively, as X)
gi1 = di - xi a
9;j = 6;J + (X,,+1)2' X,
r
1
= gil, 1 < i, j, k < m. j Assuming that c is a normalized geodesic, the equations of a geodesic d2ck
dt
x1`
m
+ i,7=1
i
ckgs;
dt d
=0,
1 < k < m,
are reduced to a system of second order ordinary differential equations of constant coefficients (1.34)
1
dCk+ck=0,
since dc-i
E g'j dt dt
1,
1 < k < m,
ij=1
holds.
The system (1.34) is easily solved. For example, let v = (vi, ... vt, 0) E TT S m C TpEm+1 = Em+1 be a unit length t a n g e n t v e c t o r at p = (1, 0, ... , 0) E Sm. Then the geodesic c satisfying initial conditions c(O) = p, c'(0) = v is, as a curve in 1E'n+1, given by cs(t) = sin t vs, c'+i (t) = cost, 1 < i < M. This geodesic c is nothing but the great circle (or a portion of the great circle) obtained as the intersection of S' and the plane spanned by the position vector (0,. .. , 0, 1) and (vi, ... , v"', 0) in Em+1. From the above observations, we see that the exponential map expp is defined in
-
34
1. ARC LENGTH OF CURVES AND GEODESICS
the entire tangent plane TpS'm, and is an injection from B,r (0) C TpSm
onto Sm\{q} (q is the antipodal point of p). It also maps the entire boundary of B,,(0) into q. ExAMPI.E 1.23. Let M be the upper half
Hm= ((x',... ,x')
IZt >0}
of the mrdimensional number space fl(m . Let g be a Riemannian metric
in HI defined, relative to the coordinate functions (x') of lRm, by m
(x:n)2
1:(dXs)2. --.1
The Riemannian manifold (H'4, 9) thus obtained is called the veal hyperbolic space of dimension m or the Poincare upper half-space.
FIGURE 1.2. Geodesics in S'm
From the definition, the components gi,, and g'i of g relative to the coordinate functions (x') are given by = (xm)-2bij. 9ti (x,"a)2d".
=
90
Via simple computations, we see that -('m)-1,
I';ni = =--(x'm)-1,
1,,.n
l
_ (xm)-1,
1
1.4. GEODESICS
35
are 0. Hence, the
and the rest of the connection coefficients equations of geodesics are given by 2 dc' dcm
d2c'
1 < i < n2 -
0,
dt2 - c"` dt
fm-1
1
d2C"'
dt
dC'
dt} -
+c.n
dt2
2
dcm
dt
2
}
=0.
1 m2 (dc')
Assuming that c is a normalized geodesic, since (c, )2
dt=1,
s-1
we finally get (1.35)
d2c' dt2 d2c'ii dt2
2 dc' dcm c" dt dt 2 cm
dcm
dt
_ 2
0,
1 < i < m - 1,
+c"' =0.
It is easily verified that the general solution of the system (1.35) is vm-1, 0) and w = (w1, ... , wm-1, 0) given as follows. Let v = (vl, . 7
be vectors in R'"-1 C Rm, and let lvi = 1 as a vector in E. Then given constants a > 0 and b, the solution is given by
c'(t)
=
Cm (t) = et+b, wit
1 < i < rn - 1.
H"
FIGURE 1.3. Geodesics in Hm
36
1. ARC LENGTH OF CURVES AND GEODESICS
The curves represented by these solutions move, in the former case, along the hemi-circle of radius a with center w, perpendicular to E", and in the latter, along the half-line passing through w and parallel to the x' axis. These represent the geodesics in (H"=).
Consequently, we see that for a given p E H"`, the exponential map expp is defined in the entire tangent plane T, and expp : H"' is a C°i diffeomorphism from TpH12 - R"s onto H"`. TTHm In general, the exponential map is always a local diffeomorphism. Namely, we have THEOREM 1.24. Given each point p, there is a positive number e such that the exponential map
expp:BE(0)CTpM-+M gives rise to a C°° difeomorphism from B,(0) onto the open subset expp(B,.(0)) of M. PROOF. We compute the derivative d(expp)o of expp at 0 E TpM. Identifying To(TM) with TpM, we get, for a given v E TpM,
d(expp)o(v) =
=
d (expp(tv)) dt
= z=o
d
dt W1, P, tu)) 1t=0
=V.
dt(So(t,p,v))
t=o
Hence, d(expp)o is the identity map of TpM. Applying the Inverse Mapping Theorem, we see that there is an a such that expp becomes 0 a C°° diffeomorphism of Be(0) onto expp(B.(0)). If we regard
as an open subset of IRm via the linear isomor-
phism TpM =R I , we see from Theorem 1.24 that the exponential map expp defines a coordinate neighborhood (expp(Be(0)), (expp I BE(0))-1)
about the point p. Namely, we can., using geodesics, introduce a natural local coordinate neighborhood system about each point p of a Riemannian manifold M. In general, if, given a neighborhood V of 0 E TM, the exponential map expp IV gives rise to a C°° diffeomorphism (expp IV)-1) or from V onto U = expp(V), we call the pair (U, simply U a normal coordinate neighborhood of p. Also, given a normal
1.4. GEODESICS
3T
coordinate neighborhood (U, 0), choosing, for example, an oarthonormal base lei,... , e,,,} for the tangent space TM, and identifying m
O(q)=>x'(q)eiE V, qEU, s=1
with (xl (q), ... , xm (q) ), we can define a local coordinate system (zt ).
We call it a normal coordinate system about p (relative to {ej). Theorem 1.24 can be refined as follows. THEOREM 1.25. At each point p in M, there exist a neighborhood
W and a positive number 6 > 0 such that given a point q E W, the exponential map expgat the point q is a C°° difeotnorphism from B (0) C TqM onto expq(Bs(0)) D W. Hence, W may be regarded as a normal Coordinate neighborhood about every point in W
PROOF. Given p E M, take e > 0 and the neighborhood V of p in Theorem 1.20. In the neighborhood U={(q,v)jgEV,vETTM, IvI<E} of (p, 0) in the tangent bundle TM, define a map F : U -+ M x M by F(q, v) = (q, expq v),
(q, v) EU.
Note that F(p, 0) = (p, p). As was seen in the proof of Theorem 1.24, d(expp)o = I; namely, the derivative d(expp)o at 0 E TM is the identity map. Hence, the derivative dF(p,o) : T(p,o)TM - TpM x TpM of F at (p, 0) E TM is an isomorphism. In fact, by the definition of F, it can be readily seen that the matrix representation of dF(p,o) is given by I
0
I I
By the Inverse Mapping Theorem, there exist a neighborhood U' C U of (p, 0) In TM and a neighborhood W' of (p, p) in M x M such that F is a C°° diffeomorphism from U' onto W'. We may assume that U' is given in the form
U'={(q,v) (qE V',
VETqM, Ivl
1. ARC M'NGTH OF CURVES AND GEODESICS
38
holds for q E W and BQ(0) C T.M. From the definition of F, this
0
implies expq(Ba(0)) D W.
1.5. Minimal length property of geodesics As was seen Example 1.21, the geodesics in the Euclidean space
E"' are nothing but straight lines. It is well known that a line in the Euclidean space Em has the property that it has minimal length among the curves joining any two points in the line. Conversely, this property characterizes the lines in Em. We see that the minimal length property locally holds for geodesics in a general R.ietuannian manifold M. As preparation, we note the following. Let (M, g) be an mr R.iennannian manifold and let u : O -r+ Al be a C°° map from an open subset 0 C Ra into Al. By generalizing the notion of the vector fields along a curve as defined in §1.3, a correspondence
V = {V(p)}moo that assigns to each point p E 0 a tangent vector V(p) E Tom ,) M of M at the point u (p) is called a vector field along the map u. Using the coordinate functions (x, yy) in R2 and a local coordinate system (x`) about u(x,y) in M, we can express V as m
(1.36)
V (X' y) =
v`(x, y)
( 8:ri
When each u' (x, y) is a C°° function of (x, y), V is called a C°° vector field along u. In terms of TM, a C°° vector field V along a C°° map u : 0 --* M simply means that V is a C°° map from O into TM such
that p E O, holds, where ar : TM - M is the bundle projection of TAI. Given a C00 vector field V along u, the restrictions of V to the curves u(x, ) and u(-, y), respectively, give rise to a C°° vector field along each curve. Hence, we can consider a covariant derivative of V along these curves. In order to distinguish these covariant derivative, it o V (p) = u(p),
we denote the covariant derivative of V along the curve u(x, ) by and the one along the curve u(-, y) by
83l
D
D , respectively.
For each point (xo, yo) E 0, the map defined by x
u (x, yo)
determines a COO curve u(-, yo) from each connected component of
0 n (y = Yo} into M. We denote by
the C°° vector field along
1.5. MINIMAL LENGTH PROPERTY OF GEODESICS
39
yo) defined by the tangent vector (1.37)
E TT(x,IVO)M
du(x,yo) ((....) (x,bo)
yo). Since the tangent vector
of this C0° curve
at any (x, y) E 0 from (1.37), we see that
c7u
(x, y) is defined
gives rise to a C°O
vector field along u. Similarly, au defined as the tangent vector field of the curve u(x, ) gives a C°° vector field along u. Then the following holds.
LEMMA 1.26. Given a C°° map u : 0 C R2 - M from an open subset 0 of R2 into M, we have
D au D 8u ay ax= ax ay PRooF. Given a local coordinate system (xi) in M, we express u(x, y) by
u(x+y) = (u'(x,y),... ,um(x,y)) By definition, we get
D%
m
k=1
a2uk y
m
k
ij=1
aus auj
ax
8 ax
Since the Levi-Civita connection is symmetric, ri? = r? . Hence, the
Dan = Data ay .
right hand side is symmetric in x, y; consequently, ay
8x
ax
When the exponential map expp : TpM -- M is a COO map in a neighborhood V of 0 E TM, the image B,(p) = expp B,(0) under expa, of the a neighborhood Be (0) C V of the origin 0 is called the geodesic ball of radius a about p in M. The image Si(p) = expp S((0) of the boundary
SS(0)={vIvETpM,
Iv(=e} of BE(0) is called the geodesic sphere of radius a of M at p. As is evident from the definition, Sf (p) is a submanifold (hypersurface) of M which is C°° diffeomorphic with the m -- 1-dimensional sphere Sm-1 C Rm.
1. ARC LENGTH OF CURVES AND GEODESICS
40
From the following Gauss lemma, we see that any geodesics emanating f mm a point p meet the geodesic sphere S, (p) perpendicularly.
L1A 1.27 (Gauss). Suppose that expp v is defined for a v E TpM. Then for each w E TpM T (TpM), we have (dexpp).(w)) = 9p(v, w)PROOF. First of all, note that the identification TpM Tn (TpM) is obtained by assigning each w E TpM to the tangent vector 4(0) E TV(T,M) of the line c (t) = v + tw. Where expp is defined, we set, for (t, s) E Ra,
u(t, s) = expp(t(v + sw)).
We obtain a C°° map u : O C Ra - M from an appropriate .ndghborhood 0 of the origin of R2 into M. FYom the definition, under the identification TQ (TpM) '" TM, we have (1,0) = (dexpp), (w).
(1,0) = (dexpp)q,(v),
Eksce, for the lemma to hold, it suffices to show that
r?u ou
at' as
9
(1,0) = 9p(V,w)
DSu
= 0 holds, since the curve t f--+ u(t, s) for each s is the geodesic whose initial vector is v + sw. Also by Proposition 1.17, we have of $,u
9
(o'
) (t°8) =9p(V +8W,V+sW).
On the other hand, since 0
D su = D Ou by Lemma 1.26, we get as at at 6.9
9(at' ) =9(a-t'ata ) =g(8t' 1a 8u au 2889
et' at
.
From this and the above equation, for any t, we have g
(' ') -5i ,TS
(t, 0) = g,(V, W).
)
I.S. MINIMAL LENGTH PROPERTY OF GEODESICS
41
On the other hand, since u(0, s) = expp(0) = p holds for any a, we have '5 T
Consequently, g
(8
,
, ) (0,0)=0.
8 (t, 0) = tgp(v, w) holds. When t = 1, this
0
equation gives the desired result.
Under the above preparations, we will see that geodesics are locally length minimizing curves. THEOREM 1.28. Let U be a normal coordinate neighborhood of p E M and Id B C U be a geodesic ball centered at p. Let c : [0,1] -. B
denote a geodesic in B emanating from p. Then for any pie smooth curve w : [0, 1] - M joining c(O) = 1 and c(1), L(c) < L(w) holds, and the equality holds if and only if c([O,1]) = w([0,1] ). PROOF. First, consider the case w((0,11) C B. For each t E (0, 11, we may assume that w(t) q& p. Otherwise, we can disregard the interval (0, t) and apply the arguments below.
Since expp is a COO diffeomorphism onto U, the curve w can be uniquely expressed as wp(t) = expp(r(t) v(t)), t 0 0. Here, v : (0,1) -* TpM is a piecewise smooth curve in S1(0) C TpM, and r : (0,1] -. R is a piecewise smooth, positive real valued function. Setting fir, t) = expp(rv(t)), we get w(t) = f (r(t), t), and except at a finite number of t,
w'(t) = Lr (r(t), t)r'(t) +
f
(r(t), t).
Since Lemma 1.27 and Iv(t) I = 1 imply 9
M,
8 tf
}(r,t)=0,
12 I(r,t)=1,
we get (1.38)
Iw(t)12
= Ir'(t)12 + 18t I (r(t), t) > Ir'(t)12.
Hence, (1.39)
I
1
I w'(t) I dt > j 1 Ir'(t)Idt >
f ' r'(t)dt = r(1) - r(e)
holds. Since r(1) = r(c), we get L(w) > L(c) by letting c -- 0.
1. ARC LENGTH OF CURVES AND GEODESICS
42
If L(w) = L(c), the equality in (1.38) and (1.39) holds. Hence,
for any t, 18f / ti{r(t), t) = 0 and fir, (t)I = r'(t) > 0 must hold. This implies that v(t) = 0, that is, v(t) is a constant vector, and w is nothing but a simple reparametriztion of c. Consequently, we get w([0,11)=c([0,IDWhen w([O, 11) is not contained in B, there is a parameter value
tj E (0,1) at which w intersects with the boundary of B. Thus, we have the inequality L(w) > L(wi[0,t11) > p > L(c), where p is the radius of B.
0
We note here that Theorem 1.28 is a local result, and that a geodesic may not always be length minimizing when its length gets larger. In fact, in the case of the geodesics in the unit sphere S'r` of Example 1.22, it is clear that a geodesic c emanating from a point p is no longer length minimizing when it goes beyond the antipodal point of p resulting in L(c) > ir. On the other hand, we see that a length minimizing piecewise smooth curve is a geodesic, as stated below. THE0R.EM 1.29. Let c : [a, b] - M be a piecewise smooth curve with a parametrization proportional to arc length. For any piecewise smooth curve w joining c(a) and c(b), if L(c) < L(w) holds, then c is a geodesic.
PROOF. Given t E [a, b), let W be the neighborhood of c(t) given in Theorem 1.25. We can choose a sufficiently small interval I C [a, b] with t E I so that c(I) C W and cII : I W represents a piecewise smooth curve joining two points in some geodesic ball. Then, by the assumption and the first half of Theorem 1.28, the length of cit equals the length of the geodesic joining these two points. Hence, we see that cuI is a geodesic from the fact that the parameter of clI is proportional to are length, combined with the second half of Theorem 1.28. Since 0 t E [a, b] is chosen arbitrarily, the entire c is a geodesic.
Given two points p, q in a Riemannian manifold M, the infimum of the lengths L(tv)'s of all piecewise smooth curves w's joining p and
q is expressed by d(p,q) and called the distance between p and q. Namely, the distance between p and q is defined to be d(p, q) = inf{L(w) I w is a piecewise smooth curve joining p and q}.
SUMMARY
43
If M is connected, the distance d(p, q) is well defined, since there are piecewise smooth curves joining p and q (Exercise 1.9 at the end of
this chapter). In this case, the function d : M x M -- JR, which assigns d(p, q) to (p, q) E M x M, gives rise to a distance function on M. In fact, by the definition, d(p, q) > 0 holds, and d(p, q) = d(q, p) is evident. Given p, q, r E M, the triangle inequality d(p,r) <_ d(p,9) + d(9, r)
also holds from the definition and the properties of infimum. Hence, we only have to show that d(p, q) = 0 if and only if p = q. To this end, since p = q clearly implies d(p, q) = 0, it suffices to show that p 0 q
implies d(p, q) > 0. Now assume that p # q and take the geodesic ball B,(p) about p. If q E B, (p), d(p, q) > 0 from Theorem 1.28. If q is not in B, (p), Theorem 1.28 also implies d(p, q) > E > 0, since any piecewise smooth curve joining p and q intersects with the geodesic sphere S, (p). In either case, d(p, q) > 0, if p # q. From the above arguments, we also see that for any p E M and a sufficiently small E, the geodesic ball BE (p) and the geodesic sphere SE (p) of radius E, respectively, are the sets defined, in terms of the distance d, as
B,(p)={gEMI d(p,q)<E}, S., (p) = {q E M I d(p,q) = e}. From these observations it follows that the metric topology defined
in M by the distance function d coincides with the topology of the differentiable manifold M (Exercise 1.10 at the end of this chapter).
Summary 1.1 The definitions of length L(c) and energy E(c) of a curve in a Riemannian manifold M. 1.2 The first variation formula and the Eider equation giving the critical points of variational problems. 1.3 The definition of the Levi-Civita connection V, i.e., a unique symmetric linear connection compatible with the Riemannian metric on a Riemannian manifold M. V defines the parallel transport of tangent vectors along curves in M. 1.4 The geodesics in a Riemannian manifold M give rise to the exponential map expp, which defines the normal coordinate system about p. 1.5 The geodesics are locally length minimizing curves in a Riemannian manifold M.
1. ARC LENGTH OF CURVES AND GEODESICS
44
Exercises 1.1 Given local coordinate systems (xi) = (xl,... , xm) and , rn) about a point x in an r-dimensional C°'° mani(X') = fold M, show the following: (1) Between the bases of the tangent space TM, transformation equations a (;-}
g
p-1 n
8 oxi
a
8 z
hold.
(2) Given the components (') and (x) of a tangent vector v E TAM with respect to (x1) and (x ), respectively, transformation equations
ax
_
(x)"
j=1
9=1
hold.
1.2 Prove that a C°° manifold M satisfying the second axiom of countability possesses Raemannian metrics. 1.3 (The fundamental lemma in the theory of variation) Let f : [a, b] -, R be a continuous function defined in a closed interval [a, b]. Assume that
f.
f (t)q(t)dt = 0
for any C°0 function i, : [a, b] - R. Prove, then, that f - 0. 1.4 Regarding a linear connection V on a C°° manifold M, prove the following.
(1)Let (x`)_(xl,...,xm)and (z)_(r1,...,Y') be two local coordinate systems about a point x E M. The connection coefficients {Pjk } and {I z, } with respect to the the coordinate systems satisfy the t ransorrmmation rule in
{rik}=E P=1
q.r=1
r''
OX9 ax,
I<&,j,k<m.
(2) Conversely, if a family of C°° functions {I Jk} is given to satisfy the transfiirmation rule (1) in each coordinate neighborhood,
EXERCISES
45
there exist a unique linear connection V on M whose connections coefficients are rfk. 1.5 (1) Given X, Y, Z E X(M) and f, h E C°°(M), show the following. (i) [f X, hY] = fh[X, YJ + f(Xh)Y h(Y f )X.
-
(ii) [X,[Y,ZJ]+[Y,[Z,X]]+[Z,[X,Y]] =0. (2) Given X, Y E X (M), set .LxY = [X, YJ and call it the Lie derivative of Y by X. What is the difference between the Lie derivative LX Y and the covariant derivative V X Y?
1.6
Regarding a linear connection V on a C°O manifold M,
prove the following. Given v E T_- M, choose a C'° curve c : [0, lJ - M
such that c(O) = x and c'(0) = v. Express the parallel transport by Pt : TxM T t)M along cj[0, t] for t E [0, lj. Then the covariant derivative V,,Y of Y E X(M) in the direction of v satisfies IY(it) - Y=). t-o 1(Pt t
V21Y = lim
1.7 Given the tangent bundle TM of a Riemannian manifold M, prove that there exists a unique C° vector field + in TM satisfying the following condition. The integral curve ap(t) of $ passing through (p, v) is given by cp(t) _ (c(t), c'(t)), where c(t) is the geodesic such that c(o) = p and c'(0) = v. This vector field % is called the geodesic spray. The local one parameter group of transformations Ot generated by ' is called the geodesic flow on M. 1.8 Let M be a Riemannian manifold and let (x1, ... , xm) be a local coordinate system around a point p E M. Prove the following. (1) A necessary and sufficient condition for (x') to be a normal coordinate system is that any geodesic passing through p is expressed by linear equations
ci(t)=vet,
f
(2) If (x') is a normal coordinate system, the equations
f (p)=0, Y
CHAPTER 2
First and Second Variation Formulas In Chapter 1, we computed the first variation formula for the energy functional of a curve in a Riemannian manifold when the image
of the curve was contained in a coordinate neighborhood. We also showed that a geodesic was characterized to be a map given as a critical point of the energy functional. In this chapter, we compute the first variation formula for the general case, and, obtain the second variation formula. As the concept of covariant differentiation was naturally induced from. the first variation formula, we see that the second variation formula naturally relates to the notion of curvature. Making use of this fact, we can investigate the effects of the curvature of a Riemannian manifold on its topological structures.
2.1. The first variation formula Let (M, g) be a Riemannian manifold of dimension m, and let c : [0, a] - M be a piecewise smooth curve of M defined in a closed interval [0, a] (a > 0). In §1.2, we computed, regarding the problem of variation, the first variation formula for the length and energy functionals when c was a C curve whose image was contained in a coordinate neighborhood of M. In this section, using the covariant differentiation defined in §1.3, we compute it for a more general piecewise smooth curve c, when the image of c is not necessarily contained in any single coordinate neighborhood.
First of all, in order to compare the length or energy of curves, we define a family of curves as follows.
DEFINITION 2.1. Given a piecewise smooth curve c : [0, a] - M,
a continuous function f : [0.a] x (-E, E) ---+ M (E > 0) is called a piecewise smooth variation of c if it satisfies the following conditions: (i) f (t, 0) = c(t) t E [0, a]. 47
2. FIRST AND SECOND VARIATION FORMULAS
48
(ii) There is a partition of the interval 0 = to < tl < < tt+l = a such that f is C°° in each domain of definition (ti, ti+1] x (-Elf)?
i=0,...,k.
(iii) f (0, s) = c(0), f (a, s) = c(a), s E (-E, e). In particular, if f is a COO map, f is called a C°° variation or a smooth variation. (In this case, by definition, c is a C°° curve.) Let f : [0.a] x (-e, c) -- M be a piecewise smooth variation. For each s E (-E.f), we set f, (t) = f (t, s), t E [0, a]. From (ii), fs : [0, a] - M becomes a piecewise smooth curve. From (i) follows fo = c. From (iii), we see fs(0) = c(0) and f. (a) = c(a). Consequently, we may consider that the piecewise family if, I !e1 < e}
thus obtained defines a "deformation" of the curve c, keeping both terminal points c(0), c(a) fixed.
C(0)
FIGURE 2.1. A piecewise smooth variation Also for each t E [0, a], if we set
It(s) = f (t, s),
s E (-E, E),
we get, keeping s = 0, C°° curves ft : (-e, e) - M passing through the points c(t). The set of all the tangent vectors to these curves at s = 0, denoted by V(t) = of (t, 0),
t E [0, a],
defines a piecewise smooth vector field V along the curve c. This V is called a variational vector field associated with f. From condition (iii), it is evident that V(O) = V(a) = 0. Conversely, the following holds.
[0, a] - M be a piecaewise smooth curve, and let V be a piecewise smooth vector field along c such PROPOSITION 2.2. Let c
:
2.1. THE FIRST VARIATION FORMULA
49
that V(O) = V(a) = 0. Then there is a piecewise smooth variation
f : [0, a] x (-c, c) --e M of c so that V is the variation vwtor field associated with f.
PROOF. Since the image c([0, a]) of c is a compact subset of M,
there is a positive number S > 0 such that the exponential map exp,(t) is defined at any t E [0, a] for any tangent vector v E Tc{t)M with [vj < b. In fact, if we take the normal coordinate neighborhood Wt and a positive number St > 0 at each c(t) as shown in Theorem 1.25, we get an open cover { Wt }, t E [0, a] of the compact subset c([0, a] ). We choose a finite subcover {W;}, i = 1,... ,1. Then the desired 6 is obtained by setting d = min (St , ... , 81). Define N = maxtE IO,aI I V (t) j, and choose a positive number a in such a way that e < 6/N is satisfied. Then the desired map [0, a] x M is defined by (-C, C) f (t, s) = exp(c(t), sV(t)) = expCt) sV(t).
f (O, s) = c(0), f (a, s) = c(a) hold, since clearly V(O) = V(a) = 0 from the definition. This f is the desired variation. In fact, from the definition of the exponential map, exp(c(t), 8V(t)) depends upon (c(t),sV(t)) in a C°° manner (see §1.4); hence, f defines a piecewise smooth variation. The property of the derivative of the exponential map shows that the variation vector field associated with f is given by
of (t, 0) _
--(exp,(t)
sV(t))Is_o = (dexpc(t))oV(t) = V(t).
0 Given a piecewise smooth variation f : [0, a] x (-E, e) -' M of a piecewise smooth curve c : [0, a] - M, define functions L : (-e, e) --
kandE:(-e,E)-+Q8by k
(2.1)
ro I(t,8)k = E i=0 ,/t
k+t 1
,
lot
(t, 8) Ia,
k
E(s) _ 2 o
f(t,s)l2dt
I(t,8)I2dt.
By the definition, L and E, respectively, are nothing but the functions
that assign to each s E (-e, e) the length L(ff) and the energy E(f,) of the curve f, obtained from the deformation of the curve c. By condition (ii) of the variation, L and E are C°° functions.
so
2. FIRST AND SECOND VARIATION FORMULAS
At this point we point out the relationship that holds between the
length and the energy of a curve. In general, for a given C' curve c : [a, b] -+ M, the length and the energy of c are defined, respectively, by
L(c) = a
E(c) =
Ic(t)'Idt,
'J
b
Ic(t)'12dt.
2 a
By setting h1(t) = 1 and h2(t) = Ic'(t)I in the Schwarz inequality 2
b
hih2d a
b
<
h12dt a
J h22dt a
we get an inequality L(c)2 < 2(b - a)E(c). Here, the equality holds if and only if h2 is proportional to h1 = 1;
namely, the parameter t of the curve c is proportional to the are length. Even if c is piecewise smooth, we see that a similar result holds by decomposing c into a union of C°° curves. From the above inequality, we get the following lemma regarding minimality of the length and energy of curves. LEMMA 2.3. Given p, q E M, denote by c : [0, a] - M a geodesic joining p and q. Then given an arbitrary piecewise smooth curve w : [0, a] - M joining p and q, if L(c) < L(w) holds,
E(c) < E(w) also holds. The equality holds if and only if w is also a geodesic of the minimal length.
PROOF. Since c is a geodesic of the minimal length, the above inequality implies 2aE(c) = L(c)2 < L(w)2 < 2aE(w). Hence, we get E(c) < E(au). Since the equality Here clearly implies L(W)2 = 2aE(c), we see that the parameter of w is proportional to arc length. On the other hand, since L(c) = L(w) holds, from Theorem 1.29, we see that w is a geodesic of the minimal length. The converse is self-evident.
0
Fiom Lemma 2.3, an energy minimizing curve c has a parametrixation proportional to its arc length. We see from this that it is more
convenient to use E as a functional than L directly, even when we study the variational problems regarding the arc length of curves. We,
2.1. THE FIRST VARIATION FORMULA
61
therefore, compute the first variation formula of the energy functional E. THEOREM 2.4 (The first variational formula). Let c : [0, aj -* M be a piecewise smooth curve, and let f : [0, a] x (-E, c) --+ M be a piecewise variation of c. Then regarding the function E : R that assigns to each curve its energy, the following holds:
E'(0) =
ra
- Ja
g (V,
Dc) dt
k
dDc' )) = _g (V(t,), Dc' (ti) t (ts d
.
i=1
Here, V is a variational vector field of f, and
Dc'
(tt ),
Dc'
(t_
) are,
respectively, given by
Dc' dt (t`)
dDc' t
t>tt
t
dt (t)
(t)
tt>ti
alt
W.
PROOF. Using the definition (2.1), we compute
dE
_
1
of Of
d
k
w 2 - =ads j
(rat ' 8t )
9
.
In each domain of the definition [t,.4+13 x (-E,E), Of 10t and Of/Os define C°° vector fields along the curves f,,, ft, respectively. By Proposition 1.15 and Lemma 1.26, we have d da9
8f 8f rat
' at
-
2g
Dof If
D Of _ D Of
as 8t '
as et
at)
cat as
Hence, in each [tj.t,+1], we get d
te+l
ds it"
(,
of
at 8t }
rt'+,
=J 2
2g
tj+1
D
Of of
a N, 5i) dt D
8f Of 50
Noting that Proposition 1.15 implies d
D Of
dt9(a'cat) -9(atof' cat)+gas'3t5t)'
2. FIRM AND SECOND VARIATION FORMULAS
52
integration by parts gives us, in each [t;, 4. .F1],
(D Of Of dt
atas'at) of of u+
9
as' at
j
Consequently, we get
(2.3)
=
eI of k
g f Daf
c,+1
-
g i
as at at } dt.
nof Daf , J9ru)dt. ^
Setting s = 0 in (2.3) and noting c(t) = f(t, 0), V(t) = 8s (t, 0) and V(0) = V (a) = 0, we get (2.2) from the definitions ofd
(tt), (t ).
From the first variation formula in Theorem 2.4 follows COROLLARY 2.5. A necessary and sufficient condition for a piece-
wise smooth cvroe c : [0, a] -+ M to be a geodesic is E'(0) = 0 for any pieoewise smooth variation f of c.
PRooF. If c is a geodesic in M, from the definition, de = 0 and
(tt) =
(t{ ), 1 < i < k; hence, necessity is obvious. In order to see sufllciency, we assume that E'(0) = 0 holds for any piecewise smooth variation f . Choose a piecewise smooth function
dt
h:[0,a]-'JRin each tt(0
0 fort 96 to hold. Let Dc' dt be a piecewise smooth vector field along c. Given a piecewise smooth variation with V as its variation vector field (see Proposition 2.2), we have from (2.2)
V(t) = h(t)
E'(0)=-
12dt
ah(t) I Jo
= 0.
Hence, we see that Dc'/dt = 0 in each open interval (t{, tf+i ). Since c is a piecewise smooth curve, Dc'/dt = 0 in [ti, ti+i]. Consequently, c I [ti, ti+i] is a geodesic.
2.1. THE FIRST VARIATION FORMULA
53
Next, consider a variation V such that V (O) = V (a) = 0 and in
each ti (1 Si
d (t') - dt (tt )
Since each c ( [t;, t,+1] is a geodesic and (2.2) implies
E'(0) _
-E 1=1
I(t)
2
dt (ti )
I
= 0,
we see that c is cl in each t;. On the other hand, since Dc'/dt = 0 in each [t;, ti+11, we also see that c satisfies the equation of geodesics. By the uniqueness of the solutions to the geodesics equation, we see that c is c0° and a geodesic in [0, b a1.
Corollary 2.5 implies that a geodesic in a Riemannian manifold, in general, is characterized as a critical point of the variational problem regarding the energy E of the curves. REMARK. Given two points p and q in a Riemannian manifold M, if we consider the set fl(M; p, q) of the piecewise smooth curves joining
p and q as a "manifold", the energy functional E : 1Z(M; p, q) -' 1[2 can be regarded as a "function" defined in this manifold. Also given c E SZ(M; p, q), we may consider a piecewise smooth variation f = {fe 151 < E} along c and the set of the variational vector fields
V of f as defining a "curve" in 1(M; p, q) such that it passes the point c = fo at s = 0 and the "tangent space" of 1Z(M; p, q) at c, respectively. In this setting, since we may consider the first variation formula as defining the "derivative" of the function E in the direction V, we can characterize a geodesic joining p and q as a critical point of this function.
In general, the tangent space, namely, the set of all piecewise smooth vector fields along c, of fl(M; p, q) at c is not finite dimensional; hence, in order to regard SZ(M; p, q) as a manifold, we need the notion of "infinite dimensional manifolds". In fact, we can introduce the structure of an infinite dimensional manifold modeled on a Hilbert space in an appropriate completion of 1Z(M; p, q). We may, in that setting, discuss the variational problem of the energy functional E from the view point of "Morse theory". On the development of Morse theory on infinite dimensional manifolds, one may consult, for instance, Palais [17], Smale [211, Palais and Smale [18], Schwarz [221 or Nagano [301 (see Prospects for Contemporary Mathematics starting on page 183).
2. FIRST AND SECOND VARIATION FORMULAS
54
2.2. Curvature tensor In this section, we discuss the curvature tensor and various kinds of curvatures in Riemannian manifolds, as preparations for the second variation formula of the variational problem regarding the energy of curves. Let us assume, for the time being, M to be simply a C°'O differentiable manifold of dimension m. First, we define a tensor field on M as a generalization of a vector field. Given x E M, let T=M and TTM* denote the tangent space of M and its dual space at x, respectively. In §1.1 we defined the tensor product TM* ® T. ,M* as the vector space over 1C consisting of all the bilinear forms. Generalizing this definition, we define the tensor product
x . xTxM* _R r times
J0 times
as the set of all (s + r) linear maps
T",(x) = TM* ®... ®TTM" ®TxM ®... ®TxM r times
s times
defined in the direct product of s copies of Tx M and r copies of TT M* .
Each element of T(x) is called a type (r, s) tensor over TIM. As can be seen readily, T(x) gives rise to a vector space over R under the naturally defined addition and scalar product. In particular, when r = 1, a = 0, T°(x) is identified with TIM, and when r = 0, s = 1, T; (x) is identified with TTM*. Let (x*) be a local coordinate system about x. `Then the bases
of TM and T2M* are given, respectively, by
(
11
i
and (dri)z.
J
x
Hence, an element T E T(x) is uniquely determined by the values
T(( (2.4)
,
),... ,[ T
We express these values by
`
Ee
)
(dx )I
,(tom''
x
15 ii,
.
,
ts,ji, ... J'< M.
and call them the components of
T(x) in (x'). Now let
K=(dx=')a®...®(dz'P)x0(
1!
®...®(Jr/
2.2. CURVATURE TENSOR
55
be a tensor of type (r, s) over TM defined by, for v1, ... , v8 E TAM,
wi,...,wr E Tx , K(v1,... ,us,w1,... ,Wr)
=dx (v1)...dx (v )w1 ((&71l
...wr
Ix
((ir))\\ x
Then it is easily seen that these rnr+s number of tensors form a basis for TT(x). Hence, the dimension of TT(x) is mr+e Using the com-
ponents T's given in (2.4), any T E Tr (x) can be written in the form of
T=
, jl,
to
.. ®(dx'- )x
(di-'t )x
,7r
a axe
a
}x
-
axir
x
Let U be an open subset of M. A correspondence T = {T (x) }ZEV
which assigns to each point x E U a type (r, s) tensor T(x) E T; (x) is called a type (r, s) tensor field of M in U. In particular, if M = U, T is simply called a type (r, s) tensor field of M. Let (x') be a local coordinate system about a point x E U. When the components Ti= I.-?" .is (1 < il, ... > ig, jl, - - , j, <_ m) of T (x) E TT(x) represented as in (2.5) are always C°° functions in (xt), we call it a type (r, s) C°D tensor field of M in U. For example, the Riemannian metric on a C°° differential manifold M given in Definition 1.1 is nothing but a type (0, 2) C°° tensor field of M. Given an open subset U in M, we denote by T' (U) the set of all type (r, s) C°° tensor fields of M in U. By the definition, we can identify T' (U) the set of all C°° vector fields of M in U. Also as in the case of 1(U), for given T, S E V, (U) and f E Coo (U), we, respectively, define the sum S + T of S and T and the product of f and T as follows:
(fT)(x) = f (x)T(x), x E U. (T + S)(x) = T(x) + S(x), It is readily verified that T(U) is a C°°(U) module under the above operations. PROPOSITION 2.6. A type (0, s) tensor field T E TO(M) is in one- to-one correspondence with an s linear map T : 1(M) x . . x 1(M) - C°°(M) satisfying, for any ft E COO(M) and Xt E 1(M), (2.6)
T(f1X I, ... , f3Xs) = ff ... f8T(Xl,... , Xq).
2. FIRST AND SECOND VARIATION FORMULAS
56
Also, a type (1, s) tensor field T E T' (M) corresponds, in one-to-one fashion, to an s linear map T :1(M) x . x 1(M) -+ 1(M) satisfying (2.6).
PROOF. We begin by showing the first half. For a given type (D, s) tensor field T = {T(x)}XEM and X1, ... , X. E 1(M), set
T(X1i... , X8)(x) = T(x)((X1),,, ... , (X,)x),
x E M.
Then T (X1i ... , X8) defines a C0° function in M. In fact, if we express
Xa, using a local coordinate system (xi) of M, as m
X{=
X
k=1
1
and denote by Ti1,..i,'s the components of T(x) in (xi), we have m
T(X1i... ,X,)(x) =
C11(x... ,b
(x)7'i1...is(x);
implying that T(X1, ... , X8) is a C°° function of (xi). Furthermore, the map
T : 1(M) x
x 1(M) , C°°(M)
obtained as above clearly satisfies (2.6) from the s linearity of T(x).
Conversely, let T : 1(M) x .. x 1(M) - C°°(M) be an s linear map satisfying (2.6). Then the following Lemma 2.7 implies that for given X1i ... , X8 E 1(M), the value T(X1, ... , X8)(x) of
T(X1i... , X8) at a point x E M is completely determined by the values of (X1)., ... , (X8)., alone. Hence, if we choose X1, ... , X8 E 1(M) such that (X1) ., = vi for given v1, ... , v8 E TM and set
T(x)(v1,... ,u8) =T(X1i... ,X,)(x), we get an s linear map
T(x) :TM x
xTzM -}fit.
This defines the desired type (0, s) tensor field T = {T(x)},,EM of M. That this T is a CO° tensor field can readily be verified from the definition by choosing the basic elements { (8/8x ),,} as the above vi.
The second half can be shown in a similar manner to the first half, if we note that a type (1, s) tensor T(x) E T' (M) over TM corresponds to an s linear map T(x) :TAM x . x TZM --+ TxM, as given in Exercise 2.2 at the end of this chapter.
2.2. CURVATURE TENSOR
5T
LEMMA 2.7. Let T : X(M) x ... x X(M) - C°°(M) or T : X(M) x ... x X(M) -- X(M) be an s linear map satisfying (2.6) and Let x E M. Then given X1, ... , X8, Y1, ... , Y8 E X(M), q X1 (x) = Y(x), (1 < i < s), then
T(X1,... ,X,)(x) = T(Y1,... Y,) (x) holds.
PROOF. In either cam, because of the s linearity of T, it suffices
to show T(X1, ... , X1, ... , X8)(x) = 0 for some i with X{(x) = 0. For each term on the right hand side of s
T(Y1,...,xi -Y,... ,Xs)
T(X1,... ,X5) -T(Y1,... ,Y8) =1 becomes 0 at x.
In what follows, we assume i = 1 for the sake of simplicity. First, we verify that if X1 = 0 in an open subset U of M, then
T(Xl,... , X,8) = 0 in U. As was given in the proof of Lemma 1.8, given x E U, choose an f E C°°(M) with f(x) = 0 and f . 0 outside U. In this case, since fX1 .= X1, (2.6) implies at x that T(X1, .. , X.)(x) = f (x)T(X1,... , Xs)(x) = 0.
X8) = 0 in Since x is an arbitrary point in U, we have T the entire U. Next, we verify that T (X 1 i ... , X8) (x) = 0 if XI(x) = 0. Let (x1) be a local coordinate system about x and let X1 = FM 1 t k8/8xk. For k = 1, ... , m, choose Ek E X(M) and fk E C°°(M) in such a way that Ek = d/8xk and f k = t;k in a sufficiently small neighborhood U of x. Define X t E X (M) by X' = Fm t f k Ek. Then fk(O) = t k = 0, since X1 = Xi in U and X (x) = 0 from the definition. Consequently, we have T(X1i ... , Xa)(x) T(XI, ... , Xe) in U. From (2.6), we get. at x M
T(X1, ... , X8)(x) = > fk(x)T (Ek, ... , Xs)(x) = 0. k=1
Under the preparations above, we define the curvature tensor of a Riemannian manifold. In what follows, we assume that M is the Riemannian manifold (M, g) of dimension m and V denote the LeviCivita connection of M.
2. FIRST AND SECOND VARIATION FORMULAS
58
DEFINITION 2.8. Given X, Y E X(M), we set
R(X, Y) = V xV y - V yV x - Vjx,vx].
We define a map R:X(M)xX(M)xX(M)-+ X(M) by R(X, Y, Z) = R(X, Y)Z,
X, Y, Z E X(M).
This R is called the curvature tensor of M or (M, g).
For example, if M is the rn-dimensional Euclidean space El, R(X, Y, Z) = 0 for any X, Y, Z E X(M) - In fact, when Z E X (M) is regarded as an Rm valued C°° function Z = (f 1, ... , f'n) in M, we have
V VyZ = (XYf',... ,XYr). Hence from the definition of [X, Y], we get
R(X, Y) = VxVy - VyVX - V[x,yJ = 0. Therefore, we may consider the curvature tensor R representing a quantity that measures the degrees of deviation of M from the Euclidean space El. As we can also see from the following lemma and Proposition 2.6,
R is a type (1,3) C°O tensor field over M. This is the reason why R is called the curvature tensor of M.
LEMMA 2.9. R : X(M) x X(M) x X(M) -+ X(M) is a trilinear map, and for any X, Y, Z E X(M) and fl, f2, f3 E C' (M), we. have R(f1X, f2 Y, f3 Z) = f1f2f3R(X, Y, Z)
PROOF. It is obvious from the definition that R is linear with respect to the sum and the scalar product in each variable. As for the property regarding the product with a CO° function, say f3s we have
R(X,Y,f3Z) = R(X,Y)(f3Z) = VX VY(.f3Z) - DYVX(f3Z) - VIX,YJ(f3Z)
= f3(VXVYZ - DyV Z - D[X.y]Z) + (XYf3)Z - (YX f3)Z - ([X, Y]f3)Z = f3R(X, Y, Z).
Similarly, we can verify the above for fl, f2
0
The curvature tensor R satisfies the first Bianchi identity given as follows:
2.2. CURVATURE TENSOR
59
PROPOSITION 2.10. For any X, Y, Z E 1(M), R(X, Y, Z) + R(Y, Z, X) + R(Z, X, Y) = 0.
PROOF. From the symmetric property of the Levi-Civita connection, the left hand side equals
VXVyZ - VyVXZ - V{X,y)Z +VyVZX - VzVYX - v(y,z]X + VzVXY - VXVZY - V[Z,X1Y = Vx[Y, ZJ - V1y,z] X + Vy[Z, X] - V 1Z,xjY +VZIX,YJ - V[X,y1Z
= [X,fY,Z]]+[Y,[Z,X]]+[Z,[X,Y]]. Hence, we obtain the desired conclusion from the Jacobi identity for the vector fields.
R is a C°° tensor field of type (1,3). Set
R(X,Y,Z,W) =g(R(X,Y)Z,W), X , Y,Z,W E 1(M). This defines a map
R : 1(M) x 1(M) x 1(M) x 1(M)
C°°(M). We readily see from Proposition 2.6 that R is a C°° tensor field of type (4,0). This R is called the Riemannian curvature tensor of M or (M, g). We have the following with respect to R. PROPOSITION 2.11. For any X, Y, Z, W E 1(M), we have (i)
R(X,Y,Z,W) = -R(Y,X,Z,W),
(ii)
R(X, Y, Z, W) = -R(X, Y, W, Z), R(X, Y, Z, W) = R(Z, W, X, Y).
(iii)
PROOF. (i) follows readily from R(X, Y) = -R(Y, X). (ii) is equivalent to R(X, Y, V, V) = 0, where V = Z + W. Since the LeviCivita connection is compatible with V and g, we see R(X,Y, V, V) = g(VxVYV,V) - g(VyV V, V) - g(V(x,YJV V) = 2 {XYg(V, V) - YXg(V, V) - [X, Y]g(V, V)} = 0.
As for (iii), the first Bianchi identity implies that R(X, Y, Z, W) + R(Y, Z, X, W) + (Z, X, Y, W) = 0.
2_ FIRST AND SECOND VARIATION FORMULAS
60
By replacing X, Y, Z, W in this equation in the order X -., Y --, Z W -+ X, we obtain a set of four equations. Adding those equations together with (i) and (ii), we get
2R(Z,X,Y,W)+2R(W,Y,Z,X) = 0. Hence, R(Z,X,Y,W) = R(Y,W,Z,X) holds. Given a local coordinate system (z') in a local coordinate neigh-
borhood U in M, the curvature tensor of M and the Riemannian curvature tensor, in terms of the components in (x'), are expressed, respectively, as follows: in Rajkdx` (& d. T'
R
dal
dx 0dx'` 0dzl.
R= In U, we have
R
8
a
a
'"
E Milk !=1
11a a
a
8
Rtjkt = 9 R 8xs , axj } axk 8xl
jia
_ E X,011 . r=1
Flurthennore, it is readily verified from the definition (see Exercise 2.4 at the e n d of t h i s c h a p t e r ) that the components R , 's of the curvature tensor are given by
a
m
8
irrjk - Tiyr ik) r-1 The equations in Propositions 2.10 and 2.11 are, respectively, given as (2.7)
jk -
Mjk = axi U
I'iik +
Rjk+Rpk,+Rij=O, Rijkt = -RjkU, Ri,ki = -Rijlk, Rijkt = Rktij. RJ 1A.nx. Since some authors define the curvature tensor R as
R(X, Y)Z = VyVxZ - VxVYZ + V(x,y1Z, or express a
s
R (axi' x, }
a axk =
rn
1 l=1
a
2.2. CURVATURE TENSOR
61
it is imperative to pay attention to the positions of the subscripts and superscripts in symbols and component representations.
For a given x E M, denote by a a two-dimensional subspace of the tangent space TIM. Then given an orthonormal basis {v, w} for or with respect to gI, the value of K(v, w) = R(x)(v, w, w, v) = 9x(R(x)(v, w) w, v)
is determined, independent of the choice of {v, w}, only by a. In fact, let {v', w'} be another orthonormal basis for a. Then {v', w'}, in terms of {v, w}, can be expressed as v` = cos OV + sin 9w,
w' _
sin 9v ± cos 9w.
From the property of R in Proposition 2.11, we readily see that R(x)(v, w, w, v) = R(x) (v', w', w', v'). We denote this value by K(a) and call it the sectional curvatum of or. From the definition, the sectional curvature K(a) is determined by the curvature tensor R. Conversely, the curvature tensor R(x) at a point x E M is completely determined, if the sectional curvatures of all two-dimensional subspaces a C Tx M are given (see Exercise 2.6 at the end of this chapter).
EXAMPLE 2.12. Let M be the in-dimensional Euclidean space Em. Since the curvature tensor R - 0, all the sectional curvatures are 0.
EXAMPLE 2.13. Let M be the unit sphere Sm C E'+1 with the induced metric g from E11+1. In terms of the local coordinate system
defined in Example 1.22, the components g,, and r t of g and the connection V are, respectively, given as
gcl=aci+(
xixj +1)2,
rks,=xk gij,
1
Hence, a simple computation, using equation (2.7), yields
R4jkt=g.a9,,k-9119ik, 1
(H'", g). From the result of the computation in Example 1.23 and equation (2.7), we get
Rjkt = -(9iIgjk - 9jcgik),
1 < i, j, k,1 < 'rra,
2. FIRST AND SECOND VARIATION FORMULAS
62
from which we see the sectional curvatures of (H"', g) are always -1.
As seen above, a Riemannian manifold (M, g) whose sectional curvature K(o) is constant, independent of the two-dimensional subspace a C TsM is called a space of constant curvattrn Given z E M and v, w E TxM, the curvature tensor R gives rise to a linear transformation of the tangent space TIM, TsM 3 t '--> R(x)(t, v)w E TxM.
The trace of this transformation for each (v, w) E TxM x TxM given by m
trace{t
R(x) (t, v)w} _
9x(R(x)(ei, v)w, es) i=1
defines a bilinear form on TIM. Here, {et} is an orthonormal basis for TxM with respect to gx. As is well known, the value of the trace does not depend on the choice of the basis. The value of this trace is denoted by Ric(x)(v, w), and the type (0, 2) tensor field Ric = {Ric(x)}$EM of M defined by the bilinear form Ric(x) E :,x* x TxM* on each TM is called the Ricci tensor of (M, g). We note that the Ricci tensor is a symmetric tensor field of type (0, 2). In fact, from Propositions 2.10 and 2.11, we get m
Ric(x) (v, w)
_
9x(R(x)(ei, v)w, e1) i=1 M
= >{9x(R(x)(ei, w)v, eY) - 9x(R(x)(v, w)et, ei)} i=1
= Ric(x) (w, v).
If we choose an orthonormal basis lei} such that v = em for each unit vector v E TxM, we get
m-l
m-1
Ric(x)(v,v) _
g,, (R(x)(ei,v)v,ej) i=1
K(e-i,v). i=1
Hence, this implies that R(x) (v, v) is nothing but the sum of the sectional curvatures K(v, ej) over the orthonormal basis { ei } (1 < i < m - 1) for the the orthogonal complement of v. This quantity is called the Ricci curvature of M in the direction of v.
2.2. CURVATURE TENSOR
63
Furthermore, the trace S(x) of the Ricci tensor Ric given by
S(x) = trace{v
Rie(x)(ei, ei)
Ric(x) (v, v)} _ i=1
is called the scalar curvature of M at x E M. S(x) is nothing but the sum of the Ricci curvature in the direction of each e; in an orthonormal basis {e,} for T,,M. The C°O function S determined by the scalar curvature S(x) at each point in M is simply called the scalar curvature of M.
Let (x') be a local coordinate system of M in a local
rdi-
nate neighborhood U. if we express the Ricd tensor of M using the components in (x') as in
Ric = E Rjdz' 0 dx', i j=1
then we have in U m
Rij
! k=1
Rk kij
m
kl 9 R
k,l=1
As observed above, although the curvature tensor R of a Riomanifold (M, g) was defined for vector fields using the LeviCivita connection of M, it is actually defined as a tensor field of M; hence, its values are basically determined for the tangent vectors..Furthermore, as is seen readily from equation (2.7), the curvature tensor R and its derivatives, such as the sectional curvature, the Ricci tensor Ric and the scalar curvature S, are "geometric quantities" determined
only by the Riemannian metric g on M (see Exercise 2.8 at the end of this chapter). Finally, we will reconstitute the definition of tensor fields from the viewpoint of vector bundles. In a similar fashion to the tangent bundle seen in §1.4, given an m-dimensional C"O manifold M, the totality
TM` = U,,,TpM* = {(P, U;) I P E M, W E TPM*} of the dual spaces TpM* of the tangent space TpM over p E M possesses a structure of a 2m-dimensional C°° manifold. The projection it : TM* -- M defined by 74P, W) = P,
(p, w) E TM",
64
2. FIRST AND SECOND VARIATION FORMULAS
is a C°° map from TM* onto M. This set TM` is called the cotangent bundle of M. If, using a local coordinate system (x') of M in a local coordinate neighborhood U, we write w E TM* at q E U as m
w
(dx')q, £=1
a local coordinate system of TM' in TM* I U is given by the 2mtuple of functions It is readily verified that TM` naturally becomes a 2m-dimensional C°° manifold under these local coordinate
neighborhood systems and that the projection r is a C°° map. More generally, if we consider the totality
?a =U MTr(p)=t(p,T) IpEM, TETe(p)) of the type (r, s) tensors in TTM at each point p E M, Te also naturally carries the structure of an m + m''+'-dimensional C°° manifold. The projection Tr : T; - M defined by
r(p,T) = p, (p. T) E Tg (M), gives rise to a C"'° map from Te onto M.T. is called the type (r, s) tensor bundle of M. Let (x`) be a local coordinate system of M in a local coordinate neighborhood U about p. If we express T E Ts (M) in the form given
in (2.5), then the tuple (xii Tii.. a) consisting of x' and the m''+" components T12...$. gives rise to a local coordinate system in 71.(M) U = W-1(U). It can be shown in a similar manner to the cases of TM
and TM" that ?; (M) naturally becomes a CO manifold and that the projection r is a C°° map. For the tensor bundle TT(M) of type (r,s) over M, r-I(p) T,'(p) is called the fiber over p. A C°O maps : M - Ts (M) with the property
ar o s(p) = p,
pEM
is called a C°° section of 7,(M). The set of all CO° sections of TS (M) is denoted by r(TT (M)). Then, given 81, 82 E r(Ta (M)) and f E C( M), we define a sum 81 + 32 of 81 and s2 and a product f al off and s l, respectively, by
(Si +S2)(p) = si(p) + S2(p) (fsl)(Y) = f(j)s1(p), It is readily seen that I'(7 (M)) is a C°°(M) module.
p E M-
2.3. THE SECOND VARIATION FORMULA
66
over M is, by the definiA type (r, a) tensor field T = (T(p)) tion, a correspondence that assigns to each point p E M a type (r, s) of T defined by tensor T(p) E T8 (p) such that the components Ti' (2.4) are C°0 functions in each coordinate system (x=) in M. Hence, as seen readily from the definition of the local coordinate systems in T,' (M), T, in fact, gives rise to a C°O section
T:M-3 p
T(p)ET (M).
Conversely, a C°° section of T, (M) is a type (r, s) C0° tensor field over M as clearly seen from the definition. Thus, we can identify r(T, (M))
with the set 1. (M) of all type (r, s) C°° tensor fields aver M. Fbr the tangent bundle TM, for example, r(TM) is, by the definition, nothing but the set X(M) of the C°° vector fields in M; hence, it can be identified with T' (M). As we have seen above, by introducing the notions of the tensor bundle 1r(M) and the space F(7y (M)) of its C"'° sections, we can not only make the definitions of vector fields or tensor fields more succinct, but also we can treat them in a unified way.
2.3. The second variation formula Under the preparations in the previous sections, we obtain the second variation formula for the variational problem on the energy of curves in this section. First of all, we note the following. Let (M, g) be an m dimensional
Riemannian manifold. As in §1.5, let u : 0 M be a C°° map from an open subset O C R2 of R2 into M. Let V be a C°° vector field along u. By restricting V to the curves u(x, ) and y), we get a C°O vector field along these curves. Hence, for V, we may consider -V along the curves. On the other covariant derivatives and
hand, for the curvature tensor R of M, R ( au, "u) V is naturally defined as a C°O vector field along u.
LF iMA 2.15. Given a C°° map u : O C R2 --> M from an open subset O of R2 into M and a C°° vector field V along u, we have
DV=R(5,i) 5; Z V-D 5i FX
2. FIRST AND SECOND VARIATION FORMULAS
66
PROOF. If we express V in the coordinate functions (x, y) of R2 and a local coordinate system (.Ti) in M as m f
v'(x,y) t ; ; ) 10
V(x,y) =
'
by the definition we get
D V-
D
8
m
vi
1
8Bl
D DV
m 8vi
axi
8+ 8x
e,1
v,
D8
t ax
8v' D a 8v= D 0 axay 8xi + ay ax 8x' + az 5i azi } dI a2v4 8
8x
DD a t
Hence, if we subtract the equation obtained from the second equation by interchanging x and y from the second equation itself, we get
a OyV - Oy a v
_ c"` *(D D
DD 8 ey8x}
=Lv {ax8y
8Xi.
On the other hand, if we express u(x, y) = (u=(x, y)) in the local coordinate system (a') of M, we get
au
8x -
l m
8u& a 8z 8x1 '
rn
8n
ay - j=L
8u1 a 8y axe
Noting Preposition 1.10, we get
D8
a
TY 8xi
axi ,
1=1
DDa
D
axv 8xi = 8x
mau3
a
°
axi
M auk auJ
".r
1=I
-f- >
Jk-i
ay
a
"` au'
ax 61Y
E'"a2u1 axay °
1=1
8 axe
a S axi
2.3. THE SECOND VARIATION FORMULA
B7
Hence, if we interchange x and y in the second equation and subtract the resulting equation from the original equation, we have
DD
ft0y
DD d 8yax 8:r, 8uk
' (V ea V 3.3
8y
9.k=1
since
8
(0
R
8
8r' _ (V,!. V i
,
holds from the definition of R, noting 1-0
DD_DD -by
8r) V
_
m
l
- V,
'0
a ] = 0, we finally get
v iau kaw
8r
8u u = R ax Oy
)
V
a R (8xk '
a
8
8x`
V.
For a geodesic c : 10, a] - M in M, the first variation E'(0) of the energy functional E for an arbitrary piecewise smooth variation f : 10, a] x (-e, e) -. M is always 0 as seen in §2.1. Namely, geodesics are the curves that give the critical points of the energy functional. As in the extremal problem of calculus, it is needed to investigate the second variation E"(0) of the energy functional at the critical point c, in order to study the behavior of the energy functional E about a geodesic c. In what follows, we calculate the second variation formula.
THEOREM 2.16 (The second variation formula). Let c : [0, a] -
M be a geodesic in M and let f : [0, a] x (-E, E) -} M be a piecee smooth variation of c. Reganiing the energy functional E : (-c, E) --s R, the following holds: q (V, 2 E"(O) (2.8)
jTt2
k s=1
(
di-
dt
2. FIRST AND SECOND VARIATION FORMULAS
68
Here, V and R are, resp , the variation vector field of f and DV (t=) are the cunmftm tensor of M. Furthermore, DV (t;) and given, resper.vely, by
(t;) = t
(t;) _
(t),
(t).
tdti
t>t{
PROOF. If we differentiate the equation (2.3) in the proof of Theorem 2.4 in s, we get
D 8f 8f
d2E _k_
ds2 inn L 8s 8s' 8t
(2.9)
t
+t
+ ,=O
14
(Of D Of ti+1 as, as at ) it.
D 8f D 8j)dt(8f D D Of
-a9
9(8s 8s' at 8t
8s' 8s 8t 8t l
dt.
If we set s = 0, since c is a geodesic, we have D do .- 0. dt dt From this follows that the third term of the right hand side equals 0. Noting that the variation f leaves the end points c(0), c(a) fixed and
that 8f /8t(t, 0) = c'(t) is smooth, we see that the first term of the right hand side also equals 0. On the other hand, from Lemmas 1.26 and 2.15, we get D D Of _ D D B,f D Of _ D Of 8f 8f 8,f R( as, as 8t at 8t as at + 8s ' 8t) at 8s 8t 8t From this readily follows that the second term on the right hand side becomes
o
8fD8f'--g V(t;),DV(t`)_DV(t:) 8s F. at }
it-
i=I
(
at
8t
Furthermore, we get at s = 0,
D D8f _ D D as & &
V + R(V,
dc dc dt )
dt;
dt dt hence, the forth term of the right hand side can be rewritten as
f
9
j
8s' 8 8t at j
2
tit
tR
/ V, dt [
Ja 9 \V (t)' dt2 Henee, we have obtained the desired equation (2.8).
dt.
0
2.4. EXISTENCE OF MINIMAL GEODESICS
69
COROLLARY 2.17. Let c : [0, a] -+ M be a geodesic in M, and let The energy functional E : (-E, E) -i R satisfies
f : [0, a] x (-E, E) -p M be a piecewise smooth variation of c. p
E"(0) = J {g(v', V') - g(R(V, c')c', V) }dt.
(2.10)
0
Here, V and R are, respectively, the variational vector field of f and the curvature tensor of M, and V' = DV/dt. PROOF. Noting, as in the proof of Theorem 2.4, that d (V, DV dt dt g
l _- g
DV DV D2V + g V' &-2 dt ' dt
holds holds with the variational vector field in each interval [ti, tt+I], we readily get (2.11) g
(V, D2V
+ R(V,
dt
o
k
g E {-o
V,
DV ) dt
{9(V, , V) - g(R(V, c')c', V)}dt. ti
We substitute this into (2.8) to get (2.10).
0
Equation (2.10) implies that the second variation E"(0) of the energy functional E depends only on the variational vector field V along the geodesic c and the curvature tensor R of the R.iemannian manifold M.
REMARK. As was stated in §2.1, if we regard the set fl(M; p, q) of all piecewise smooth curves between two given points p and q in a Riemannian manifold M as a "manifold", and the energy functional E : 1I(M; p, q) - R as a "function" defined in this manifold, we may consider the second variation formula (2.10) as the "He sian"of the function E at the critical point c. The "index" of this Hessian and the topology of Q(M; p, q) are closely related to each other, for example, see Milnor [12].
2.4. Existence of minimal geodesics Let (M, g) be a connected yn-d mensional R.iemannian manifiold.
As seen in § 1.5, the distance between two given points p, q E M is
2. FIRST AND SECOND VARIATION FORMULAS
70
defined as the infimum of the lengths L(w) of piecewise smooth curves w joining p and q; namely,
d(p, q) = inf {L(w) I w is a piecewise smooth curve joining p and q }.
The topology in M induced from the distance function d coincides with the topology of M as a differentiable manifold. Theorems 1.25, 1.28 and 1.29 together yield that there exists a unique minimal geodesic c, i.e., L(c) = d(p, q), joining any two points p and q in a sufficiently small geodesic sphere W of M.
Given two points p, q E M, there may not always exist a minimal geodesic joining p and q. For example, consider the Riemannian manifold obtained by removing the origin from the Euclidian space E'11. Even if such a minimal geodesic exists, it may not be unique, as can be seen readily in the case of an antipodal pair in the unit sphere SR`.
In this section, we investigate the existence problem of minimal geodesics. We begin with conditions for any two points p, q E M of a Riemannian manifold M to have a minimal geodesic joining them. DEFtNITION 2.18. Let M be a R.iemannian manifold. If the exponential map expp at p is defined for any tangent vector v E TpM at any point p E M; namely, any geodesic c(t) emanating from p is defined for every t E R, M is called geodes°icaldy complete. For example, as seen in Examples 1.21, 1.22, 1.23, the rn-dimen-
sional Euclidean space E'", the unit sphere (STI, g) in E"'+', and the m-dimensvonal real hyperbolic space (H, g) are all geodesically complete. They are also complete as a metric space (M, d) with the distance d induced from the Riemannian metric; namely, any Cauchy sequence with respect to d converges. The following important result called the Hopf-Rinow theorem shows that these two notions of completeness are not only equivalent, but also the exact condition which assures the existence of minimal geodesics.
TxE OREM 2.19 (Hopf-Rinow). Let (M, g) be a connected Rie-
mannian manifold and let p be a point in M. Then the following conditions are equivalent: (i) The exponeniialznap exp, at p is defined in the entire tangent Spam
Tit.
(ii) Any bounded dosed subset of (M, d) is compact.
2.4. EXISTENCE OF MINIMAL GEODESICS
71
(iii) (M, d) is a complete metric space. (iv) (M, g) is geodesically complete. In what follows, an (M, g) satisfying one of the above conditions, therefore, all of them, will simply be called a complete Riemannian manifold.
THEOREM 2.20 (Hopf-Rinow). If (M, g) is a complete and connected Riemannian manifold, for given two points p, q E M, there is a geodesic c joining p and q such that L(c) = d(p, q); namely, there is a minimum length geodesic joining p and q. We first prove Theorem 2.20.
PROOF. Set d(p, q) = r. We only need to treat the case r > 0. Given a point p, under the condition (i) of Theorem 2.19, we show that for any q E M, there is a minimum length geodesic joining p and q.
FIGURE 2.2
Let W be a normal coordinate neighborhood of the point p as given in Theorem 1.25. For sufficiently small 6 > 0, let Ba(p) C W denote the geodesic ball of radius 6 with p as its center. Since the boundary S = SS (p) = {x E W1 d(p, x) = S} of Ba (p) is a compact set and since the distance d(q, x) from a point q is a continuous function in S, there is a point xo E S such that d(q, xo) assumes the minimum. Using the exponential function expp, xo can be expressed as xo = expp(dv),
v E TpM, IvJ = 1.
From the assumption, the normal geodesic c(t) = expp(ty) is defined at all t E JR for this v. The proof will be done if we can show c(r) = q.
2. FIRST AND SECOND VARIATION FORMULAS
72
In fact, since L(c) = r, c will be a minimum length geodesic joining p and q. In order to show c(r) = q, we see that the equation
d(c(t), q) = r - t
(2.12)
holds for each t E [0, r]. Setting t = r in (2.12) we see that c(r) = q since d(c(r), q) = 0. Now set I = It E [0,r]1 (2.12) holds}.
Clearly, 136 0 since 0 E I. Also from the definition, I is a closed subset of [0, r]. Hence, assuming to E I, it suffices to verify that if to < r, there is a sufficiently small 8' > 0 such that (2.12) holds for to + b'. From this, we get sup I = r. Since I is a closed subset, r E I; consequently, we get I = [0, r]. This completes the proof. In order to see that (2.12) holds at to+8', we consider the geodesic
ball B (c(to)) of radius b' and center c(to) and its boundary S' _ S',g(e(to)). As before, there is a point xo' E S' such that the distance d(x, q) between q and x E S' is minimized at xfl. It suffices, then, to see xo' = c(to + 6') for this xfl'. In fact, from d(c(to), q) = r - to due to to E I and from d(c(to), q) = 8'
+ DES' d(x, q) = 8' + d(xo', q).
we see that r - to = 5' + d(xo', q) = 6' + d(c(to + b'), q).
(2.13)
From this, we get
d(c(to + 8'), q) = r - (to + 8'); consequently, we see that (2.12) also holds at to + 8. Next, we show c(to + b') = xo'. From the first equality in (2.13) and the triangle inequality of the distance function, we have d(p, xo') >d(p, q) - d(xo', q) = r - (r - to - 6') = to + 6'. We finally see that d(p, xo') = to +6', since the length of the piecewise
smooth curve obtained by connecting the portion of the geodesic c between p and c(to) and the minimum geodesic c' between the points xo' and c(to) is to+b'. Hence, by Theorem 1.29, this piecewise smooth
curve is nothing but a geodesic joining p and xo'. In particular,
0
c(to + 8') = xo', which was what needed to be shown. We next prove Theorem 2.19 in the order of (i) (iv)
(i).
(ii)
(iii)
2.4. EXISTENCE OF MINIMAL GEODESICS
73
Let A C M be a bounded closed set. (ii). PRooF. (i) Since A is a bounded subset of (M, d), the diameter p = diam(A) = sup{d(x, y)lx, y E Al is finite. Hence, there is a geodesic ball B centered at a point p such that A C B. From the condition (i) and Theorem 2.20 follows
AC BCexppBr(0) for the r neighborhood B,.(0) C TpM of the origin 0 in the tangent space T.M. Since the closure B,.(0) of B,.(0) is a compact subset of T,pM, the image expp B,.(0) of the exponential map exp, is a compact subset of M; hence, its dosed subset A is also compact.
Let {pk} be a Cauchy sequence in (M,d). Since from the definition [p&) is a bounded subset of M, we see from (ii) that {pk} is relatively compact. Therefore, {pk} has a convergent (ii)
(iii).
subsequuence. Since (Pk} is a Cauchy sequence, it also converges. (iii) (iv). Let p E M be an arbitrary point of M, and let c be a normal geodesic emanating from p. For c, we set
I+ = it > 01 c is defined in [0, t] }.
Then we show t+ = sup I+ = +oa. Assume t+. < +oo, and let {tk } c I+ be a sequence converging to t+. From the assumption, given any E > 0, for sufficiently large k, ! with I tk - ti l < e, we have d(c(tk), C(tt)) < L(c I [tk, t1]) = Itk - tj I <
Hence, {c(tk)} is a Cauchy sequence of M, and {c(tk)} converges to a point q E M from (iii). If we tale the normal coordinate neighborhood W and 5 in Theorem 1.25 about the point q, we see that c(tk), c(ti) E W for sufficiently large k < 1, and that c I [tk, 41 is a minimum geodesic joining c(tk) and c(t1). If we note here that exp,(t,,) is a C"° diffeomorphism from Bj(0) C TTt,,)M onto expC(tki(Ba(0)) D W for each tk, we readily verify that c can be extended beyond t+. This contradicts the definition of T+. Hence, t+ = +oo. Similarly, for I_ = {t < 0 1 c is defined in It, 0]}, we can show that inf I- = -oo. Thus, we have shown that M is geodesically complete. 0 (1) is self-evident. (iv) From Theorem 2.19 follows
74
2. FIRST AND SECOND VARIATION FORMULAS
COROLLARY 2.21. A compact connected Riemannian manifold M is complete.
A geodesic c : [0, a] -- M in M with c(0) = c(a) = p is called a geodesic loop in M with a base point p. Namely, a geodesic loop c is a geodesic in M whose initial point p(0) and terminal point c(a) coincide. In general, the tangent vector c'(0) and c'(a) do not coincide. A geodesic loop c with c'(0) = c(a) is called a closed geodesic of M. In other words, a closed geodesic c is a CO° map from a circle S' into M such that the geodesic equation Vec' = 0 is satisfied at each point of the one-dimensional manifold S1. We now investigate the existence problem of such closed geodesics. First, we review the following definition.
DEFINITION 2.22. In general, a continuous map c : [0, a] -+ M with c(0) = c(a), namely, a continuous curve in M with the identical initial and terminal points, is called a loop in M. Given two loops
co:[0,a]-Mand ci:[0,al -+M in M, a continuousmapf;[0,a]x [0,11 -+ M satisfying the following two conditions (i) and (ii) is called a frm homotopi/ from co to cl : (i) f (t, 0) = co(t), f (t, 1) = city, t E [0, a]. (ii) f (0, s) = f (1, s),
s E 10, 1].
When there is a free homotopy such as above, co and cl are called free hornptopic, and we denote it by co ~ ci. A free homptopy between two loops is different from the homotopies defining the fundamental group 7r(M) of M. We point out that the "deformation" f. (t) = f (t, a) does not necessarily keep the base
point f9(0) = f, (a) of f, fixed. It can be readily seen that the relation being free homotopic among the loops is an equivalent relation. Given a loop c in M, the equivalent class [c] = {ci I ci is a loop free homotopic to c}
in this equivalence relation is called the free homotopy class of c. The set consisting of all the free homotopy classes is denoted by Cl (M).
The free homotopy class of a trivial map is called the trivial class, and the rest are called nontrivial classes. Regarding the existence problem of closed geodesics, we have
THEOREM 2.23. If (M, g) is a compact Riemannian manifold, there exists a closed geodesic with minimum length in each nontrivial free homotopy class a E CI (M).
2.4. EXISTENCE OF MINIMAL GEODESICS
75
PROOF. Let d denote the infimum of the length L(w) of the piece-
wise smooth loop w belonging to a. Choose a sequence {cj } C a of piecewise smooth loops c3 belonging to a such that L(c3) --+ d = inf L(w).
(2.14)
Here, each Cj is defined in the closed interval [0,1]. Furthermore, without loss of generality, we may assume that c j [t J _ 1 i t3 ] is a geodesic in each interval [t3 _ 1: t3], where it is a C°° curve, and that the parameter t is an affine parameter proportional to the are length. We denote by K an upper limit of the arc length L(c3) of e j. Then for any t1 < t2 E [0,1], we can easily verify that I
t2
Ic'3(t)Idt < K(t2 - t1). J Therefore, the sequence {c3 } is an equicontinuous family. Since M is compact from the assumption, Ascoli's theorem implies that by d(cj(tt),c (t2)) <_
choosing a subsequence of {c3 } if necessary, {c3 } converges uniformly
to a continuous loop co : [0,1] , M. Note co E a. Since the image c([0, 11) of co is a compact subset of M, we can < tk = I of [0, 1] in such a way choose a partition 0 = to < t1 < that each co I [ti_1, ti] is contained in the normal coordinate neighborhood Wi defined in Theorem 1.25. If we then replace co I [ti-1, ti] by the minimum geodesic c' joining co (ti-1) and co(ts) in Wi, we get a piecewise smooth loop c : [0,1] -p M in M such that c' = c I [t1_1, ti] for each i (1 < i < k). As is clearly seen from the construction, c
is free homptopic to co. Hence, c E a and the length of c satisfies L(c) > d. We now verify that L(c) = d holds. We assume L(c) > d and set F > 0 to be e = (L(c) - d)/(2k + 1). From the definition of d and the fact that c j converges to co uniformly, for sufficiently large number j, we see that
L(c3) - d < c,
d(cj(t), co(t)) < c (t E [0,1])
holds. If we set cj = c3 I [ti_ 1, ti] for c3, we get k
E(L(c?) + 2E) = L(c3) + 2k < d + (2k + 1)E k s=1
76
2. FIRST AND SECOND VARIATION FORMULAS
Hence, we see that
L(c) + 2f < L(c') holds for some i (I < i < k). This implies that the length of the piecewise smooth curve obtained by joining the minimum geodesic between co (ti-1) and c j (t{-1),
the curve c`c and the minimum geodesic between cc(ti) and c. (t;) is
shorter than the length of c. It contradicts that c' is the minimum geodesic joining co(ti-1) and co(ts). Consequently, we get L(c) = d. Since a is a nontrivial free homotopy class from the assumption, L(c) = d > 0. If we can show that c is a dosed geodesic, this c is what we want. Now parametrizing c by arc length, we see from the way it was constructed that c : [0, al -s M is a piecewise smooth loop such
that ea& c' = c I [t;-1, tj4 is a geodesic and that it has the shortest length among the curves in the free homotopy class a. Hence all we need is to verify that c is C°° about the point p;, = c(t;) for each
i(f
FIG JR.E 2.3. A piecewise smooth variation
Now let us assume that c is not C°0 at pi, for instance. Namely, assume the tangent vectors e'(tl) 36 c`'. We consider a geodesic ball B of center pi with a sufficiently small radius as shown in Figure 13. Take two points qI, 42 E e([tj-1, ti=ll) n B near p and join them by the minimum geodesic in B. Then we get a piecewise smooth loop in
a with a shorter arc length than c. This contradicts that c has the n3dnimum length in a. Therefore, c is C°0 at pt.
0
Like the sequence {cs} of piecewise smooth loops in the proof of Ileorem. 2.23, a sequence consisting of elements in the domain of definition of L satisfying (2.14) is called a minimizing sequence with respect to a given functional L. A method to prove the existence of a critical point of the functional L by directly showing the existence of
25. APPLICATIONS TO RIEMANNIAN GEOMETRY
77
a converging subsequence for a given minimizing sequence as in the proof of Theorem 2.23 is called the direct Method in the calculus of variation. When the Riemannian manifold (M, g) is not compact, we should
note that Theorem 2.23 does not always hold true. In fact, for example, consider a surface of revolution M in E3, asymptotic to the axis of revolution as in Figure 2.4. Then we see readily that there is a nontrivial flee homotopy class a E C1 (M) with no minimum length loop.
FIGURE 2.4. A surface with no closed geodesic
REMARK. Since a geodesic in the unit sphere (St', g) in E'"+1 is a great circle (or a portion of a great circle), as seen in Example 1.22, there are infinitely many closed geodesics in (S"', g). However, since S' is simply connected, we cannot conclude the existence of closed geodesics from Proposition 2.2. In general, whether there are infinitely many geometrically mutually distinct closed geodesics in an arbitrary compact Riemannian manifold (M, g) is an important research topic in Riemmannian geometry. The studies on this subject have a long history, including one from the view point of Morse theory; see, for instance, Klingenberg [8].
2.5. Applications to Riemannian geometry As seen in §2.3, the second variation formula regarding the energy of curves is naturally related to the concept of curvature in R.iemannian manifolds. Using it, one can investigate the effects of the curvature
of Riemannian manifolds on topological structures. In this section, we discuss, as an example, the topological structures of Riemannian
2. FIRST AND SECOND VARIATION FORMULAS
78
manifolds with positive sectional curvature using the second variation formula. Let (M, g) be a connected Riemannian manifold of dimension m. If M is complete, the following holds regarding its diameter diam(M) = sup{d(p, q) I p, q E M}. THEOREM 2.24 (Myers). Let (M, g) be a complete, connected m-
dimensional Riemannian manifold If them exists a positive number k > 0 such that Ric > (m - 1)k2g
with respect to the Ricci tensor R.ic of M, the following hold:
(i) diaan(M) < it/k. (ii) M is compact. (iii) The fundamental group 7r1 of M is finite. PROOF. Let p, q be any two points of M. Since M is complete and connected, from Theorem 2.20, there is a minimum geodesic c : [O,1] --- M connecting p and q. We will verify that the length L(c) of c is less than or equal to r/k. From this follows d(p, q) < r/k. Since p and q are arbitrary, we get diam(M) < it/k; consequently, (i) is shown. From (i) follows that M is complete and bounded. Hence, Theorem 2.19 implies that M is compact, proving (ii). We will reach a contradiction, assuming L(c) > r/k. The parallel displacement of an orthonormal base lei (0), ... , em-1(0), c'(0)/l} at the point p = c(0) along c gives rise to cOQ vector fields el (t), ... , em-i(t), along c such that e11(t) - c'(t)/l. For each t E [0, 11, {e1 (t), ... , e1z(t)} forms an orthonormal base for the tangent space T(t) M. Now we define a vector field Vj along c by VJ (t) = (sun irt)es (t),
I < j < m - 1.
Since VJ (0) = V1 (1) = 0, as was seen in Proposition 2.2, there is a C°° variation of c whose variation vector field is V. for each j. Let E7 (s) = E(f9 s)) denote the function that assigns to f; its energy. Since each e; (t) is a parallel vector field along c, a simple calculation using the second variation formula in Theorem 2.16 yields 2
E/' = - J g(V,, V-' It + R(V,, c')c')dt 0
(2.15)
=
r1 0
sine rt(7r2 -12K(e?(t), e,, (t)))dt.
2.5 APPLICATIONS TO RIEMANNIAN GEOMETRY
79
Here, Vj" = D2 V;/dt2, and K(ej (t), e,,.(t)) is the sectional curvature of the two-dimensional subspace determined by the orthonormal base elements {ej(t),e,,.(t)} in the tangent space TT(t)M. If we add the equations (2.15) over j = 1,... , m - 1, we readily get from the definition of Ricci curvature m-1
I
Ej"(0) = J sin2wt((m - 1)ir2 -12 R.ic(c(t))(em(t), e,..(t)))dt. 0
j=r
From the assumption Ric > (m
kl>ir,weget
- 1)kag, and since we also assumed
12 Ric(c(t))(em(t), em (t)) > (m
- 1)k212 > (m - 1)7r2;
consequently, we get m-1
E Ej"(0) < j=1
fI
J0
sin27rt((m
- 1)zr2 - (m - 1))dt = 0.
This implies that there is some j (1 < j < m -1) with which Ej"(0) < 0 holds. This contradicts that c is a minimum geodesic. In fact, as seen in Lemma 2.3, c has the minimum energy; namely, E(c) < E(w) holds for any piecewise smooth curve w connecting p and q. Therefore,
Ej"(0) < 0 for each variation f3 implying 1 < 7r/k. In order to show (iii), let Cj : M , M be the universal covering map, and let g = w'g denote the Riemannian metric induced from g through w in the universal covering manifold k of M. Then w gives rise to an isometric and locally diffeomorphic map from (M, 4) onto (M, g). Noting that the Levi-Civita connections and Ricci tensors hold unchanged under w, we see readily that (M, g) also satisfies the conditions in the theorem (see Exercises 2.8 and 2.9 at the end of this chapter). Consequently, (ii) implies that k becomes compact. From this we see that w-1(p) (p E M) is a finite set; hence, the fundamental group 7r1(M) is finite.
0
As a special case of Theorem 2.24, we have COROLLARY 2.25. Let (M, g) be a complete and connected Ric-
moannian manifold of dimension m. If the sectional cun'atmw K of M satisfy K > k2 > 0, M is compact, diam M < sr/k, and the fundamental group ir(M) is finite.
so
2. FIRST AND SECOND VARIATION FORMULAS
In Corollary 2.25, condition K > k2 > 0, regarding the sectional curvature, cannot be weakened simply to K > 0. In fact, the parabolic surface of revolution in the three-dimensional Euclidean space E3 given by
M={(z,y,z)EE'31z=z2+y2} is complete and the sectional curvature K > 0 everywhere, but not compact. We discuss another application of the second variation formula. THEOREM 2.26 (Synge) - Let (M, g) be a compact, connected, and
orientabk Riemannian manifold of even dimension. If the sectional curvature K of M is always positive, then M is simply connected PROOF. First of all, we note the following. For the piecewise smooth variation f : [0, a] x (-E, c) -f M of a geodesic c : [0, a] -M given in Corollary 2.17, if we do not assume condition (iii) of Definition 2.1; namely, if the deformation obtained from the family { f, } of curves in the variation f does not necessarily keep the end points c(0), c(a) fixed, the variation vector field V (t) of the variation f may not satisfy V(O) = V (a) = 0. Consequently, the first term on the right hand side of the equation (2.9) in the proof of Theorem 2.16
is not necessarily equals 0. Also the terms at t = 0 and t = a in the second term on the same right hand side remain nonzero. These
terms correspond to the cases at t = 0 and t = a in the first term on the right hand side of equation (2.11), as one can see in proof of Corollary 2-17.
Keeping these in mind, if we repeat the arguments in the proofs of Theorem 2.16 and Corollary 2.17, we see readily that the second variation formula of the enemy function E fur the general variation f, which does not necessarily keep the terminal points c(0), c(a) fixed, is given by (2.16)
E"(0) _
10
f{g(V, V') -g(R(V, c')c',V)}dt
D
- g (Ts , f (0, 0), c"(0)) + g
(D ,
f (0, a), c'(0)
Here, V` = DV/dt and R is the curvature tensor of M. Now we assume that there is a nontrivial free homotopy class a E CI (M) of a loop in M. By Theorem 2.23, there is a minimum closed geodesic c in a. Let t be the arc length parameter and let e : [0,1] -- M
2.5. APPLICATIONS TO RIEMANNIAN GEOMETRY
81
represent the arc length parametrization of c. As was seen in Proposition 1.15, the parallel transport Pc : TT(o)M TC(i)M = T'(o)M defines a linear isometry of TT(o)M onto itself. Since c is a dosed geodesic, Pc(c'(0)) = c'(l) = c'(0) holds. From the assumption, M is orientable; therefore, Pc defines an orientation preserving orthogonal transformation I cl(0)i -, c'(0)1 PC
in the orthogonal complement c'(0)-i C Tc(o)M of c'(0) in TCo)M. On the other hand, M is even dimensional from the assumption. Hence c'(0)1 is odd dimensional. As is well known in linear algebra, the orthogonal transformation PP I c'(0)1 has 1 as an eigenvalue, and it leaves the corresponding eigenvector v # 0 fixed. Namely, we have
P°(v) = v. Consequently, by transporting v parallelly along c, we obtain a vector field V(t) along c which is perpendicular to c at each point c(t). If we consider the variation of c given by f (t, s) = exp0t) sV(t),
(t, s) E [0, a] x (-e, c),
as treated in Proposition 2.2, noting V' - 0 and v(O) = V(a), the second variation formula (2.26) combined with the assumption K > 0 on the curvature readily yields
E"(0)
j
g(R(V, c')c', V )dt < 0.
Of course, E'(0) = 0 for the first variation formula of E. Hence, we see that E(fe) < E(c) for sufficiently small 0 < s < c. With the length of f8, this implies that L(fs)2 < 2lE(f$) < 21E(c) = L(c)2. Since it is clear from the definition that f, and c are free homptopic C°° curves, this yields the contradiction that the length L(c) of c is the minimum in a. Consequently, we see that M possesses no nontrivial 0 five homotopy class; hence, M is simply connected. We point out the following regarding the assumption in Theorem 2.26. The three-dimensional real projective space P3(R) is a compact orientable C°° manifold, and it possesses a natural R iemannian metric g so that the universal covering map tii : S3 -> P3(R) gives rise to an isometric and locally diffeomorphic map. With respect to this g, the sectional curvature K of P3(R) is always positive, but P3(R) is not simply connected. Hence, the assumption that the manifolds
2. FIRST AND SECOND VARIATION FORMULAS
Sa
are even dimensional cannot be removed. The assumption that they are orientable cannot also be removed. Indeed, two-dimensional real projective space P2(R) is a nonorientable C'0a manifold, and it possesses a Riemannian metric g such that the sectional curvature K is always positive as in the case of P3(R). However, p2(R) is not simply connected (see Exercise 2.10 at the end of this chapter).
Summary 2.1 The first variation formula regarding the energy E(c) of a curve c and characterization of the geodesics as the critical points of the energy functional E. 2.2 The de initions, the l3iema nian curvature tensor, and various curvatures on a Riemannian manifold. 2.3 The second variation formula of the energy E(c) of curves. 2.4 Existence of a minimal geodesic joining two arbitrary points p, q E M in a complete Riemannian manifold M, and existence of a al closed geodesic within each nontrivial free hoinotopy class of n a closed curve in a compact Riemannian manifold. 2.5 Theorems of Myers and Synge regarding the topology of complete Riemanniau manifolds of positive curvature.
Exercises Let (xi) and (J{) be local coordinate systems about a point x in a C°O manifold M. Show the following:
2.1
(1) Between the components 7'' andi...i, of a tensor field T E T$(M) of type (r, s), a transfiamiation identity &k3 ftk1 . fr 1
1
klt ,*. LI,.
Itr
holds.
(2) Given two local coordinate systems (xi), (z) in an open sub-
set U, that the components of T E ;(M) with respect to (x{) are C'° implies that the components with respect to (r) are C°° . 2.2 Show that the space T.1 (x) of all tensors of type (1, s) over the tangent space Tz to a C0° manifold M is canonically isomorphic
to the vector space Hom (TM x ... x TM, .,x) of all s linear maps from the direct product TxM x ... x TxM into T.M.
EXERCISES
2.3
83
A linear connection
V : X(M) x X(M) - X(M) in a C°° manifold M does not define a C°° tensor field of type (1,2) over M. Why not? 2.4 Let (xi) denote a local coordinate system in an m-dimensional Riemannian manifold M. Show that a=
'jk = ax rjk - aaxe r4k +
m
(rirr,k - l1 rr k) r=1
holds regarding the components of the curvature tensor R2.5 Let o C TxM be a two-dimensional subspace of the tangent space TaM of M at a point x E M. Show that the sectional curvature K(a) of a is given by
K(cr) =
gx (R(v, w)w, v) ga(v,v)gx(w,w) - 9x(v,W)
with respect to any base {v, w} for o. 2.6 Show that the curvature tensor at a point in a Riemannian manifold is completely determined by the sectional curvatures of all possible two-dimensional subspaces in the tangent space at the point. 2.7 Suppose that the parallel transport between two arbitrary points in a Riemannian manifold M is determined independently of the choice of piecewise smooth curves between the points. Then show that the curvature tensor R of M is identically 0; namely, R(X, Y)Z = 0,VX,Y, Z E X(M).
2.8 Let co : M -+ N be a C°° diffeomorphism from a Riemannian manifold (M, g) into another Riemannian manifold (N, h). If gx(v,w) = hp(x)(dPx(v),&Px(w)), Vv,w E TxM, holds at each point x E M, cP is said to be isometric. Given an isometric diffeomorphism cp : M -+ N, show the following:
(1) Let V and V' be the Levi-Civita connections of M and N, respectively. Then d&o(V xY) = v4(x)dco(Y),
X, Y E X(M).
Also let R and R' be the curvature tensors of M and N, respectively.
2. FIRST AND SECOND VARIATION FORMULAS
84
Then dsp(R(X, Y)Z) = R'(dco(X ), d(p(Y))d4p(Z),
X, Y, Z E X(M),
holds. (2) If c is a geodesic of M, V (c) is a geodesic of N. 2.9 Let (M, g) be a Riemannian manifold and let c:,1 : M ---; M be a COO covering of M. Prove the following:
(1) M has a Riemannian metric g so that w becomes a locally isometric diffeomorphism. (2) (M, g) is complete if and only if (M, g) is complete.
2.10 Let M be a compact and connected even-dimensional Riemannian manifold. Assume that all the sectional curvatures of M are positive. Prove that M is either simply connected or the fundamental group irl(M) = Z2.
CHAPTER 3
Energy of Maps and Harmonic Maps In this chapter, we define a functional called the energy of maps in the spaces of smooth maps between Rieman.n.ian manifolds and discuss harmonics maps which are the critical points of the functional. The energy of a map is a spontaneous generalization of the energy of a curve as seen in Chapter 1. The harmonic maps include, for examples, harmonic functions, geodesics, minimal submanifolds, isometry, holomorphic maps, etc.
3.1. Energy of maps Let (M, g) be an mrdimensional Riemannian manifold. Denote
by (xl, ... , x') a local coordinate system in an open subset U in M. The local coordinates of a point x E U in M are expressed as (x')(1 < i < m). At each point x E U,
forms a base for the tangent space TIM, and
f(dxl)z,... ,(dxm)r} forms the dual base for the dual space TIM*. Namely,
W)x((ax'/zl
=o,
1
In U, the Riemannian metric g of M is given by m
g = > g=i dxidxi
.
ij=1
Here the components g1j of g in (x') are C° functions in U defined by (a
al
gs'
g
axs' ax? 85
3. ENERGY OF MAPS AND HARMONIC MAPS
86
and (g`j) forms a positive definite m x m symmetric matrix at each point in U. Let (gij) denote the inverse matrix of (g1 j ). Namely, we set (3.1)
E9'k9kj = b`,
94k9kj _ Sj
1 < i, j < M.
k=1
k=1
The R.iemannian metric g, induces a natural linear isomorphism between the tangent space T,,M and the dual space TXM* defined by
5:T,,M-+TM*,
(3.2)
0:TxM* _,TTM.
In fact, for given X. E TM, wx E TM', we may set XT(Y3)
°
9x(wz1Yz) = w1(YY), YY E T1 M.
91(X1,Yx),
If we, using the local coordinate system, express m m a X1 _ X'(x)(; )x, wx =
X , ws are, respectively, given by
m m XX = E(Eg;j(z)X3(x))(dx`)., i=1 j=1 m in w!
_ E E 9ii (x)wj (x) i=1
j=1
Using this linear isomorphism, an inner product gx in TZM* dual to the inner product gx in TxM is defined, for given wx, es E T,,M*, to be 99(Wx, 8x) = 9, (WI' 61).
From this definition combined with (3.1) and (3.3), an easy computation yields 9x((dx{)a, (dx')x) = g1 (x).
Namely, the inverse matrix (gij(x)) of (gj(x)) is precisely the matrix representing the components of the inner product g.*, in TxM*. Let (N, h) be an n-dimensional Riemannian manifold and let u : N be a C° map from M into N. Denote by (y1, ... , y") a M local coordinate system in an open subset V in N. Express the local
coordinates of a point y E V in N by y = (y") (1 < a < n). Then
3.1. ENERGY OF MAPS
87
for a given x E U with u(x) E V, the local coordinates of u(x) are, in terms of C°° functions Ua = yk o u in U, expressed as (3.4)
u(s) = (ul (xl, ... , 2m), ... , un (x1, ... , xm)).
The Riemannian metric h of N is also expressed in V as n
h = E h.,fldyadyI3. a,0=1
Let us discuss the differential of u at a point x denoted by
du= : TM -+ duz is a linear map from the tangent space TTM into Tu(,,)N. If we represent u as (3.4) using the local coordinates, we get n
du, (()) =
(8x=) (x) ( ay ) UkX) ,
I
< 97 b.
In other words, due is the linear map represented by the n x m matrix
((8ua/8x')(x))
A.s is well known, the linear space Hom(T.M, Tu(s) N) consist-
ing of all linear maps from TsM into Tu(,,)N is naturally linearly isomorphic to the tensor product ,,s* ® Tu(s)N. In fact, it is obtained by assigning to f E Hom(TxM, Tti(s)N) a bilinear map f t in TsM* ®TT(s)N defined by f I (v, w) = w(f (v)),
v E TM, w E Tu(s)N*.
Hence, the differential dux of u can be regarded as an element of TJ M* 0 T,,(s)N. Namely, we have (3.5)
du,, E Hom(TsM,T,,(s)N)
TJM* OTT(z) N.
Since a base for TXM* ®Tu(,,)N is given by (3.6)
{(dxt)s ®(8/ a)L(s) I < i < m, 1 < a < n},
due is represented by sn
(3.7)
n
l i)// (x)(dx`)z ®(
du,, _ m .L ( i=1 a_1
)// (xy
On the other hand, the inner products gi in T=M* and hu(x) in Tu(s)N naturally induce an inner product (, ) in the tensor product
3. ENERGY OF MAPS AND HARMONIC MAPS
88
TXM* 0 T,,(,,)N. In fact, for the base in (3.6), we set
(dx')x u(x)
®(
g'3ha#(u(x))
u(x)
x
and extend it bilinearly to arbitrary elements. In other words, we define the components of the inner product (, ) as the tensor products (g'"(x))®(ha,9(u(x))) = (gaj(x)(ha0(u(x))) of the matrices representing the components of gx and hu(x). We denote by IduxIx the norm
of du with respect to this inner product. As seen readily from the definition of the inner product and (3.7), 1dti
is given by
j
Idu, I = (du,, duz)= n M
a
E g''(x)hao(u(x))
l (x)
(x)
From these facts, we see the following. Given a C°° map u : M -> N from a R.iemannian manifold (M, g) into another Riemannian manifold (N, h), we consider the induced vector bundle u-ITN by u over M from the tangent bundle TN of N. u-ITN is the vector bundle over M whose fiber over x E M is the tangent space Tu(g) N of N at u(x). Next denote by T*M the cotangent bundle of M, and
consider the tensor product T*M ® u-'TN of T*M and u-'TN. T*M®u-iTN is nothing but the vector bundle whose fiber at x E M
is TTM 0 Tu(z)N. Denote by I'(T*M 0 u-'TN) the space of all C°° cross sections of the vector bundle T*M 0 u-'TN. Namely, r(T*M®u-iTN) is the set of all CIO maps a : M --> TM* ®u-ITN such that a(x) E TXM* 0 Tu(.,)N holds at every x E M. Define a map du : M -+ TM* ®u-iTN by setting du(x) = du, for the differential du,, of u at each point x in M. From (3.5) and the local representation (3.7) of dux, we see readily that du determines a C°O cross section of the vector bundle T*M ® u-iTN. Namely, we have
du E r(T*M (g u-'TN).
On the other hand, a fiber metric (
,
) is naturally defined in
the vector bundle T*M 0 u-'TN from the inner product (, )z in the tensor product TZM* 0 Tu(x)N. In fact, given a,' a E I'(T*M 0 u-'TN), we simply define (a,' a) (x)
= (a{x),' a{x)) z, x E M.
89
3.1. ENERGY OF MAPS
Given du E r(T*M®u-1TN), we can speak of the norm Idul with respect to this fiber metric (, ) . From the definition, ldul (x) = Idu,, Iz, and Idul is given by (3.8)
auQ
n
m
Idu12
= E E 9 haA(u) (axi) i,j=1 a,3=1
(2)
Under the above observations, we defined the energy density of a map u as follows. DEFINITION 3.1. Given a C°° map u : M -p N from a R.iemannian manifold (M, g) into another Riemannian manifold (N, h), a C°° function e(u) E C°°(M) defined by
e(u)(x) = 2 Idul2(x),
x E M,
is called the energy density function of u or simply the energy density. It is clear from (3.8) that the energy density e(u) is a C°° function. Let {eI, ... , em}, {e'1, ... , e',b} be, respectively, orthonormal bases
for the tangent spaces TM, ,,,(,)N with respect to gz and hu(,). We express du,, in these bases as n
dux(ei) _
E
'M.
or=1
Then it is readily verified, from its definition, that Idul(x) is given by m 1du12(x)
i=1 a=1
Consequently, we may regard Idul2 (z) as representing the square sum of the "rate of expansion" for the differential du,, : TTM -+ T, (z)N of u in the mutually orthogonal directions. This is the reason why we call e(u) the energy density of u. DEFINITION 3.2. Let (M, g) be a compact Riemannian manifold. Then the integral of e(u) given by (3.9)
E(u) =
e(u) d1ts
Jet is called the energy or the action integral of the map u. Here, p9 represents the standard measure induced in M from the Riemannian metric g (see Exercise 3.1 at the end of this chapter).
90
3. ENERGY OF MAPS AND HARMONIC MAPS
Let C°D(M, N) denote the space consisting of all the CO0 maps from the Riemannia manifold M into the Riemannian manifold N. .
If M is compact, the energy E(u) E R is defined by (3.9) for each u E C°°(M, N). Hence, the energy of maps is regarded as defining a functional
E:C(M,N)-+R. Our objective in the following is to find maps which are critical points of this functional E.
3.2. Tension fields As preparation for obtaining the first variation formula charaaoterizing the critical points of the energy functional E, we consider the second fundamental form of a map and its tension field. Let (M, g) and (N, h) be Riemannian manifolds of dimensions m
and n, respectively, and let u E C'0 (M, N) be a C°° map from M into N. Let TM` 0 u`ITN denote the tensor product of the cotangent bundle TM' of M and the bundle u-1TN induced over M by u from the tangent bundle TN of N. As seen in the last section, TM* 0 u ITN has the fiber metric L , ) induced naturally from the metrics g, h. First, we verify that there is a naturally induced connection compatible with this fiber metric (, ); in other words, (, ) is parallel with respect to the connection. The Riemannian metric g induces the Levi-Civita connection V in the tangent bundle TM of M. Namely, as in Theorem 1.12, given CO° vector fields, or C°° cross sections X, Y, Z E r(TM) of TM, them exists a unique connection V which satisfies the following conditions: (3.10)
X9(Y, Z) = 9(VxY, Z) + 9(1', VxZ),
(3.11)
VXY - VYX = [X, 11.
In fact, given a local coordinate system (x') of M, we define the connection coefficients {r!,) of the Levi-Civita connection V in (a-') by in
(3.12)
V.
&zs
From (L27), r k is given by rn
ag;!
r;; = 21 E gki !=1
C
a9;!
ax, + axe
- '09ij a24
3.2. TENSION FIELDS
91
Here, gi3's are the component of g in (xi) and gii's are the components of the inverse matrix of (gi,) For a given Y E P(TM), if we define
VY(X) = VXY, X E r(TM), we obtain the following tensor field of type (1, 1): VY E r(TM* (& TM)!---' Hom(TM,TM).
Consequently, the Levi-Civita connection V of M defines a map
V : r(TM) _ r(TM* (9 TM), which assigns a CO° tensor field VY E r(TM* (& TM) of type (1,1)
to a C°° tensor field Y E F(TM) of type (1, 0). We call VY the covariant differential of Y. We can define, from the Levi-Civita connection V in TM, a con-
nection V* in TM* as follows. First, we note that from the linear isomorphism in (3.2), we obtain bundle isomorphisms
b:TM -,TM*, between the tangent bundle TM and the cotangent bundle TM*. Using these bundle isomorphism, given w E r(TM*) and X E r(TM), we define V*Xw E r(TM*) by V*Xw(Y) _ (Vxwo)b (Y),
(3.13)
Y E r(TM).
It is easily verified that V*Xw satisfies the same computational rules as the covariant differentiation VXY in the tangent bundle TM. Call V*Xw the covariant derivative of w by X. For given w E r(TM*), define
V*w(X) = V*xw, X E r(TM). Then we obtain a type (0, 2) tensor field V*w E r(TM® (9 TM*)
Hom(TM (9 TM*).
The map
V*: r(TM*) -. F(TM* (& TM*) that assigns the type (0, 2) C°° tensor field V*w to the type (0,1) Co' tensor field w is called the connection in TM* determined by V. From (3.13) and the definitions of b and #, we see that V *Xw is given through a simple calculation by (3.14)
V*XW(Y) = Xw(Y) - w (VXY) .
3. ENERGY OF MAPS AND HARMONIC MAPS
92
Consequently, we may accept (3.14) as a definition for V *Xw. Rewriting (3.14), we obtain
Xw(Y) = V*Xw(Y) + w(VXY).
This equation explains that the connection V in TM and the connection V* in TM* are in a mutually dual relationship. From the definitions of g* and V* and from (3.10), we can readily verify that, given X E r(TM) and w, 0 E r(TM*), the equation (3.15)
Xg*(w, 0) = g* (V* XW, 0) + g* (w, V*X0)
holds. Namely, we see that V* is a connection compatible with the fiber metric in TM*. Let U be a coordinate neighborhood in M and let (x') denote a local coordinate system in U. Then we see that (3.14) yields in
(3.16)
V * .-
dxk
= - E r 3 dx1,
1 < i, k < m.
j=1
We note that the connection coefficients of the connection V* induced in TM* from V are given as negative the connection coefficients of V. Let K E r(TM® ® TM*) be a tensor field of type (0, 2) and let L E I'(TM ®TM) be a tensor field of type (2, 0). As a generalization of the above, we can define, by covariantly differentiating L and K,
a tensor field V*K E r(TM* 0 TM* 0 TM*) of type (0,3) and a tensor field VL E r(TM* 9 TM (9 TM), respectively, by (317)
V*K(X,Y, Z) = XK(Y, Z) - K (VXY, Z) - K (Y, V Z) , VL(X, w, 0) = XL(w, 0) - L (V*xw, 0) - L (w,V*X9) .
Using these definitions, if we define K and L, respectively, by
V*xK(X,Y) =V*K(X,Y,Z),
VxL(w,0) =VL(X,w,0),
we can easily verify that the covariant derivatives V * x K and V x L
of K and L in X satisfy the same rules of calculation as V xY. In particular, if V * K = 0 and V L = 0, K and L, respectively, are called parallel tensor fields with respect to the connections V* and V. For example, g and g* are tensor fields of type (0, 2) and (2, 0), respectively. Following (3.17), a calculation of their covariant differentials V*g and Vg*, from (3.10) and (3.15), yields
V*g = 0, Vg*.
3.2. TENSION FIELDS
93
Namely, g is a parallel tensor field of type (0, 2) with respect to V*, and g* is a parallel tensor field of type (2, 0) with respect to V. In other words, that the Levi-Civita connection and g are compatible is nothing but that g and g* are parallel tensor fields with respect to V* and V, respectively. Given a local coordinate system (x') in a local coordinate neighborhood U of M, we express the components of the type (0, 3) for field V*g and the type (2, 1) tensor field Vg' by °igjk and Vigjk, respectively. Namely, V i9Jk and Dig1 k, respectively, are C°° functions in U defined by
(k) a
°igjk=V`y` Vi9'jk = Vg'
g j,b
xk =V
Ole
= V 8m' g! (dxi, dxk).
From (3.12), (3.16) and (3.17), simple computations yield that V,gjk and Vigjk are given, respectively, by V1gjk =
gjk
axi
m
-
m
i jglk 1=1
(3.18)
°igjk = axk + J
m
- E tlk t
1=1
m
j+ilgik +
1=11
I'ikl911. 1L=1
Consequently, we may regard (3.18) as definitions for Vigjk and
Vigjk. That g and g' are parallel with respect to V' and V, respectively, means (3.19)
Vigjk = 0,
1 < i, j, k < m.
Vigjk = 0,
We note that from (3.18), these equations are nothing but the componentwise expressions of (3.10) and (3.15) in the local coordinate system.
We can uniquely define a connection 'V in the vector bundle u_ I TN induced from the tangent bundle TN of N by a C° map
u : M -+ N from M into N (see Exercise 3.2 at the end of this chapter). In fact, let U and V be coordinate neighborhoods of M and N such that u(U) C V, and (x') and (y*) denote local coordinate systems in U and V, respectively. For each I < a < n, (3.20)
(&Yaoul(x)=
}
'6y" UW
,
3. ENERGY OF MAPS AND HARMONIC MAPS
94
defines a C°° cross section of u-1TN over U. At each point z E U,
(r° u}) (x)' ... ,
(ayn o u
) (x) }
gives rise to a base for the fiber TT()N of u-'TN over x. Then we can define a covariant differentiation ' V as such satisfying, for each
1
l
I0 8 88(r
o
u} (x)
Qdu.((
)/
If we represent by 1''#,y the connection coefficients of V' in (ya), a simple calculation yields, from (3.20) and (3.21),
` a.1 By-Y
/
n
NE a=1
C
(0=1
Namely, 'V is nothing but a linear connection in u-'TN whose connection coefficients are given by
l
u`h E r(u-'TN' (9 u`1TN`) defines a fiber metric in u-1TN'. As in (3.17), we see that the induced connection 'V defines the covariant derivative of u'h
'Vu'h E r(TM" 0 u-1TN' (9 u`'TN'). We express the components of 'Vu* h in terms of local coordinate
systems (as) and W) in M and N, respectively, by
Vshafl(u) = 'Qu*h G2'' 5i o u, a
49
oy
ou
ou,ou 19
3.2. TENSION FIELDS
95
From the definition of the Levi-Civita connection V' and from h being compatible in N, we can easily verify that (3.22)
V{ha,,o(u) = 0,
1 < i < m,
1 < o, 8 < n.
In other words, the induced connection 'V is compatible with the fiber metric u* h.
Next, we define a connection V in the tensor product TM' u-1TN using the connection V' in TM* and the induced connection ' V in u 1 TN as follows. First, we note that a C°° section of TM' ®u1TN, in general, is given as a linear combination of tensor products of a Ca° section of TM' and a C°° section of u-1TN. Consequently, in order to define the desired connection V, it suffices to define a covariant differential
V(w ® W) E r(TM' 0 TM* ® ul1TN) for the section w ® W E r(TM' ® u-'TN) obtained as the tensor product of w E r(TM') and W E r(u`1TN). Using V* and'V, we define V(w 0 W) by (3.23)
V (w ®W) = (V'w) ®W + w ® ('VW).
Under the above definition, the covariant derivative Vx(w ® W) of we W in X is given by
V( w ®W) = (V'Xw) ®W + w ®('VxW), X E r(TM). It is easy to verify that the covariant derivative satisfies the same computational rules as the coraviant derivative V Y in TM. Hence, (3.23) gives rise to a connection
V : r(TM' ®u-1TN) --+ r(TM* ®TM* 0 u-1TN) in the tensor product TM* ®u -1 T N. It is almost trivial from the definition that this connection V is
compatible with the fiber metric (, ) in TM* 0 u-1TN. In fact, if we note that the components of ( , ) in the respective local coordinate systems (x*) and (y') in M and N are given by g3kh,,,0(u)'s , it is easy to see, from (3.23), (3.19) and (3.22), that the covariant derivative of (, ) regarded as a C°° section of (TM' u-1TN)' ® (TM' (& u 1TN)' TM ® TM ® u-'TN* 0 u-1TN' satisfies the following equation (3.24)
Vsg?kh.,6(u)=O,
1 < i, j, k < m,
1
As was seen in the last section, the differential of a COO map u E C°° (M, N) from M into N defines a C°° section du E r(TM' ® u-ITN) in the vector bundle TM' 0 u-ITN. Considering
3. ENERGY OF MAPS AND HARMONIC MAPS
94
the covariant differential of du by the connection V in TM we obtain a C°° section
-1TN,
Vdu E r(TM® ® TM* (& u-'TN)
in the vector bundle TM' (9 TM* ®u-1TN. This Vdu is called the second fundamental form of the CO° map u. Since TM® ® TM® ® u`1TN is isomorphic to the bundle of hommmcphianm Hom(TM (9 TM, u-1TN) as a vector bundle, Vdu is nothing but a u-1TN valued tensor field of type (0, 2) over M. Here, we remark the following.
Ln4mA 3.3. Given u E CO°(M, N) and X, Y E r(TM), we have
Vdu(X, Y) ='Vxdu(Y) - du(VXY).
(3.25)
PROOF. Since du E r(TM* 0 u-1TN) is given as a linear combination of sections in the vector bundles TM* and u 1 T N, it suffices to consider the tensor product w 0 W E r(TM* ® u-ITN) of w E r(TM') and W E r(u-1TN). From definition (3.23) of the ounnection V in TM' ®u-1TN and (3.14), we can readily verify that
(V(w0W))(X,Y) = (V"xw®W +w0'VvW)(Y) (Xw(Y) - w (VXY)) ® W + w(Y) ®'VxW ='Vx((w ®W)(Y)) - (w ®W)(VXY). Hence, we have
Vdu(X,Y) = (Vx) (Y) ='Vxdu(Y) - du(VxY). D
La 3.3 Implies that we may regard (3.25) as a definition of Vdu. On the other hand, if we express, in local coordinate (,Ti) aM ()inMandN,duEr(TM*®u-1TN)andVduE r(TM* 0 TM' 0 u -'TN), respectively, as in
n
d=1 0=1
C
a
fto
0 U,
m rr Vdu = L > V, Vjua ' dx' 0 dx n
(3.27)
i4=1 a-- I
we have the following.
®A4.Q 0 u,
3.2. TENSION FIELDS
97
LEMMA 3.4. For each 1 < 1, j < m, 1 < holds:
V'Vju«
N°
["b
(u«
Lr
n, the following
n+
'
a
OUP 19U7
L r01(u) ax' az - a {a2y - k=1 ? ajk +0,7=1
PROOF. From the definition of Vdu and (3.27), we get
u«da
e$$
ou.
j=ia=1 On the other hand, from the definition of the induced connection 'V, (3.16) and (3.26), we get m
V
du =V
axjdxj ®
0U
j=1 a=1
a2u«
a;-axj
j=1
d
®
(8)ay,
a o + au« ftj V
ou+ auaj dxj O'V 15x-
in
n
a2u«
= E1 0-1 {l
m
k &ua
m
E r'7 a2;k + f3,ry=1 k=1
dsj
awt
iv out
a)
auo &7 (u 5T, axj
dx' ®..a ou. The proof follows from this.
0
We see, in particular, that Lemma 3.4 and the symmetric property of the induced connection imply V'V1ua = VjV'u«. In other words, the second fundamental form of a C°° map u is a symmetric u-1TN valued tensor field of type (0, 2) in M. Namely, we get the following.
COROLLARY 3.5. Given U E C°° (M, N) and X, Y E F(TM),
Vdu(X,Y) = Vdu(Y, X) holds.
Let lei,...
, en j be an orthonormal basis for the tangent space
TAM of M at each point x E M. For the second fundamental form
98
3. ENERGY OF MAPS AND HARMONIC MAPS
Vdu of a C°° map u, as readily seen, m
traceVdu(x) _
Vdu(x)(ej, e=) i=1
is uniquely determined independently of the choice of the orthonormal
basis, and gives rise to a C°O section of the vector bundle u-1 TN. If we express Vdu as in (3.27), we see that this follows from that trace Vdu is given by n
(3.28)
m
trace Vdu =
g{' ViV fu" er-Z
ij=1
ou
)
10
in terms of the local coordinate systems (x') and (ye) of M and N, reqwtively. DEFINITION 3.6. Given a C° map u E C110 (M, N),
r(u) = trace Vdu E F(u-' TN) is called the tension field of u.
Finally, we make a remark regarding the induced connection 'V
in the induced bundle u 'TN. LEMMA 3.7. Given u E C°°(M, N) and X, Y E r(TM), we have
'Vxdu(Y) -' V ydu(X) = du([X, 11). PROOF. From Lemma 3.3 and Corollary 3.5 we get, noting (3.11),
'Vxdu(Y) -' Vydu(X) = du(VxY - VyX) = -du([X,Yl).
In the above arguments, we employed the Levi-Civita connections
V V' of M and N, respectively, the connection V' in TM* induced from V, the induced connection'V in u-'TN determined by V. Furthermore, we also used the connection V being compatible with the natural fiber metric (, ) in TM` 0 u 'TN. In what follows we will denote all of these connections by the same symbol V, unless there is a fear of confusion. It is obviously more convenient to uniformly employ the same symbol.
3.3. THE FIRST VARIATION FORMULA
99
3.3. The first variation formula Under the preparations in the previous sections, we obtain the first variation formula for the energy E(u) of a Cd° map u E C°°(M, N). Let (M, g) be a compact rn-dimensional Riemannian manifold, (N, h) an n-dimensional Riemannian manifold, and I = (-E, e) (e >
0) an open interval in the real line R. Given a C°° function u E C°0 (M, N), a CO0 map F : M x I - N is called a C°° variation or a smooth variation of F if
F(x, 0) = u(x),
(3.29)
x E M,
is satisfied. Given a variation as defined above, set
x E M, t E I.
Ut(x) = F(x, t),
Then ut gives rise to a CO° map ut : M -> N, and uo = u holds from (3.29). Furthermore, since F(x, t) is of C°° in t, the family of maps jut I t E I} C C°°(M, N) obtained as above defines a smooth variation of the given map u = uo in the space of maps C°° (M, N). In what follows, we denote by F = jut } t E I a C°° variation of u E C°° (M, N)
for the sake of simplicity. When a C°° variation F = {ut}tEI of N u E C°° (M, N) is given, at each x E M, ut (x) = F(x, t) : I defines a C°° curve in N, passing through u(x) at t = 0. Consequently, the set of the tangent vectors to these curves at t = 0, denoted by
ut(x) =
V (t) = t=o
&
(x, 0) E Tu(s)N,
x E M,
defines a C°C section V E r(u-'TN) of the induced bundle u -'TN. In other words, V (x) defines a C°° vector field in N along the map u. The tangent space T(x,t) M x I of the product manifold M x I at a point (x, t) is naturally identified with the direct sum TxM ED TtI of the tangent spaces TM and TtI. In this setting, V is nothing but the vector field given by V (X) = dF(,,o)
0,
j
0
j,
(x, 0) E M X I.
This vector field V E r(u-'TN) is called the variational vector field of the map u.
Given a C°° map u E C°°(M, N), let V E r(u ITN) denote a C°° vector field in the induced bundle u-'TN. Then, for a sufficiently
3. ENERGY OF MAPS AND HARMONIC MAPS
100
F(x. t)
FIGURE 3.1. A variation of a map and its variational vector field
small e > 0, a C'°° variation F = {ut}tEj of u can be defined by F(x, t) = exp,3(.) (tV (x)), (x, t) E M x I. With this variation, dI
V (X) _
ut (x),
x E M,
t=0
holds. Hence, the set r(u-'TN) of the C°° sections of the induced bundle u ITN is nothing but the space of the variational vector fields along U.
Given a C°° variation F = {ut}tEl, we investigate the change of the energy functional E. Since F is a C°° variation, the energy
E(ut)
2
Jr I dut I2dpg
becomes a CO° filnction in t for each ut E C°Q(M,N). Regarding the first variation of E(ut), we have
THEoiu M 3.8 (The first variation formula). Let F = {ut}tEl be a CO° variation of a C°° map u E C°° (M, N). Then
5E(ut)I
(3.30)
where V = d
r
t_o
M = -J (V,T(u))dp9,
ut is a variation vector field of u, and T(u) is the
I
too
tension field of u. (,) is the natural fiber metric in the induced bundle
u-'TN.
3.3. THE FIRST VARIATION FORMULA
101
PROOF. Let F(x,t) = ut(x) : M x I -+ N be a COO map defining the C°° variation of u. Consider now the vector bundle T(M x I)* ®F-1TN over M x I. As was seen in the previous section, T(M x I)* 0 F-'TN admits a natural fiber metric (, ) and the standard connection V compatible with the metric. Under the Tx M®Tt I, we denote the covari-
natural identification T(x,t) (M x I)
ant differentiation with respect to the connection V in the directions (8/8x4, 0) E T(x,t) (M x I) and (0, d/dt) E Tx,t (M x I), respectively, by
vt = V(o,d/dt).
Vi = v(a/ax+,0),
As before, denote by (xi ), (ya) the local coordinate systems in M and N, and by gij, h,,.,g the components of the Riemannian metrics g, h, respectively. Since V is a connection compatible with the fiber metric (,) , for each 1 < i, j, k < m, 1 < a, < n, from (3.24) follows (3.31)
vigjkha#(ut) = 0,
Vtg3khafl(ut) = 0. Since from the definition, E(ut) is given by
E(ut) =
2
f
M
;dµ9 exi
L is=1 a,Q=1
noting (3.31),
dtE(ut) t=0 =
2
f
M
n
a
9s'ha18(ut) a 4 8zj
bl
(ij=1 a,4=1
m
dp9
t=0
a
n
&i (7xj ) 1t=0
M
On the other hand, applying Lemma 3.7 to the C°° map F and the vector fields (8/8x4, 0), (0, d/dt) over M x I, by noting that [(0, d/dt), (8/8x4, 0)] = 0, we get V(o,d/dt)dut
\ \a '' 0)) -- V(a/ax+,o)dut `(0' 4))
=dut(
(0
dt)'(
=0.
i'0))
Consequently, for each 1 < i < m, 1 < a < n, we get (3.32)
vt Our = 8xt
via
8t
3. ENERGY OF MAPS AND HARMONIC MAPS
102
If we write a variational vector field of u as n
V=1:
a=1 Va
ou,
=a
noting V,
we get It=01
m
d'IE(ut)
=
a
IEE
94'h(u)©i
M j j_1
= fm
M
n-+
E
L
g'jha0(u)ViVQ
i,j=1 a,$=1
att t _ o
dig
(quo W
d1A9
= J(VVdu)d9. Here (, ) on the right hand side denote the natural fiber metric in the vector bundle TM* 0 u-'TN. Hence, from the following lemma, we obtain (3.30) as the first variation formula of E(u ).
0
LEMMA 3.9. Given u E COO (M, N) and V E r(u-'TN), we have fm
(VV, du)djas =
" fkt
(V,
r(u))dµ9.
Here, the symbol (, ) on the left and right hand sides, respectively, represent the natural fiber retrace in T M* 0 u-1 TN and u-1 TN. PROOF. Let X be a C°O vector field over M given by m
m
X=E i=1
n
E E E 9'hag(u)V i=1 j=1 0,6=1
'
Denote the covariant differentiation of X by in
VX = E ViXj dxt ®tax ij=1 i VEX' (see Exercise 3.3 The divergence of X is given by divX = at the end of this chapter). Then noting the identity
V(V0du)=V 0du+V0Vdu
3.4. HARMONIC MAPS
103
together with (3.24), (3.26), (3.27), and (3.28), we get from a simple calculation m n auls
E ij=1 a,3=1
divX = E
M
9'3h«p(u)ViyVa
a
.i
n
+ E E giiha0(u)VaV,Vjuu3 i,j=1 a,J=1
=(VV, du) + (V, r(u)).
Green's theorem (see Exercise 3.4 at the end of this chapter) implies JM
divX dµ9 = 0;
hence, we get the desired result.
From the first variation formula of Theorem 3.8, we get the following-COROLLARY
3.10. Given u E C1 (M, N), a necessary and sufficient condition for the first variation of E(ut) of an arbitrary COQ
variation F = tvt}tEJ to satisfy dtE(ut)t_o = 0 is for the tension field of u to identically vanish, i.e., r(u) L 0. PROOF. It suffices to note that we can take any CO° section V E
I'u-'TN as the variation vector field in the first variation formula (3.30).
Corollary 3.10 indicates that a C°° map u E COO (M, N) with r(u) = 0 is a critical point of the energy functional E : C°°(M, N) -R.
As seen from the proofs of Theorem 3.8 and Lemma 3.9, we used
the connection V in the vector bundle TM* 0 u-1TN and Green's theorem regarding the divergence of a vector field. However, the basic
idea for the proof of the first variation formula for the energy E(u) of a map is essentially the same as in the case of the first variation formula for the energy E(c) of a curve.
3.4. Harmonic maps From the first variation formula for the energy of a map obtained in §3.3, we have seen that the critical points of the energy functional are given by COO functions whose tension fields vanish identically.
3. ENERGY OF MAPS AND HARMONIC MAPS
104
Such maps are generally called harmonic maps. In this section, we discuss the definition of harmonic maps and some examples. In what follows, (M, g) and (N, h), respectively, denote connected Riemannian manifolds of dimensions m and n, and denote by u : M N a C°D map from M into N. We begin with a rigorous definition of a harmonic map.
DEFINITION 3.11. A C°° map u E C°°(M, N) is called a harmonic map if its tension field r(u) is identically 0; namely,
r(u) = trace Vdu - 0
(3.33)
holds in M. (3.33) is called the equation for harmonic maps.
When M is compact, a harmonic map u is nothing but a map which is a critical point of the energy function E : C°° (M, N) - R. In fact, as was seen in Corollary 3.10, the map u E COO (M, N) being harmonic means that
=0 dtE(ut) holds for an arbitrary C°° variation F = {ut}tEl of u. From Lemma 3.4 and (3.28), we see the following, regarding the equation for harmonic maps. Let (xs) and (y') denote local coordinate systems in M and N, respectively. With these local coordinate systems, we express the map u by
u(x) = (u1(x1,... ,xm),... ,u"(x1,... ,2m)) = (U12 (x')), and denote the tension field r(u) of u by n
r(u)°i / o u E r(u 1TN).
r(u) _ a=1
Then r(u)° is given by
r(u)
m
02ua °C
=1 9 =,j
= Lu°1 +
=j 8xk
8x'8xi
Lr L.
i,.1=1 Q,'T=1
8U.*
r
k
k=1
=jrja
9
Q-f (u)
+Q,y=1
1a ply
(u
)
8U,0 OU7
ax' 8xj
OuQ ou7
8x' axj
Here, r'k and r'Qy, respectively, represent the connection coefficients of the Levi-Civita connections in M and N, and A is the Laplace
3.4. HARMONIC MAPS
105
operator in M (see Exercise 3.3 at the end of this chapter). Consequently, equation (3.33) for harmonic maps is given, in the local coordinate systems in M and N, as m
(3.34)
A0 +
n
7 gjjr'G
(u)
5T Xi
=0, 1 < a < n.
From (3.34), using the local coordinate systems, we see that equar tion (3.33) for harmonic maps is a system of second order quasi-linear elliptic differential equations. The nonlinear parts are second degree polynomials of the first order partial derivatives of ua. If we set 9aJrn
$'r (u) 1,J=1 0,'Y= I.
ou auk axs 92;j '
and decompose (3.33) formally as
au' + r(u)(du, du)a = 0,
1 < a < n,
we understand the characteristics of the equation for harmonic maps better. Note, however, that the above decomposition is invariant only in coordinate transformations of M, but not in those of N. Examples of harmonic maps appear in various problems in differential geometry. In what follows, we take up some of these examples. EXAMPLE 3.12 (Constant maps and identity maps). The
sun-
pleat example of a harmonic leap is a constant map. In fact, let u : M -* N be a constant map. Then there is a point q E N such that
u(x)=q, xEM.
The derivative du of u is identically 0; hence, the tension field of u is also 0. Especially, if M is compact, the energy E(u) of a constant map u becomes 0. The converse also holds true. Namely, the constant maps are nothing but the maps that give the absolute minimum value of the energy functional E : C°° (M, N) -; R. Now, if we assume u : M - M to be the identity map
u(x) = x, x E M, then dug : TAM -i TM is the identity map for each x E M. Hence, 7(u) = 0, and u is a harmonic map. EXAMPLE 3.13 (Harmonic functions). Assume (N, h) to be the one-dimensional Euclidean space E. Namely, E is the real line R with the Euclidean inner product. A C°° map from M into N is nothing
100
3. ENERGY OF MAPS AND HARMONIC MAPS
but a COO function u : M -p R. As was seen in Example 1.14, r= 0 in the equation (3.34) for harmonic maps. Thus, u being harmonic is equivalent to Au = 0 (Laplace equation) regarding the Laplace operator of Al. Consequently, a harmonic map
u : M R is nothing but a harmonic funchon in Al. Fbr a given function u : M -i R, the energy E(u) defined by (3.9) is called the Dirichlet integral of u. In general, it is well known as the Diriclrlet principle in analysis that the Laplace equation Lu = 0 is the Euler equation, namely, the first variation formula for the Dirichlet integral. In this context, we may consider the equation r(u) = 0 for harmonic maps as a generalization of the Laplace equation for functions in the case of maps between Riemannian manifolds. This is the reason why a map u with r(u) = 0 is called a harmonic map.
EXAMPLE 3.14 (Geodesics). Next, we assume (M, g) to be the one-dimensional Euclidean space E. In this setting, a C0° map fircm Al into N defines a C°'° curve u : J --s N. Considering t as a coor-
dinate function in R, we get i Jk = 0, as in Example 3.13. Hence, the equation (3.34) for harmonic maps reduces to the equation for geodesics in N
Consequently, a harmonic u : R -+ N is a geodesic in N, and t is nothing but an affine parameter for u. On the other hand, if we assume (M, g) to be the one-dimensional
unit sphere S1 C E2, a COO map u : St - N determines a smooth loop in N. A harmonic map u : Sl -+ N becomes a closed geodesic in N. The energy E(u) of u is precisely the energy of curves defined in §1.1. The equation for harmonic maps r(u) = 0, therefore, can be regarded as a generalization of the equation for geodesics in the case of maps between Riemannian manifolds. The next example is also best understood if it is viewed from the same aspect.
EXAMPLE 3.15 (Minimal submanifolds). Let u : M - N be a C' °° immersion of a Riemannian manifold (M,g) into another (N, h). As was seen in Example 1.2, theme is an induced metric u`h in M.
3.4. HARMONIC MAPS
107
When especially g = u'h, namely, at each x E M, gx(v, w) = hu(x)(dux(V), dux(w)),
v, w E TIM,
holds, we call u an isometric immersion from M into N. Given a C°° immersion u : Al -* N, we may define the normal vector bundle of M
TM-i = U dux(T2M)1 xEM
using an orthogonal decomposition
Tu(x)N = duI(TTM) @ duI(TM)l, x E M, of the tangent space Tu(I)N of N at u(x) with respect to k(x). Then, as seen easily from Lemma 3.3 and the definition (3.31) of the induced connection, we see that the second fundamental form
©duEI'(TM`®TM'(& u-'TN) of u defined in §3.2 coincides with the ordinary second fundamental form
AEF(TM*9, TM`0TM1) when M is regarded as a Riemannian submanifold of N. In fact, since
u is a local imbedding, by identifying X E r(TM) with du(X) E r(u-'TN) about each point x E M, (3.25) represents the identity (©'XY)(x) _ (V xY)(x) + A(X, Y)(x), x E M, expressing the decomposition of the covariant derivative V'XY into
the TM and TM' components. We see that the tension field r(u) of u coincides, up to a constant, with the mean curvature vector field H = trace A/m of the Riemannian submanifold M of N.
In general, when the mean curvature is identically 0, we call the Riemannian submanifold M a minimal submanifold of N. Consequently, an isometric immersion u : M -+ N being harmonic is equivalent to M being a minimal submanifold of N. In particular, if u; M -+ N is an isometric diffeomorphism, Vdu = 0; hence, u is a harmonic map (see Exercise 2.8 in Chapter 2). EXAMPLE 3.16 (Riemannian submersions). We consider a some-
what opposite situation to Example 3.15. In general, a C°° map u : M -+ N is called a submersion if the derivative du, : TxM --+ Ta(e) N is surjective at each x E M. It is clear from the definition that
m < n holds between the dimensions of M and N if there is a submersion from M onto N. Also the inverse function theorem implies
3. ENERGY OF MAPS AND HARMONIC MAPS
108
that u-1(u(x)) is an (m - n)-dimensional C°° submanifold of M for each x E M. We call u-i (u(x)) a fiber passing through x E M. Let u: M --+ N be a submersion from a Riemannian manifold (M, g) onto another (N, h). At each x E M, denote by V,: the tangent space to the fiber u-1(u(x)) at x. If we orthogonally decompose the tangent space T=M at x with respect gx, we have
T1M=V, (DHH, xEM. VS and H. are called, respectively, the vertical component and the horizontal component at x. In particular, if the restriction duz IHx : H. - Tu(r) N of the derivative du,, of u to H= at each x gives rise to an isometric isomorphism, u is called a Riemannian submersion. Given a vector field x E r(TN) over N, a vector field X E I'(TM) in M satisfying X E H=,
dux(. (x)) = X(u(x)), x E M
is called the horizontal lift of X.
Now let u : M -- N be a Riemannian submersion. Then the following proposition implies that u being harmonic is equivalent to each fiber u-1(u(x)) being a minimal submanifold of M.
PRoPoSmoN 3.17. Let u : M - N be a Riemannian submersion from a Riemannian manifold (M, g) onto another Riemannian manifold (N, h). Then a necessary and sufficient condition for u to be a harmonic map is that the fiber u-1(u (x)) passing through x for each x E M is minimal as a Riemannian submanifold of N. PROOF. Given x E M, denote by (yn) a local coordinate system
of N about u(x). Applying the Schmidt orthonormalization to the natural frame {8/8yP } determined by the coordinate system at each
point x E M, we get a set lei
I
1 < i < n} of C° orthonormal
vector fields around y. At each point y, {e'i(y) I 1 < i < n} forms an orthonormal base for T.N. Such a set {e'i} as defined above is called an orthonormal frame field. Denote by {e1i... , e,,} the horizontal lift of {e'1, ... , e,,), and e x t e n d it to a n orthonormal {ei , ... , en, eni1, ... , c. } around x. Then, from the definition and Lemma 3.3, the tension field r(u) of u is given by m
T(u)
M
Vd21(ei, ei) i=1
(T.,du(ez)
_ i,=1
- du(Ve1ei)}.
109
3.4. HARMONIC MAPS
Noting the definition of the induced connection 'V and Exercise 3.7 at the end of this chapter, we see that V.,du(ej) = V'e',e'i = du(Vepei)
holds for each 1 < i < n. For n + 1 < i < m, e -j
to the
vertical component at each point. Hence, 'Ve,du(ei) = 0 follows. Consequently, we obtain
t(u) _ - > du(Veiet) = -du
Veiei)
.
s=n+1
From this follows that a necessary and sufficient condition for r(u) = 0 is given by m
V,e, ei E Vz = (du=)-' (0),
x E M.
i=n+1
On the other hand, if we regard the fiber u-1(u(x)) passing through x as a Riemannian submanifold in M, we see that m
trace A(x) = the H. component of E Ve; ei (x) in+1 holds from the definition of the second fundamental form A. Consequently, we see that each fiber u' 1(u(0)) must be a minimal subman0 ifold of N in order for u to be a harmonic map.
Finally, we look into the case of complex manifolds. Here, we presume that the reader is acquainted with the definition and fundamental properties of Kahlerian manifolds (see Kobayashi 1261 for those). EXAMPLE 3.18 (Holomorphic maps). Let (M, g) and (N, h) be Ki hlerian manifolds. Denote by m and n the complex dimensions of M and N, respectively. Let (zi) = (z1, ... , zm) and (wi) = (w1, ... , wn) be local complex coordinates systems of M and N, respectively. Then the Kahler metrics g and h in M and N are expressed as
in
g=2
m
g,jdz`dzj, iJ=1
h=2
h,,,Rdzadz"0. a,j=1
If u : M - N denote a C° map from M into N, and if we express u with respect to the local coordinate systems as u(z) = (u1(z', ... , zm), ... , un(z1, ... , zn=)) _ W (z`)),
3. ENERGY OF MAPS AND HARMONIC MAPS
110
we can express the equation of harmonic maps (3.34) for each a as m
gt-
l
192u°i
aziazi
A
ra, A
aua
y
/
+ B,C
19UC r BC(U) suB azi az1
= 0.
Here, the index A runs through 11,... , m,1, ... , rn} and B, C run through { 1, ... , n, 1, ... , n}. 174 and r'BC are, respectively, the con-
nection coefficients of the Levi-Civita connections of V and V' with respect to the local complex coordinate systems.
Since M is a Kahlerian manifold, for any A, r = 0 holds. Sim ilarly, 'rBc = 0 holds for all T BC's other than I ,, since N is also a Ki hlerian manifold. Consequently, with respect the complex coordinate systems, the equation of harmonic maps is given by (3.35) in g..
=,i=1
82ua az=azi + ,y=1
a
OUP oy
u
az Ozj
= 0,
1:5 a!5 n.
Let J denote the almost complex structures in M and N. A C°° map u : M --+ N satisfying J o du = du o J is called a holomorphic map, and similarly, an antiholomorphic map if J o du = -du o J. From the definition, u being holomorphic is equivalent to each ua being a holomorphic function of (z'), namely, aua /az' = 0. Similarly, being antiholomorphic is equivalent to each uA being antiholomorphic,
namely, &a/az* = 0. Hence, we see from (3.35) that a holomorphic or antiholomorphic u : M - N from a Kahlerian manifold M into a Kahlerian manifold N is a harmonic map.
3.5. The second variation formula In this section we obtain the second variation formula regarding the variational problem of the energy of maps. As in the first variation formula in §3.4, let (M, g) be a compact m-dimensional Riemannian manifold, (N, h) an n-dimensional Rie-
mannian manifold, and I = (-e, e) an open interval. Assume that u E C°°(M, N) is a harmonic map from M into N. Let F = {ut}tEl be a COQ variation of u. First, we note the following. Given a variation vector field
dl V= dt _ ut(x) E F(u-'TN)
3.5. THE SECOND VARIATION FORMULA
111
of u, we have the covariant differential VV E I'(TM* (9 u-iTN) with respect to the induced connection in u-1 TN. Furthermore, if we consider the covariant derivative of VV with respect to the connection
V in TM` f& u-1 TN, we obtain VVV E r(TM* e TM' ®u-'TN) as the second order covariant derivative of V. If we express VVV, in local coordinate systems (x`) and (y)) in M and N, respectively, as
vvv = E E vivjv° dxi®dx'®iou, i,j=1 a,3=1
we obtain as the trace of V V V (3.36)
m
m
trace VVV = E E 9ijViVjVo "=1
0uE
r(U-1TN).
i,j=1
On the other hand, given a variation vector field V and the curvature tensor RN of N, if we set
(RN(V,du)du)(x)(v,w) = RN(V(r),du,.(v))du1(w), v,w E TIM, RN(V, du)du gives rise to a C°° section of the vector bundle TM* 0 TM* 0 u-1TN. From this, we define, as the trace of RN (V, du)du, trace RN (V, du)du (3.37)
M
q =jRN
V, du
( &Xi
du
OXJ
E r(u-'TN).
Under the above preparations, we get the following theorem. The idea of the proof for the second variation formula regarding the energy of maps is a sentially the same as in the case of the second variation formula for the energy of curves in Theorem 2.16.
THEOREM 3.19 (The second variation formula). Let u E C°°(M, N) be a harmonic map and let F = {ti }=e, be a C0° variation of U. Then r d2 _ = fM (V, trace (VVV + RN (V, du)du))dicg d 2 E(ua) t=o holds. Here, V and RN are, respectively, the variation vector of u and
the curvature tensor of N, and (, ) denotes the natural fiber metric in the induced bundle u-1T N. Also trace (V V V + RN (V, du)du) is the C°O section of u-1TN defined in (3.36) and (3.37).
3. ENERGY OF MAPS AND HARMONIC MAPS
112
PROOF. As in Theorem 3.8 (the first variation formula), we consider the vector bundle T (M x I) - & F-' T N and the connection V compatible with the natural fiber metric (, in it. First of all, from Theorem 3.8 and Lemma 3.9, we note that
dt
E(ut) _ -
,trace Qdut dµ9
J (
holds for each t E I. By differentiating this equation in t, we get
E(ut)
_-
(3.38)
trace Vdik dµ9
dt
Jrht
/plc out , trace Vdu: }dµ9 /`
(Vttrace Vduc )d
9.
!JJ
V6 represents the covariant differentiation with respect to V in the direction of (0, d/dt) E T(x,c) (M x I). Noting the definition of the curva-
ture tensor R' of N and ((0, d/dt), (8/8x', 0)] = 0, we see, regarding the covariant differentiation with respect to the induced connection
in F-1TN, that
V (O,d/d)V
((i.
(8f 0x1 ,o)du t
= Qi/& ,o)V (o,d/dt)d
+R" holds.
(dttt
a
((i,
0
((0)) ,du((o))) du`(o))
On the other hand, noting Lemma 3.7 and ((0, d/dt),
(8/8x`, 0)] = 0, we get 0(O,d/dt)d
((azi
((o))
=
((os
d ))
_
3.6. THE SECOND VARIATION FORMULA.
113
Hence, we get
V(o,d/dal V(8/e=d,o)dtct ((
_
7 r o/
,o)dtt ((0,d/dt))
+ RN dut
((o))
,
dut
((,0))) dut ((o)). 8x= 8xl
From this, we see readily that (3.39)
Vt trace Vdut = trace
(vv
dutl
+ RN
J
.
If we set t = 0 in (3.38), since u = u0 is a harmonic map, we get r(u) = trace Vdu = 0;
implying that the first term on the right hand side equals 0. On the
other hand, since 'v = att ! _ by definition, using (3.39), the second c-o
term of the right hand side can be rewritten as ,trace (VVV + RN(V, du)du))dµ9; hence, we obtain the desired result.
0
Theorem 3.19 implies that the second variation d2/dt2 E(ut)lt0 of the energy of the map is determined by the tensor product of the variation vector V along the harmonic map u and the curvature tensor RN of the Riemannian manifold N. As seen in §2.5, we were able to investigate the topological structures of Riemannian manifolds of positive curvature using the existence theorem for closed geodesics in compact Riemannian manifolds and the second variation formula for the energy E(c) of curves. In a similar manner, one can study the structures of Rienlannian manifolds using harmonic maps and the second variation formula for the energy of maps. For example, Micallef and Moore [13) recently proved "the topological sphere theorem" for the Riemannian manifolds of positive curvature under a pointwise pinching condition. Their proof utilizes the existing theorem for harmonic spheres regarding the second homotopy group Ira (M) by Sacks and Uhlenbeck [20] in combination with Morse theory on infinite-dimensional manifolds.
Siu and Yau [24], by applying a similar idea to Kir manifolds, solved the so-called the "Frankel conjecture". In their arguments, the
114
3. ENERGY OF MAPS AND HARMONIC MAPS
second variation formula has provided an important tool to measure stability of harmonic maps. For example, see Urakawa [31] for the relationships between the second variation formula and the stability of harmonic maps.
Summary 3.1 Definition of the energy E(u) of a C°° map it a Riemannian manifold M into another Riemannian manifold N. 3.2 A fiber metric and a compatible connection V are naturally defined on the vector bundle TM* ® u-1TN. Using this connection
V, the second fundamental form Vdu and the tension field T(u) _ trace Vdu of u are defined. 3.3 The first variation formula regarding the energy E(u) of a map u and a characterization of the harmonic maps as the critical points of the energy functional E. 3.4 The equation r(u) = 0 for the harmonic maps and examples of harmonic maps. 3.5 The second variation formula regarding the energy E(u) of
a map u.
Exercise 3.1 Let (M, g) be an m-dimensional Riemannian manifold, and let Co(M) denote the vector space of real valued continuous functions defined in M with compact support. Given a locally finite coordinate neighborhood systems {(Ua, 0a)}aEA, let {pa}aEA be a partition of unity subordinate to the open cover {Ua}aEA For a given f E Co(M), set (paf/det(g)) o 0a 1dZ.... d xx . Il9 = aEA
J
. (U-)
Here, (g a) is the matrix consisting of the components of g with respect to the local coordinate system (xa, ... , x,' n,). The integral on the right
hand side represents the Lebesgue integral of a continuous function with compact support defined in the open subset cbaUa in R" `. Then p r o v e that p (f) is uniquely determined independent of the choice of local coordinate systems {(Ua,la)}aEA and partitions of unity subordinate to {Ua }aEA and that µ9 gives a positive Radon measure in M. µ9 is called the standard measure induced from the Riemannian metric g in M. µ9 is also written by h, f dµ9, and called the integral of f over M.
no
EXERCISE
3.2 Let M and N be C1110 manifolds, and let u : M -- N a C°° map from M into N. Denote by u -1 TN the bundle over M induced by u from the tangent bundle TN of N. Given a C'O0 section Y E r(TN),
a C°° section Y a u E r(u-'TN). Prove that the linear connection v' on TN defines a unique linear connection 'V on u-'TN satisfying the following condition. Given a point v E TM and Y E r(TN), 'VvY 0 U
=
holds. 'V is called the induced connection by u in u'1TN from V. 3.3 Let (M, g) be an in-dimensional Riemaunian manifold, and denote by (xi) a local coordinate system of M in a local coordinate neighborhood U. Verify the following. (1) Given a C°° function f E C°° (M), a CoD vector field grad f =
(4f)# on M defined by
g$(grad f(x),v) = tVx(v), v E T.M is called the gradient of f. In U, grad f is given by
E (ii) M (M
g
Of
a=ii
$
(2) Given a C°° vector field X E I'(TM), a C"° function div X defined in M by
div X (x) = trace {v " V,X}, VETAM is called the divenjence of X. If ViXJ's represent the components of
the covariant differential VX of X = 9 1 Xj (x'), the following holds in U:
vixi -
div X = J=1
1
7
with respect to
8k
M FX det(9iJ ) k=1
(3) Given a C°° function f E C°°(M), a COD function A f defined
in M by
f = div grad f
is called the Laplacian of f. A is called the Laplace operator. If Vf V j f's denote the components of the type (0,2) C°Q tensor field
3. ENERGY OF MAPS AND HARMONIC MAPS
116
V ' determined by f with respect to (x3), m
m
&f = > 9'ipipjf = E m
V
4
ij=l
i.j=1
Ek
set
»i
a2 f g=i
axJ m
af
L.. 1 'i
k
k= 1
det(ggj) E 9 Of
k=1
k=1
holds in U.
3.4 (Green's Theorem) Let (M, g) be a compact Riemannian manifold. For any C°° vector field X E r(TM),
f div X f µ9 = 0 holds.
3.5 Let (M, g) be an m,-dimensional Riemannian manifold. Denote by V and V* the Levi-Civita connection of TM and the dual connection in TM* determined by V, respectively. Verify the following-
(1) Given a type (r, s) C°° tensor field T E r(TT'(M)), define the covariant differential VT E r(T,-+1(M)) of T as a generalization of (3.17) by
VT(X, X1, ... , XS, wl, ... ,
X ' T(X1:...
,
X81 w1 I
... ,
tttr)
$
,VxX,,... ,X87w1,... ,wr) i=1
r
- 1: T(X1i...
,
X5, w1, ...
,
wr).
j=1
Given X E r(TM), the covariant derivative VxT of T by X defined by
VTX(Xl, ... , XS, wl, ... , wr) = VT(X, X1, ... , X8, w1, ... , wr)
satisfies the same computational rules as the covariant derivative VxY of vector fields. It also holds that
Vx(T0U) =VxT0U+T®VxU for T E r(T$ (M)) and U E r(T9 (M)).
EXERCISE
117
(2) With respect to a local coordinate system (xi) in M, express respectively. the components of T and VT by T 11:; and VkT Then VkTJ1...J' _ a Tjl...j,
axk ,l. $
;,...ib
a
m
r
Tjl......
il...j.is +
rj6
L1 kt
il...j..-jr i1...{s
b=11=1
e=1 1=1
holds.
3.6 Let (M, g) be an rn-dimensional Riemannian manifold and let R E I'(T3 (M)) be the curvature tensor of M. Then the covariant differential VR E I'(T4 (M)) of R satisfies VR(X, Y, Z, V, w) + VR(Y, Z, X, V, w) + VR(Z,X,Y, V, W) = 0.
In other words, if we express it in a local coordinate system (xi) as m 1:
VR =
VR;kldxt ®dx'
dxk (& dx!
5X 7;'
i,j,k,j,r-1
the following (the second Bianchi identity) holds:
VjRjkj + V j Rrj +
0.
3.7 Let u : M -+ N be a Riemannian submersion from a Riemannian manifold (M, g) into another Riemannian manifold (N, h). Denote by V and V' the Levi-Civita connections of M and N, respectively. Let X, k E F(TM) be the horizontal lifts of X, Y E T(TN). Verify the following:
(1) g(X(z),Y(z)) = h.( )(X(lL(x)),Y(u(x))), z E M. (2) du(]X,Y]) = [X,Y]
(3) VY = (V'xYY +
2[X,Y]
Here, (V'XY) is the horizontal lift of V'XY and [X, Y]1 is the vertical component of [X, k] . 3.8 Regard R2 = C by identifying (x, y) with z = x + uE C. Express the unit three-dimensional sphere S3 C E4 by S3
=
I(z1,z2) E C2 I IziI2 + 122
12
= 11.
Define the Hopf map 46: S3 - S2 by z2) = (2ziz2, Iz1I2 - Iz2I2) E C x R, (zl, z2) E S3. Show that 0 is a harmonic map with respect to the Riemannian metrics on S3 and S2 induced from the standard Euclidean metric.
118
3. ENERGY OF MAPS AND HARMONIC MAPS
3.9 Given an immersion W : M -+ N from a Riemannian manifold (M, g) into another Riemannian manifold (N, h), cp is said to be conformal if cp*h = e2Ag for some C°° function p E COO(M). Assume
that M is of dimension 2. If yo : M - M is a conformal diffeomorphism, then E(u o cp) = E(V) for any C' map u E C°°(M, N), and furthermore, cp is a harmonic map.
3.10 Let (M, g) be a compact two-dimensional Riemannian manifold, and let (N, h) be an n-dimensional Riemannian manifold. Given a COO imbedding u : M -- N, the area of u is defined to be A(u) = IM dAu-h.
Prove that the inequality A(u) < E(u) holds, and that the equality holds if and only if u is conformal.
CHAPTER 4
Existence of Harmonic Maps In this chapter, we consider the existence problem of harmonic maps between compact Riemannian manifolds. Regarded as a generalization of the existence problem of closed geodesics discussed in Chapter 2, whether or not a given map can be deformed to a harmonic map may be ranked as one of the most fundamental questions among geometric variational problems. There is an effective technique called the heat flow method for deforming a given map to a harmonic map. In this chapter, we first explain the approach of the heat flow method. Then, using this method, we prove that any continuous map from a compact Riemannian manifold into a compact Riemannian manifold with nonpositive curvature can be free-homotopically deformed to a harmonic map. This theorem was first proved by FAls and Sampson in 1964. We give, in this chapter, a simpler proof than the original using the inverse function theorem in Banach spaces.
4.1. The heat flow method Let (M, g) and (N, h) be compact Riemannian manifolds of dimension in and n, respectively. Let f E C'30 (M, N) denote a C°° map from M into N. As one might guess, from the examples of harmonic maps seen in §3.4, whether or not f can be continuously deformed to a harmonic map u : M -* N from M into N is a fundamental problem in the study of harmonic maps. For example, in the case where (M, g) is a one-dimensional unit sphere S' C R2, a harmonic map u : S1 -+ N is a closed geodesic in N; hence, the problem is nothing but whether or not a given smooth loop can be deformed continuously
to a closed geodesic in N. In this case, as seen in Theorem 2.23, it is well known that any f E C'00 (S', N) can be deformed to either a constant map or a closed geodesic u : S' -+ N free-homotopic to f. 119
120
4. EXISTENCE OF HARMONIC MAPS
The objective in this chapter is to study the existence of harmonic maps between general compact Riemannian manifolds, and then to prove the following theorem due to Fells and Sampson.
THEOREM 4.1 (Fells-Sampson). Let (M, g) and (N, h) be compact Riemannian manifolds. Assume that (N, h) is of nonpositive
curvature. Then for any f E CO°(M, N), there is a harmonic map : M -- N free-homotopic to f.
u
Here that f a n d 4
are free-homotopic means that there exists
a continuous map u: M x [0,1] - N satisfying u(x, 0) = f (x),
u(x, l) = u,,,,(x),
x E M.
Also N being of nonpositive curvature implies that at each point y E N, the sectional curvature K(o) < 0 for each two-dimensional subspace o C T.N. In what follows, we simply denote it by KN < 0. Unlike the existence theorem (Theorem 2.23) of closed geodesics, the condition on the sectional curvature in Theorem 4.1 is necessary. In fact, an arbitrary f E C°° (M, N) is not always free-homotopic to a harmonic map, unless KN < 0 holds everywhere in N. For example,
Eells and Woods [2] show that any map f : T2 - S2 of mapping degree ±1 from the two-dimensional torus into the two-dimensional sphere is not free-homotopic to a harmonic map regardless of the Riemannian metrics g, h in T2 and S2, respectively. Namely, it is known that there is no harmonic map it : T2 - S2 of mapping degree ±1. In the case of the existence theorem of closed geodesics, as was seen in the proof of Theorem 2.23, we were able to apply the direct method in the variational techniques. Namely, we introduced a functional L which measures the length L(c) of piecewise smooth loops and directly constructed a closed geodesic regarded as a critical point of L from a minimal sequence. However, when M is a more general compact Riemannian manifold other than S, an application of the direct method to the energy functional C°° - R inevitably encounters certain difficulties. Indeed, the reason for the difficulties is that the equation for harmonic maps is essentially a system of nonlinear partial differential equations, as opposed to that the defining equation for geodesics is a system of linear ordinary equations in the tangent bundle TN of N. However, Bells and Sampson were successful in proving Theorem 4.1 using a technique called the heat flow method modeled after Morse theory on infinite-dimensional manifolds.
4.1. THE HEAT FLOW METHOD
121
The key points of the central idea may be described as follows. First, we review the first variation formula of the energy functional E as given in §3.3. Given a C°° map u E C°° (M, N), denote a C° variation of u by F = {ut}tEJ, I = (-e, e), and let
= ut E r(u-'TN) t=o
be the variation vector field. The first variation of E(ut) is given by (4.2)
d E(ut)
(V, r(u))dta9. t=o
M
Here, r(u) is the tension field of u, (, ) is the natural fiber metric in the induced vector bundle u-'TN, and pg is the standard measure in M determined by g. If we regard M = C°° (M, N) as a manifold by ignoring details, the energy functional E : C°° (M, N) - R can also be regarded as a function defined in M. Since the variation F = {ut}tEl of u can then be regarded as determining a curve in M, the variation vector field V E r(u-'TN) defined (4.1) represents nothing but the tangent
vector of this curve F at t = 0. As was seen in §3.3, for a given V E r(u-'TN), a C°° variation F of u was defined; consequently, r(u-'TN) may be regarded as representing the tangent space TuM of M at u E M. Hence, for given W1, W2 E r(u`'TN), ((W1, W2))
=
JM
(WI, W2)dp9
would define an inner product ((, )) in the tangent space Tu,M.
On the other hand, since the first variation d E(ut)t-o of E(ut) is considered to define the derivative dE.(V) of the function E on M in the direction of the tangent vector V, the first variation formula (4.2) can be expressed as
dEu(V) = -((r(u), V)). As is readily seen from the definition of the gradient (see Exercise 3.3 in Chapter 3), this implies that the tension field r(u) of u, indeed, is
nothing but the gradient vector of the functional -E at u; namely, we see that
r(u) = -(grad E)(u).
122
4. EXISTENCE OF HARMONIC MAPS
Consequently, a harmonic map u, which is a critical point of the energy functional, is precisely a singular point (a zero point) of the gradient vector field grad E of E.
rR
FIGURE 4.1. T(u) = -(grad E)(u) Analogous to Morse theory on the finite-dimensional manifolds (see Foundations of Morse theory, Iwanami Shoten, Gendai Sugaku no Kiso [27] ), we may say that the function E in M decreases most efficiently in the direction of -grad E, namely, the tension field r(u). Consequently, one may attempt to deform a given map uo = f E C°° (Al, N) along the flow determined by the tension field r(u) in M as a method to obtain a harmonic map free-homotopic to uo. If, indeed, a deformation as above is possible, its flow ut will be given as a solution to the equation
(4.3)t
= r(ut)
The equation (4.3) is a system of nonlinear parabolic partial dif-
ferential equations, and is nothing but the equation to obtain the integral curves of the tension field r(u) regarded as a vector field in M. Analogous to the so-called classical heat equation (see Appendix §A.2(c)), the method of deforming uo along the solution of (4.3) is, in general, called the heat flow method. Consequently, the existence problem of harmonic maps is reduced to whether or not the deformation of uo along ut reaches a critical point r(um) = 0 of the energy
4.1. THE HEAT FLOW METHOD
123
functional E. Keeping the above in mind, we consider, for a map u : M x [0, T) --+ N, the following initial value problem of a system of nonlinear parabolic partial differential equations (4.4)
f8 (x, t) = r(u(x, t)),
(x, t) E M x (0, T),
tu(xO) = f (x). Here, T > 0 and f E COO (M, N) is a map given as the initial condition. Also we assume that it is continuous in M x 10, T) and is C°° in M x (0, T); namely, u E Ce(M x 10, T), N) n C"- (M x (0, T), N).
A map u satisfying (4.4) is called a solution to the initial value problem (4.4). The system of nonlinear parabolic partial differential equations in (4.4) is called the parabolic equation for harmonic maps. In order to prove Theorem 4.1, given the initial value problem (4.4) of the parabolic equation of harmonic maps, we must show the following:
(1) For any initial value f, (4.4) possesses a solution u : M x Mx10,oo). (2) Set ut (x) = u(x, t) and t -r oo. Then ut converges to a harmonic map u,,, : M - N, and f = uQ and u,,,, are free-homotopic to each other.
In this section, we first assume that (1) holds, and discuss the statement in (2). Let (x¢) and (ye) denote local coordinate systems in M and N, respectively. For a given solution u to (4.4), set ut (x) u(x, t) and define m
E
n
o
e(ut) = 2IdutI2 = 2 4,j=1 a _1
E(ut) =
r
rc(ut) = 2
K(h) = J
e(uu)dµ9,
out
2
=
1
a4
axi 8
,'(ut)dp9.
Here, gtj and h,,,o are, respectively, the components of the Rim metrics g and h in the coordinate systems (xi) and (y'), and gij represents the components of the inverse matrix of (g{?) . We also
4. EXISTENCE OF HARMONIC MAPS
124
express the map u in the local coordinate systems by / /(x', u(2, t) = (u1(x1, ... , m, t), ...'U . , un n (xI, ... , x" , t)) (u
=
t)).
As is clearly seen from the definitions, E(ut) is the energy of each uc E C'°° (M, N), and K(ut) is the kinetic energy of the deformation determined by ut. For the energy density e(ut) and the kinetic energy density rs(ut) of each ut, we have the following formula called the Weitzenbock formula. This Weitzenbock formula is a fundamental equation satisfied by the solutions of the initial value problem (4.4), and it plays an important role in the arguments that follow.
PROPOSMON 4.2. Let u E CO(Mx[O,T),N)nC'°(Mx(O,T),N) be a solution to the parabolic equation for harmonic maps (4.4), and let ut (x) = u(x, t). We have, in M x (0, T), (Weitzenbock formula for e(ut)) (1)
ee
)
=®e(ut)
- IVVut)2
in
m
E Ricer (e e1)ej
,
dut(ei) )
j=1
M
+
(RN(dut(ej), dut (ej))dut (ej), dut(ej)) i,.i=1
(2)
(Weitzenbock formula for c(ut))
8K) Ot
(4.7)
(us) - Iiat Z m
RN
dut(e;),
"Ut
) Out
dut(e;) }.
Here, L1 is the Lapla©e operator of M, and Ric1 and RN denote, respectively, the Ricci tensor of M and the curvature tensor of N. {et} represents an orthonormal basis for the tangent space at each x E M, and (, ) represents the natural fiber metric in u-TN. Before the proof of the proposition, we review the Ricci identity,
which plays an important role in the process of computation. Let u E C°°(M, N). Consider the tensor product TM* ® u-1TN of the cotangent bundle TM* of M and the vector bundle u-'TN over M induced by u from the tangent bundle TN. Let V denote the the
4.1. THE HEAT FLOW METHOD
125
connection
r(TM` (& TM' 0 u-'TN)
V : r(TM` (D u-'TN)
compatible with the natural fiber metric (
,
) in TM* 0 u-'TN.
Consider the second order covariant differential
VVdu(X, Y, Z) = r(TM' ® TM' ® TM" ® u-'TN) of du E r(TM* 0 u-'TN). As is readily seen from the arguments in §3.2, for given X, Y, Z E r(TM), VVdu is a C°° section defined by
VVdu(X,Y,Z) = (Vx(Vydu))(Z) - (Vvxydu)(Z) In local coordinate systems (x') and (ya) of M and N, respectively, we express du, Vdu, VVdu and the curvature tensors RM and RN by n
in
du
a..
dxi ®
a o u,
i=1 a=1
m
n
Vi V j u° . dx` ®dxj ®w o u,
'Vdu = i,j=1 a=1
m
n
ViVjVkua.dxi0dxj®dxk®-ou,
VVdu= i j,k=1 ck=1
as
R U;i xj) 7
NGlya a
R
a axk
-
a RMI ijk fix! !=1
a
0 ' NO ! Oho' _ E a=1R
NS
a
Then the Ricci identity asserts that for each 1 < i, j, k < rra, 1 < cx < n, the following holds:
(4.8)
V1 V Vkua - V j ViQkYla _ to Mt qua
-
Na o
N'? a R ijk R 076 ax i axj ax k . x t + a !=1 ,7,b=1 n
u6
4. EXISTENCE OF HARMONIC MAPS
128
It can be readily verified that the Ricci equation is nothing but the defining equation for the curvature tensor R° of the connection V in TM* & u-'TN. The proof is left to the reader as a review for the definition of connection V in TM* ® u-'TN (see Exercise 4.1 at the end of this chapter). Now we prove Proposition 4.2.
PRooF. Given a solution ut(a) = u(x, t) to (4.4), as in a similar manner to the proof of Theorem 3.8, we consider the connection V compatible with the natural fiber metric { , ) in the vector bundle
T(M x (0,T))` 0 u-ITN over M x (0,T). Using this connection, we denote the covariant differentiation in the directions (8/ax1, 0) E and (0, d/dt) E T(y,t) (M x (0, T)), respectively, by Ttx.t)(M
Vt = V(o d/dt).
Vi = V(8/Asi,O)I
Let (x'), (fl) be local coordinate systems of M and N. Denote by gt,, hQ,# the components of the Riemannian metrics g and h, respectively. Since V is a connection compatible with the fiber metric
(, ), we see that Vtgakh,,0(ut) = 0,
Vtg hao(ut) = 0
hold for each 1 < i, j, k < m, I <_ a, 0!5 n.. (1) From the definition (4.5), we see first that
de) =Vt
E
9=aha,o(ut) ault9 -x
:,?=1 cr)3=1
M
n
aQ
E Eg'jhj9(U7t)Vt Ari ATi
t.j=1 Q$
Paying attention to the definition of V, we apply Lemma 3.7 to the C°° map u and the vector fields (01&4,0), (0, d/dt) in M x (0, T) in a similar manner to the proof of Theorem 3.8. Since [(0, d/dt), (&/Oxt, 0)1 = 0, we see that
vt&Lr
fit, 1
4.1. THE HEAT FLOW METHOD
127
Consequeiitly, we get
n
&,(U t) bit
n
V 1l1
()''u
- E E y jhn,l=1at
01-i
i.,7=1 (Y3=1 rn
it
Since we have
VkVIe(ut) m
= VIVk
2
i.j=1 n(3=i
m
Ou'l
n
E 9 Jhnl3(ut)
IT
i
8.0t
out
= Vk z,,j=1 c 13=1
n
n
EE
(Ilt) (VkVIV11t
i,j=1 n13=1
blu3 08.1.1
+
noting ViVizlt = ViVlur,1 < i, l < in. 1 < a < in as given in Corollary 3.5, we get n
De(ut) _
9k1VkVIe(1f,t) k.1=1
_
n
n
i, j,k,1=1
nii.1
3
It
9tj
cll
+ IVV'ut12.
Since we have, from Ricci identity (4.8). VkV1V04'
=ViVkVlu,'
fit
ER r=t
At r
kit
0'11,'
i`
%, n R'
011" 0116 011"
'I Al 0.rk
4. EXISTENCE OF HARMONIC MAPS
128
by substituting this into the previous equation, we get (4.10)
i,j=1 a,6=1
9'iko(u )Vi (
ij=1 0=1
\ j,k=1
Mr auto aup 1` gklRMW
g=jhaQ(ut)
air azi
1r=1
n
+
gkivkVj4tt) 1943 +IVVueI2
hap=1(u't)
X
ia,k,i=17,5,E=1
gi,gklRNO tout tout tout au't° 7aE &A; Oxi Oxi Oxj
in
1:(Ve,r(ut),dut(ei))
+ IVVutI2
i=1
m
+
m
(du(RicM (ei, ej )ej
i=1
-
dut(ei)
j=1
(RN(dut(ei),dut(ej))dut(ej),dut(ei)). i,j=1
On the other hand, we know that out = r(ut) holds, since ut is a solution of (4.4). We get the desired equation from (4.9) and (4.10). (2) One can prove this in a similar manner to (1). We leave it to the reader as an exercise problem (see Exercise 4.2 at the end of this 0 chapter).
We note here that in the proof of Proposition 4.2, the terms containing the derivative VVVut of the highest order in the compu-
tation of (a/at - L)e(ut), for example, were replaced by the term (RN (du, du)du, du) consisting of the curvature tensor and the first order differential du. This was done using the Ricci identity combined with the fact ut(x) = u(x, t) being a solution to the parabolic equation for harmonic maps. COROLLARY 4.3. Let u : M x [0, T) -+ N be a solution to the parabolic equation (4.4) for harmonic maps, and set ut (x) = u(x, t). The following hold in M x (0.T) :
4.1. THE HEAT FLOW METHOD
129
(1) If N is of nonpositive curvature KN < 0, and if, furthermore,
there exists a constant C such that Ric' > -Cg, then (
at
)
< Le(ut) + 2Ce(ut).
(2) If N is of nonpositive curvature KN < 0, then OK(ut)
at
< Ar.(u,t)
PROOF. (1) We only need to note that the fourth term of the right hand side in (4.6) -< 0 because of KN < 0, and that due to RicM > -Cg we have M
dut
E RicM (et, e3 )ej
> -Cdut (e{).
j=1
0
(2) This follows easily from (4.7).
We remark here that there always exists a constant C E R such
that RicM > -Cg, since M is compact. From Corollary 4.3, we readily see the following.
PRoPOSMON 4.4. Let u : M x (0, T) -+ N be a solution to the parabolic equation (4.4) for harmonic maps, and set ut (x) = u(x, t). Then the following hold in M x (0.T) : (1) E(ut) is a monotone nonincreasing function. Namely.
d E(ut) = -2K(ut) < 0. (2) If N is of nonpositive curvature KN < 0, then
d E(ut) = -2dK(ut) > 0. Namely, E(ut) is a convex function and K(ut) is a monotone nonin-
crewing -
PROOF. (1) By the first variation formula in Theorem 3.8, we have
E(ut)
JM c 8 'T(ut))dtAg JM(t&)dµS =
-2 K(ut).
4. EXISTENCE OF HARMONIC MAPS
130
(2) From (2), Corollary 4.3, and Green's theorem, we get
K(h) _
JM K(u1)d/L9 =
I
JM
)dµ9
JM
u)dp= 0. 0
Proposition 4.4 implies the following. Let N be of nonpositive curvature KN < 0. Further assume that the parabolic equation (4.4) for harmonic maps has a solution it : M x [0, oo) N for T = oo. Noting (1), Proposition 4.4 and E(ut) > 0, K(ur) --, 0 must hold as t - 0 for Ut(x) = u(x, t). Hence, 5u /c9t --> 0 must hold as t --+ 0. However, since u is a solution to (4.4), this implies that T(ut) -> 0 holds as t -, 0. Thus, when t oo, ut converges to u. E C30 (M, N). Consequently, if r(ut) - r(u.), r(ti ) = 0 must hold. In other words, vt converges to a harmonic map u... . In what follows, we verify step by step that there exists such a solution as described above to the initial value problem (4.4) of the parabolic equation for harmonic maps.
4.2. Existence of local time-dependent solutions Let (M, g) and (N, h), respectively, be compact R.iemannian manifolds of dimension m and n. First, we note that the following holds on the differentiability of solutions to the equation for harmonic maps.
THEOREM 4.5. If a C2 differ the equation for harmonic maps
able map u : M
N satisfies
r(u) = 0,
(4.11)
u is a C°° differentiable map. PRooF. We may locally verify differentiability of u at each point x E M. We choose a coordinate neighborhood V about x and a coordinate neighborhood W about u(x) so that u(V) C W holds. Denote by (xi) and (ye) local coordinate systems in V and W, respectively. With respect to these local coordinate systems, equation (4.11) for harmonic maps can be expressed, for each ua = y O a u, as m
n
Out9 8Zt'1
ij=1 j3,7=1
Here, d is the Laplace operator of M and rc represents the connection coefficients of the Levi-Civita connection on N.
4.2. EXISTENCE OF LOCAL TIME-DEPENDENT SOLUTIONS
131
Suppose that a C2 map u satisfies equation (4.12). Since the right hand side is a C' function, it, in particular, is o-molder continuous for 0 < o < 1. Consequently, u is of C2+° from the theorem on differentiability for the solutions to linear elliptic partial differential equations (see Appendix §A.2(d)). The right hand side of (4.12) becomes of C'+°; hence, u is of C2+a from the same theorem. Repeating this argument, we see that a is of C°°. Theorem 4.5 tells us that we only have to create a solution of at least C2 to the equation (4.11) for harmonic maps, in order to show existence of the desired harmonic maps. With this remark in mind, we consider, with a map f : M -4 N given as an initial value, the following initial value problem of the parabolic equation for harmonic maps (4.13)
W (x, t) = r(u(x, t)), u(x, 0) = f (x),
(x, t) E hf x (0, T),
concerning a map u : M x [0, T) - N. The objective in this section is to show that the initial value problem (4.13) possesses a solution for a sufficiently small T > 0. The solution as above is called a local time-dependent solution of (4.13).
For the purpose of discussing the existence of the local timedependent solutions, we rewrite the parabolic equation (4.13) for harmonic maps in a form that is analytically more desirable. To this end, we use the Nash imbedding theorem [151 which shows that an arbitrary compact Riemannian manifold can be isometrically imbedded in Euclidean space of sufficiently high dimension. In other words, we may assume, without loss of generality, that the Riemannian manifold (N, h) is realized as a submanifold of the q-dimensional Euclidean space Rq for a sufficiently large natural number q, and that the Riemannian metric h is nothing but the induced metric from Rq. Let denote such an isometric imbedding, and let N be a tubular neighborhood of the submanifolds t(N) C Rq in Rq. Namely, for a sufficiently small e > 0, N is an open subset (see Exercise 4.3 at the end of this chapter) of Rq defined by
N= {(x, v) I x E t(N), v E Tt(N)l, lvi < F}.
4. EXISTENCE OF HARMONIC MAPS
132
In the tubular neighborhood N, let
7r.N-,t(N) denote the projection; namely, 7r is the map that assigns to each z E N
the closest point in t(N) from z.
Let u : M x [0, T) --i N be a map from M x (0, T) into N c. R9. Regarding u as a R9 valued function, we consider the following initial value problem for the system of parabolic partial differential equations: (4.14)
(&-a) u(x, t) = fI(u)(du, du)(x, t),
(x, t) E M x (0, T ),
u(x, 0) = 10 f (x).
Here, A is the Laplace operator of M and f is the map given as the initial condition of (4.13). fI(u)(du, du) is a vector in R9, and is defined as follows. Let (zA) be the standard coordinate system of Rq and let (x`) be a local coordinate system in M. With respect to them, we express ir(u) and u(x, t), respectively, as 7r(z) = (7r1(zi, ... , z9), .
. . ,
7r9(z1, . . .
,
z4)) = (7rA(zB)),
u(x, t) = (ul (xi, ... , xm, 01. . , u9(xl, ... , xm, t)) = (uA(xi, t)) .
Then the components of H(u) (du, du) are given by m
9
sC j
7rA
auB 7uC
g 8zB8zc (u) ax¢ iJ=1 B,C=1
T.?,
1 < A < q.
As is readily seen from Lemma 3.4, if we denote the second fundamental form of the map it by Vdu, we get (4.15)
II(u) (du, du) = trace Vdir(du, du).
Among the solutions to the initial value problem (4.14), we consider those u : M x [0, T) -i N which satisfy
u E C°(M x [0,T), N) n C2,1(M x (0,T),1V); namely, those which are continuous maps from M x [0, T) into N, and
are, furthermore, of C2 in M and of C' in (0, T). The initial value problems (4.14) and (4.13) are related to each other in the following way.
4.2. EXISTENCE OF LOCAL TIME-DEPENDENT SOLUTIONS
133
PROPOSrri0N 4.6. Let u E Co (M x [0, T), N) nC2" 1(M x (0, T), N).
If u is a solution to the initial value problem (4.14), u(M x [0, T)) C t(N) holds, and u is a solution to the initial value problem (4.13). The converse also holds true.
PRooF. Suppose that u E C°(Mx[0,T),N)nC2,I(Mx(O,T), is a solution of the initial value problem (4.14). First, we verify that u(M x [0, T)) c t(N) holds. To this end, we define a map p : N -+ R4 by
p(z)=z-ar(z), zEN, and a function p:Mx[0,T)-'Rby V(x, t) = l p(u(x, t))12,
(x, t) E M x [0, T).
From the definition of 7r, p(z) = 0 is equivalent to z E t(N). Hence, it suffices to see p(z, t) - 0. Since u(x, 0) = t o f (x) E t(N), we see cp(x, 0) = 0. Also since u is a solution to (4.14), we get at _ at (p(u), p(u)) = 2 dp
\ at
,
p(u))
= 2(dp(Liu -11(u)(du, du)), p(u)), Acp = L (p(u), p(u)) = 2(Op(u), p(u)) + 21Vp(u)I2,
where (, ) is the inner product in R. On the other hand, from a formula for the second fundamental form of composite maps (see Exercise 4.4 at the end of this chapter), we have that
Ap(u) = dp(Lu) + trace Vdp(du, du). Since r(z) + p(z) = z holds from the definition, we have that dir + dp is the identity map and Vdar + Vdp = 0. Noting these together with that the images of dir and p are orthogonal to each other, we get
tap = 2(dp(du) - trace Vdir(du, du), p(u)) + 2IVp(u)12 = 2(dp(Lu - II(du, du)), p(u)) + 21Vp(u) 12. Consequently, we have 00
at
= o - 21Vp(u)I2
'
4. EXISTENCE OF HARMONIC MAPS
134
Green's theorem yields, for each t E (0, T), jM co(, -t)dp9
=
f(, t)dp9
-2
JM
I VP(u) I2dµg < 0.
Hence, we have JM (, t)dug
fM p(., t)dp9 =
implying that cp(x, t) - 0.
Next, we verify the the second half of the assertion. To this
end, let u M x [0, T) --> N be a map from M x [0, T) into N, and set it = t o u. We must show that u is a solution to the initial is a solution to the value problem (4.13), if it : M x [0, T) --+ t initial value problem (4.14). From the definitions, it = t o u and t = 7r o t. Hence, from the formula for the second fundamental form of composition maps, we get Au = trace ©dt(du, du) + dt(r(u) ), trace Vdt = trace Vd7r(dt, dt) + dir (trace Vdt). Since t : N -t IlS9 is an isometric imbedding, noting that trace Vdt is orthogonal to t(N) at each point, we get do (trace Wt.) = 0. These equations yield
dt(r(u)) = L
- traceVd7r(du, du). it
On the other hand, since dt (d at dt
r(u) -
=
au
, we have
) = (f_)u_H(u)(dudu).
As a result, if it is a solution to the initial value problem (4.14), u becomes a solution to the initial value problem (4.13). It can be easily verified that the converse also holds. From Proposition 4.6, we see that we can get a time-dependent local solution to the initial value problem (4.13) by constructing a timedependent local solution to the initial value problem (4.14). Since the equation in the initial value problem is a system of parabolic differential equations with regard to the vector valued function, it is relatively
easy to set up a function space in which existence of solutions is to be discussed. In what follows, we construct a time-dependent local solution to the initial value problem (4.14).
4.2. EXISTENCE OF LOCAL TIME-DEPENDENT SOLUTIONS
135
Following Ladyzenskava. Solonnikov and Ural'eeva [9, p. 71. we set lip a function space in which the existence of solution is treated as follows. Given T > 0, set Q = Al x [0. Tj. Let 0 < a < 1. Given a vector valued function u, Q - R`t. set
IuIQ = sup
(x.t)EQ
(u)("")
=
I41(x. t)I.
I u(.r, t) - u(.r'. t))
sup
')°
(
x#x' (u) (a)
=
lu(a . t)
sup
-
t')I
It - PI",
(x.t).(r.t')EQ
tot'
and define the norms
IUI(
.a,2) =IuIQ IZ61(2+a.1+a/2) =IuIQ
(
4 1 8) .
by
IuIQ
Q
+ (u)(°) + (,)(n/2) + Iat-uIQ + IDxuIQ + IDruIQ (8t'tl)ta/2)
+
+ (Dxa)(1/2+°/2) + (Dru)t"/2)
+ (Otu);.() +
Here, d(x, x') is the distance between .r and in A!. and Otu represents c u/O3t. Also, Dxu and Dxu represent the first order derivative of u in Al and its covariant derivative, respectively. In terms of a local coordinate system (:ri) in Al and the coordinate functions (;y°) in Rq, Du and Dxu are, respectively, defined by
Dxu=du=
q
au"
i=1 a=1
of
d.r i ®
i3
Vy
Dxu=Odu=
oya
i.j=1 cr=1
and IDrulq and ID,2.uI2 are, respectively. given as m 2
D.xuIQ
sup
q
9j
a
u ct
i
(x.t)EAI :.j=1
Ox
q
I D=uI2 =(x.t)EA1 sup
gikyltt7i0jtt"p plu i.j.k.l=1 a=1
4. EXISTENCE OF HARMONIC MAPS
136
With respect to these norms, we define the function spaces Ca'«/2(Q, Rq) and C2+a,1+a/2 (Q,), respectively, by Ca'a/2(Q,Rq) = {u E C°(M x [0, 71) 1 ,C2+a,I+a/2(Q,R9) = {u E C2,1(M x [0,T1) I
IuIQ'
2)
< oa},
ful'+a,1+al2)
< oo},
and set
C2+a,i+a/2(Q, N) = {u E
Ca'a/2(Q, Rq) and
C2+a,1+a/2(Q,
Rq) I u(Q) C N}.
C2+a,1+a/2(Q.
Rq) are Banach spaces with norms respectively. It is easy to verify that Q 4 C2+«,1+«/2(Q, Iq) C,2+a,1+«/2(Q, N) is a closed subset of (see Exercise 4.6 at the end of this chapter). Ca'a/2(Q,1 ) and C'+a}1+a/2(.Q, Rq) are called a Holder space on Q = M x [0, T]. We now prove the following. IuI(«,a/2)'
IuI(2+a,l+a/2),
THEOREM 4.7. Let (M, g) and (N, h) be compact Riemannian manifolds. For any C2+a map f E C2+4(M, N), there exist a positive C2+a,1+a/2(M x [0, el,1V) such number e = e(M, N, f, a) > 0 and u E that u is a solution in M x [0, e) to the initial value problem (4.14). Here, c = e(M, N, f, a) is a constant dependent upon M. N, f and a.
We prove this theorem using the inverse function theorem (see Appendix §A.2(a)) in Bausch spaces. The idea of the inverse function theorem is to reduce solvability of a nonlinear differential equation to solvability of a linearized equation. First, we review the results regarding existence and uniqueness for linear parabolic partial differential equations.
THEOREM 4.8. Let (M, ,g) be a compact Riemannian manifold, and set Q = M x [0, 7']. Given a vector valued function u : Q - R9, let
Lu= A+a be a parabolic partial differential operator, and consider an initial value problem (4.17)
= F(x, t), u(x, 0) = Ax) {Lu(x,t)
(x, t) E M x (0, T),
4.2. EXISTENCE OF LOCAL TIME-DEPENDENT SOLUTIONS
137
Here, the components of Au, a Vu, b u, at u are, respectively, defined
q mg
by
f u',
E aB (x, t) axt
> bB(x, t)uB,
,
B=1
B=1 -=1
If
aB ,
1 < m, 1 < A, B < q,
bB E C 'a/2(Q, JR),
for some 0 < a < 1, then, for any F E Ca'a/2(Q,IR ), f E C2+a(M,Rq), C2+a,l+a/2(Q,R) to (4.17) there exists a unique solution u E F E such that iul(2+a,1+a/2) < Q
holds.
<-
Q
(2+a)) If (2
Here, C = C(M, L, q, T, a) is a constant dependent on
M, L, q, T, a alone.
Theorem 4.8 is a classically well known result. For more details, the reader is advised to consult Ladyzenskaya, Solonnikov, and Ural'ceva [9, p. 320] or Partial Differential Equations 1, Gendai Sugaku No Kiso, Iwanami Shoten [28J. We now prove Theorem 4.7.
PROOF. First, choose an a' such that 0 < a' < a < 1. Also we extend the projection is : N -- t(N) from the tubular neighborhood N of the isometric imbedding t : N -, Rq to a C°° map zr; Rq -- R9. We may assume here that is a constant map outside a sufficiently large ball such that N C B C Rq. This can be done by choosing N small enough if necessary. We assume that the right hand side of (4.14) is defined for this extended it by using (4.15). Also for the sake of simplicity, we express 8/& and 8/&A by o' and 8A, respectively. Step I (Construction of an approximating solution). First, by identifying f with t of, we consider the following initial value problem of a system of linear parabolic partial differential equations:
-
f(L O v(x, t) = II(f)(df, df)(x), (x, t) E M x (0,1), lv(x, 0) = f (X). Since the assumption on f yields f E C2 (M,1"), II(f)(dF, df) E C
there exists a unique solution VE
C2+a,1+a/2(M
X [0,1J, Rq)
4. EXISTENCE OF HARMONIC MAPS
139
to (4.18) from Theorem 4.8. Denote the desired solution by u. Then v approximates u at t = 0 in the following sense. Namely, we note
that v(x, 0) = u(x, 0),
8tV(x, 0) = Otu(x. 0).
Step 2 (Application of the inverse function theorem). Set Q = M x [0,1] and consider a differential operator
P(u) = Au - atu - II(u)(du, du). x [0, c], R9) satisfying P(u) = 0 is the desired
AUE solution.
Now for 0 < al < 1, we define the subspaces X and Y in C2+a,1+a/2 (Q,
R) and Ca'°a'/2(Q, ]9), respectively, as follows:
X = {z E C
2+a',l+a'/2(Q,
R) I z(x, o) = 0, atz(x, t) .= 0},
Y = {k E Ca%a''2(Q, R4) I k(x, 0) = 0}. From the definitions, X and Y are closed subspaces; hence, Banach spaces. For a given z E X, if we put
P(z) = P(v + z) - P(V), P defines a map P : X -+ Y from X into Y, as we can easily see P(z) E Y. In particular, P(0) = 0. P is Rkchet differentiable in a neighborhood of z = 0. A simple calculation (see Appendix §A.2(a)) using the definition yields that the Frechet derivative P'(0) : X -p Y is given, for each Z E X, by 9
P'(0)(Z) = AZ -- atZ -
ZAaAII(v)(dv, dv)
- 2II(v)(dv, dZ).
A=1
From this it can be readily seen that P'(0) : X -- Y is an isomorC2+a,l+«/2(Q,111:9), phism. In fact, since v E from the definition of P'(0) and Theorem 4.8, we see that for any K E Y, there exists a unique Z G
G"2+a`,I+a'/2(Q
111;9) satisfying
(P'(0)(Z))(x, t) = K(x, t),
(x, t) E M x (0,1),
Z(x, 0) = 0.
We also see that for such a Z the following holds: (4.19)
IZIQ+a ,1+a'/2) < CIKI(',a'/2).
Since K(x, 0) = 0 and Z(x, 0) = 0 hold, we have that 8iZ = 0 holds; consequently, Z E X. This tells us that P'(0) is surjective. On the
4.2. EXISTENCE OF LOCAL TIME-DEPENDENT SOLUTIONS
139
other hand, we see that P'(0) has a continuous inverse map from (4.19). Hence, P'(0) is an isomorphism. Applying the inverse function theorem (Theorem A.2) for Banach spaces, P : X -+ Y is a homeomorphism between a sufficiently small neighborhood U of 0 E X and a neighborhood P(U) of 0 E Y. In other words, there exists a positive number 6 = a(M, N, f) > 0 such that C2+a',1+a'/2(Q, R) satisfying the following there exists a unique z E IkIQ'a,/2 C'',a'12 with k(x, 0) = 0 and conditions. For any k E < 6, z satisfies P(z)k,
(4.20)
8t(z, 0) = 0.
z(x, 0) = 0,
Here, b = S(M, N, f) is a positive number determined by M, N and f. Now if we set u = v + z and w = P(v), from (4.20), we see that there exists a u E C2+a'.1+« /2(Q, R) satisfying
P(u)(x, t) = (w + k)(x, t), u(x, 0) = f W.
(4.21)
(x, t) E M x (0,1),
Step 3 (Existence of time-dependent local solutions). In order to see the existence of a desired local time-dependent solution, for a given positive number e, consider a C°° function ( : R - R satisfying ((t) = 1 (t < e), ((t) = 0(t > 2e), 0 < C(t) < 1, IK'(t)I 2/e (t E R). We note that w = P(v) E Ca,a/2(Q, R9) C C°',4'/2(Q, $4) and that w(x, 0) = 0 holds from the definition of P(v) and v(x, 0) = f. We can verify through a simple calculation (see Exercise 4.7 at the end of this chapter) that there is a constant C > 0, independent of a and w, such that (4.22)
Ibwl(a',a'/2) < Q
Cc(a-a')/21u,1(a,a/2)
Q
holds.
Set k = -(w. Then K(x, 0) = 0. From (4.22), we have Ikt(',a'/2)
< b for a sufficiently small e. Consequently, there exists a uE x [0, e], R4) such that the following special case of (4.21) holds: C2+a',1+a'/2(M
J P(u)(x, t) = 0, lu(x, 0) = AX).
(x, t) E M x (0, e),
4. EXISTENCE OF HARMONIC MAPS
140
Namely, we have obtained a solution u E to the initial value problem
(A - O)u(z, t) = II(u)(du, du)(z, t), lu(x, o) = AX).
C2+a'.1+a'/2(M
x (0, E], lRq)
(z, t) E M x (0, E),
Since we have
f E C21, (M' V),
n(uu)(du, du.) E Ca,c,12(M x [0, e], IQ),
we see by Theorem 4.8 that uE
C2+a,l+a/2(M
x [01 E],ff4).
Since u(M x [0, e']) C J for a sufficiently small positive number E' (0 < e < E), u is a solution to the initial value problem (4.14) in M x (0, tj. Applying Proposition 4.6, we see that u is a solution to (4.14) in M x [0,,E] . It is also clear from the above proof that E > 0 is a positive number depending on M, N, f and a alone. 0 From Theorem 4.7 and Proposition 4.6, the following clearly follows.
COROLLARY 4.9. Let (M, g) and (N, h) be compact Riemannian manifolds. For a given C2+a maP f E C2+a(M, N),
them exist a positive number T = T (M, N, f, a)
U E C2+a,1+a/2 (M X [0, T], N) such that
ON' (z, t) = r(u(x, 0),
> 0 and
(x, t) E M x (0, T),
I ae(x, 0) = AX) holds. Here, T = T(M, N, f, a) is a constant dependent on M, N, f, a atone.
Noting the result regarding differentiability on the solutions to a linear parabolic partial differential equation, we obtain the Mowing result on existence of time-dependent local solutions to an initial value problem for the parabolic equation for harmonic maps. THEOREM 4.10 (Existence of time-dependent local solutions). Let
(M, g) and (N, h) be compact Riemannian manifolds. For a given
C2+a map f E C2+-(M, N), them exist a positive number
'13. EXISTENCE OF GLOBAL TIME-DEPENDENT SOLUTIONS
T = T(M, N, f,a) > 0 and u E
C2+a,1+a/2(M
141
x f0,T1,N) fl
C° (M x (0, T), N) such that 5 N(x, t) = r (u(x, t)),
(x, t) E M x (O,T),
lu(x,0) = f(x) holds. Here, T = T (M, N, f, a) is a constant dependent on M, N, f, a alone.
PROOF. Let u E C2+a,1+a/2(M x [0,T],N) be the solution in Corollary 4.9. We only have to verify the differentiability of u about each point (x, t) E M x (0, T). As in the proof of Theorem 4.5, denote
by (x') and (y) the local coordinate systems about x and u(x, t), respectively. With respect to these local coordinate systems, the parabolic equation for harmonic maps is expressed for each u° = ya o u as
)'U'
m
n
Y
=,j=1 C1+a,a/2 Noting that the right hand side is from the assumption on u, we see that the theorem (see Appendix §A.2(d)) regarding differentiability on solutions to linear parabolic partial differential equaC3+a,1+°/2. tions implies that u is of This yields that the right hand C4+p,1+a12 side is of C2+-.1+a/2. Then we, in turn, see that u is of Iterating this argument gives us that u is of C°° about each point. 0
4.3. Existence of global time-dependent solutions As was seen in §4.1, in order to prove Theorem 4.1 using the heat flow method, it was necessary to show that the initial value problem of the parabolic equation for harmonic maps (4.23)
J(x,t) = r (u(x, t) ),
(x, t) E M x (0, T ),
lu(x,0) = f(x)
had a solution u : M x [0, oo) - N when T = oo. We call such a solution in M x [0, oo) as above a global time-dependent solution to (4.23). As seen in Theorem 4.10, a local time-dependent solution to (4.23) always exists. However, the parabolic equation for harmonic maps is a system of nonlinear partial differential equations; hence, existence of a global solution is not always guaranteed. In fact, to show the existence of a global solution, it becomes crucial to estimate
4. EXISTENCE OF HARMONIC MAPS
142
the growth rate of the solution u(x, t) in time t. In order to control the effect of the nonlinear terms of the equation, the curvature of the Riemannian manifold N plays an important role. In this section, we investigate the relationships between the existence of global timedependent solutions and the curvatures of M and N.
In what follows, let (M, g) and (N, h) be compact Riemannian manifolds of dimensions m and n, respectively. First of all, we discuss the maximal principle for the beat equation as a tool to estimate the growth rate of the solutions to (4.23). Let 0 be the Laplace operator
of M, and Let L = A following holds.
5i
be the heat operator. We verify that the
LEMMA 4.11. Let U E C°(M x [0,T))nC23 (M x (0,T)) be a deal valued function in M x 10, T), which is C2 in M and C' in (0, T). If
u satisfies Lu > 0 in M x (0, T), then
max it = max u
Mx(O,TJ
MX (0)
holds; namely, the maximum value of it in M x [0, T) is attained at a point in M x {0}. PROOF. Let e1, c2 > 0 be positive numbers and set
u(x,t)=u(x,t)-Eit,
Q = M x [0,T-f21.
Regarding the maximum value of ta, we have (4.24)
a u = Mmax} u.
In fact, since u is a continuous function in Q, it attains the maximum
value at a point (x°, t°) in Q. we must show t° = 0. We suppose t° > 0, and induce a contradiction. Since Lu > 0 in M x (0, t) from
-
the assumption, u satisfies at (x°, t°)
Mrk k=1
holds at (x°, t°). Here, q 's are the connection components of the Levi-Civita connection in M. Since u(x°, t°) is the maximum value of
4.3. EXISTENCE OF GLOBAL TIME-DEPE (DENT SOLUTIONS
143
u in Q, we get
0 (x°, t°) > 0, a2fi
and the matrix
(x°, t°)
afi axi (x°, t°)
is nonpositive definite. But this
contradicts fI > 0. Hence, (4.24) has been proved. Noting el, Ea > 0 are arbitrary in (4.24), we can readily obtain the desired conclusion.
0 From the Weitzenbock formula seen in §4.1 and Lemma 4.11, we have the following estimate regarding a solution u to the initial value problem (4.23) of the parabolic equation for harmonic maps.
PRoPosmoN 4.12. Let u E C,1(Mx [0, T),N)nC°0 (M)40, T),N) be a solution to (4.23) and set ut (x) = u(x, t). Assume that N is of
> -Cg for a ant 0 and that R C E R. Furthermore, let c be a positive number such that 0 < e < T. nonpositave curvature KN
Then the following hold for the energy density e(n) of u : (1) For an arbitrary (x, t) E M x (0, T), e(ut)(x) < e2Ct sup
xEM
e(f)(x)
(2) For an arbitrary (x, t) E M X 1c, T),
e(ut)(x) < C(M,f)E(f). Here, C(M, e) is a constant dependent on only M and E.
PROOF. (1) First, we note that an inequality
Le(ut) =
\A
--
e(ut) > -2Ce(ut)
regarding e(u) is obtained from Corollary 4.3 (1). If we put v(x, t) = e-2Cte(ut), we see, through a simple computation, that v satisfies Lv > 0 in M x (0, T). Hence, from Lemma 4.11, e-2Cte(ut)(x) = v(x, t) :s =m xv(x,o) = m xe(f)(x)
holds at an arbitrary (x, t) E M x [0, T).
(2) Regard e(ut)(x) and E(ut) as functions of (x, t) and t, respectively. We express them as e(u)(x, t) and E(t)(t). Furthermore,
4. EXISTENCE OF HARMONIC MAPS
144
let C be the constant in (1) and let 0 < E < t < T. For a given (x, t) E M x (0, T), we set
"(X, S) -
H(xe y,
JA,
s)e-2C(t-E)e(u)(y, t - e)d119 (y),
where H(x, y, s) is a fundamental solution to the heat equation over M. dµ9 (y) indicates that we integrate in y with regard to the standard measure 1A. in M. The property of the fundamental solution readily implies that w1 (x, s) satisfies (& - 0) W1 (X, s) = 0, (x, s) E M x (0, T), { lima10 wi (x, 8) = e-2(t-0 e(u) (x, t - E),
(see Appendix §A.2(c)). On the other hand, if we put w2(x,
s) = e-2c(a+t-E)e(u)(x, s + t - e),
we see, through a simple computation, that "(x, s) satisfies (L
- &) "(x, s) > 0,
(x, s) E M x (0, T),
11me10 w12(x, 8) = e-2C (I-.)e(u)(x, t - e).
Since by putting w3=w2-wl, we have Lw3 > 0,
w3(x, 0) = 0,
we get from Mama 4.11 W2(x, 8)
w1 (x. s),
(x, s) E M x [0, T).
If we put s = E in this inequality, we see that (4.25)
e(u) (x, t) < e2 1 H(x, y, c)e(w)(y, t - E)dµg(y)
holds for any (x, t) E M X (E, T).
As one of the properties of the fundamental solution there, we note that there exists a constant e(M, e) (Corollary A.4) dependent
on only M and >0 such that H(x, y, e) < c(M, e).
4.3. EXISTENCE OF GLOBAL TIME-DEPENDENT SOLUTIONS
145
Hence, we get from (4.25)
e(u)(x, t) < e2C'c(M, E)E(u)(t - E).
On the other hand, since E(u)(t) is a monotone nonincreasing function as seen in Proposition 4.4, we get
e(u)(x, t) < c
c(M, E)E(f ).
Putting C(M, e) = e2C`c(M, c), we obtain the desired result.
0
PROPOSITION 4.13. Let u E C2,1 (Mx 10, T),N)nC°°(Mx(o, T),N)
be a solution to (4.32). If N is of nonpositive curvature KN < 0, at any (x, t), we have at
(x,t)<suI(x,0)1. xEAf at
PROOF. From Corollary 4.3 (2), we have
Lo(ut)=
(©-
]?C(ut)>0
for ut(x) = u(x, t). Hence, we get the desired result from Lemma
0
4.11.
Propositions 4.12 and 4.13 imply that the growth rate of a solution u to the initial value problem (4.23) is uniformly bounded with respect to time, if N is of nonpositive curvature KN < 0. Namely, we get the following. be
PROPOSITION 4.14. Let u E C2"(Mx [0, T),N)nC°°(Mx(0, T),N) a solution to (4.32). If N is of nonpositive curvature
KN < 0, then, for any 0 < a < 1, there exists a positive number
C=C(M,N,f,a)>0such that
1 u(., t)I c2+a(M,jv) + 1-,j9ui ICt(M,N) <
C
holds at any t E 10, T). Here, C = C(M, N, f, a) is a constant dependent only on M, N, f and a. PROOF. As in §4.2, where we proved existence of a time dependent local solution, we consider that (N, h) is realized as a submanifold in the q-dimensional Euclidean space Rq via an isometric imbedding t : N --+ R9 and that the vector function u : M Rq is a solution
4. EXISTENCE OF HARMONIC MAPS
146
to (4.14). Then u satisfies the system of elliptic partial differential equations
Au = II(u)(du, du) + o where A is the Laplace operator in M. Noting Propositions 4.12 and 4.13, we see that the right hand side of the above equation is bounded independent of t E [0, T). In other words, we have
t)IL-(M,R4) < ci(M,N,f) Since the image of u is always contained in the bounded set N c Rq, we have I u(', t) ILOO(M,R4)<
c2(N).
Consequently, by the Schauder estimate (see Appendix §A.2(e)) for the solutions to an elliptic partial differential equation, at any t E [0, T), we have (4.26) I u(', t) ICI+a(M,R4) <_ c3(M, a)
( tE[O.T) sup I AU(', 01 L°(M,R4) + SUP
I u(', t) ILQO(M,Rq)
tE [O,T)
< C4 (M, N, f a).
On the other hand, u is also a solution to the system of parabolic partial differential equations
Lu = II(u)(du, du),
where L = A - at is the heat operator in Al. Noting (4.26), the C° norm of the right hand side of this equation is bounded independent of t E [0, T). In other words, we see that I II(u)(du, du)(., t) IC°(M.N):5 c5(M, N, f, a)
4.3. EXISTENCE OF GLOBAL TIME-DEPENDENT SOLUTIONS
147
Hence, by the Schauder estimate for linear parabolic partial differential equations (see Appendix §A.2(e)), we get, for any t E [0,T),
I u(.,t)
IC2+a(M,P9)
cs(M, a)
+I
t)I Ca(M,R )
sup I Lu(., t) Ic'=(M,RQ) + sup I u(', t) I L-(Mi +)
tE [O,T)
t E [O,T)
c7(M,N,f,a).
0 At this point, we review the uniqueness of the solutions to an initial value problem of a parabolic equation for harmonic maps. THEOREM 4.15 (Uniqueness of solutions). Let (M, g) and (N, h)
be compact Riemannian manifolds, and assume that both u1, u2 E C°(M x [0, T), N) n C2,1 (M x (0, T), N) satisfy the parabolic equation for harmonic maps
(x, t) = T(u(x, t)),
(x, t) E M x (0, T).
Then ul = U2 in M x {0} implies u1 = u2 in M x [0,T). PROOF. As in the proof of Proposition 4.14, we regard u1, u2 as vector valued function u1, U2: M - t(N) c RQ, and consider u1, u2 as solutions to the parabolic equation (4.14) for harmonic maps. Define a function cp : M x 10, T) - R by (p(x, t) = I u1 (x, t) - U2 (X, t)
12,
(x, t) E M x 10, T).
Then cp in M x [0, T) satisfies
-
- n(u2)(du2,du2))
+ 2Id(ul - U2) 12. If we rewrite 11(ui)(dui, du1) - 17(u2)(du2, due)
_ (11(u1) -n(u2))(du1,ehel)+f(u2)(dui -du2idu1) + n(u2)(du2, dul - du2),
4. EXISTENCE OF HARMONIC MAPS
148
and apply the Mean Value Theorem to 11(ul) - 11(u2), we can readily verify that
>- -csJui - uzI(Iui - u2I + Id(ul - u2)I)
+ 2Id(ul - U2)12 --C9o.
b > 0, E > 0). Also Here, we used an inequality ab < Eat + c8 and cg are constants dependent on up to the third derivatives of
the projection it : N - t(N) of a tubular neighborhood N of t(N), and dependent on the maximum values of the energy densities e(ul) and e(u2) of ul and u2 in M x [0,T). Since cp(x, 0) = 0 from the assumption, using the maximal principle (see Exercise 4.8 at the end of this chapter), we get p(x, t) = 0,
0
namely, u1Eu2
Now we can prove the following regarding existence of a timedependent global solution to the initial value problem of the parabolic equation for harmonic maps. THEOREM 4.16. (Existence of time-dependent global solutions)
Let (M, g) and (N, h) be compact Riemannian manifolds, and assume that N is of nonpositive curvature KN < 0. Then for any C+Q
map f E
C2+°.I+°/2(M
(4.2'1)
N), there exists a unique u E
x (0, oc), N) fl CO* (M x (0, oo), N) such that W (x, t) _ 'r(u(x, t)),
(x, t) E M x (0, oo),
u(x,0)=f(x)
holds.
PROOF. By Theorem 4.10, there exists a positive number
T = T(M, N, f, a) > 0 such that, regardless of the curvature of N, the initial value problem (4.27) has a solution u E x [0, T], N) n 47=(M x (0, T)) in M x [0, T]. In particular, if N is of nonpositive curvature KN < 0, we must show that this solution u can be extended to M x (0, oo). Set To = sup{t E [0, oc) 1 (4.27) has a solution in M x [0, t] }.
We will show To = oo. To this end, assume To < oo, and let {tj be a sequence of numbers that converges to To. As in the proof of Proposition 4.14, we consider N to be a submanifold in q-dimensional
4.3. EXISTENCE OF GLOBAL TIME-DEPENDENT SOLUTIONS
Euclidean space Rq, and regard each
149
t;) E C°°(M, N) as a vector
valued function u : M - R. We also express by Ot. Let positive numbers a and a' be chosen so that 0 < a < 5i a' < 1 holds. We see from Proposition 4.14 that the sequences of functions
t,)} and
t;)}
C2+a'(M, form uniformly bounded subsets in the function spaces Rq) and Ca (M, Rq), respectively. Hence, these sequences become, respectively, uniformly bounded and continuous subsets in the function spaces C2+1 (M,R9) and C'(M,R9). By the Ascoli-Arzela theorem, there exist a Subsequence {tak} of {t=} and functions
u(-, To) E C2+a(M,Rq) and Btu(-, To) E Ca(M,Rq)
such that the subsequences t=k)}
and
respectively, converge uniformly to Since for each t,k , we have
{Otu(-, tik)},
To) and Btu(-, To), as tik - To.
Otu(, t=k) = T(u(-, ttk)),
we also get at To
consequently, we see that (4.27) has a solution in M x [0, TO]. If we apply Theorem 4.10 with To) as an initial value, we see that there exists a positive number c > 0 such that an initial value problem (x, t) E M X (To, To + E),
u(x, 0)
=u(x,To)
has a solution u E C2+a,I+a12 (M X [To, To + E], N) in M x [To, To + E].
We see from their constructions that this solution and the previous solution u in M x [0, To] coincide in M x {To}, and that they give rise C2+a,l+a/2(M x [0, To + E], N) to the initial value to a solution u E problem (4.27). Noting the arguments regarding differentiability of the solutions in Theorem 4.10, we see that it is CO° in M x (0, To + E). Consequently, (4.27) has a solution in M x [0, To + c]. This contradicts the definition of To. Thus, To = oc. The uniqueness of u readily follows from Theorem 4.15. 13
4. EXISTENCE OF HARMONIC MAPS
150
4.4. Existence and uniqueness of harmonic maps Let (Al, g) and (N, h) be compact Riemannian manifolds of dimensions m and n, respectively. Given f E C2+a(M, N), we consider the initial value problem of the parabolic equation for harmonic maps (4.28)
= r(u(x, t)), f(x,t) 8u
(x, t) E AI x (0, oo),
u(x, 0) = f W.
As seen in Theorem 4.16, if N is of nonpositive curvature KN < 0, there exists a unique time-dependent global solution u E C2+a.I+a/2(AI x [0, oo), N) fl C, (Al x (0, oo), N) to (4.28). In this section, we verify that this u converges to a harmonic map that is free-homotopic to f.
PROPOSITION 4.17. Assume that N is of nonpositive curva-
ture KN < 0. Given an initial value f E C2+a(M, N), let u E C2+a,I+a12(M
x [0, oo), N) fl C,=(Al x (0, oo), N) denote the timedependent global solution to (4.28). Then there is a sequence {ti} of real numbers with ti - oo such that a sequence of COO maps {u(-, tj)} converges to a harmonic map U. E COD (M, N) five-homotopic to f. PROOF. As in the proof of Theorem 4.16, Proposition 4.14 implies that E R}, E R}, respectively, form uniformly bounded and equicontinuous subsets in
C2+a (M, N) and Ca (M, N). From the Ascoli-Arzela theorem, we see that there exists a sequence { t= } of real numbers and u,,,. E C2+a(M, N) such that t)}
and
t)},
respectively, converge uniformly to uO0 and 8tu,', as t -- ao. As noted in Proposition 4.4, since we have 8tuOC = 0, we get
T(u(., ti)) =
ti) - r(u,°) = 0;
namely, we see that uO0 satisfies the equation for harmonic maps. Since u(., ti) uniformly converges to u,°, ti) and u,0 are freehomotopic to each other for a sufficiently large ti. In fact, noting that Al is compact and that u(x, t3) and u,C(x) are contained in the same coordinate neighborhoods in N for sufficiently large ti, we can easily construct a free-homotopy as above. On the other hand, since u(., t) is continuous in t, f = u(., 0) and u(., t) are free-homotopic; hence, f 0 is free-homotopic to u°C.
4.4.
EXISTENCE AND UNIQUENESS OF HARMONIC MAPS
151
Proposition 4.17 completes our objective to prove Theorem 4.1 due to Eells and Sampson. We state the following. COROLLARY 4.18. Let (M,g) and (N, h) be compact Riernanntan manifolds and let N be of nonpositive curvature. Then any con-
tinuous map f E C°(M, N) is free-homotopic to a harmonic map u00 E Coo (M, N).
PROOF. There exists a C°° map f E C' W, N) homotopic to f (see Exercise 4.9 at the end of this chapter). Then apply Theorem
4.ltof. If we examine Proposition 4.17 more thoroughly, we see that we t) converges uniformly to upo as t -i oo without can prove that choosing a sequence {ti}. In other words, the following holds. THEOREM 4.19. Let
N
of
be
nonpositive
curvature
Kn < 0. Given an initial value f E C2+a (M, N), let u E
C2+a,i+a/2(M x [0'X ), N) n C, (M x (0, oo), N) denote the timedependent global solution to (4.28) and set ut(x) = u(x, t). Then ut E C' (M, N) uniformly converges to a harmonic map u,,p E CO° (M, N) that is free-homotopic to f. We outline the idea of the proof of this theorem following Hartman [5]. First, we note the following lemma. LEMMA 4.20. Let N be of nonpositive curvature K1e < 0, and let
f E C2+a(M x [0, 1], N). For each s E [0,1], let u(x, t, s) denote the solution to the initial value problem (x, t, s) = T(u(x, t, s)),
(x, t) E M X (0, T),
lu(x, 0, s) = f(x, s). Then the functions
v(t, s) = sup I zEM
2
as
(x, t, s) I
v(t) =
,
su
sup xEM,sE[O,1J
2
a8
are both monotone nonincreasing functions in t.
PROOF. We see that the solution u(x, t, s) to the initial value problem is C2 in s E [0,1] (Hartman [5]). Set v(x, t, s)
_ au =
I
a8 (x' t' S)
2
rn
G.. ha,o(u)
a,9=1
aua auj 018
(98
4. EXISTENCE OF HARMONIC MAPS
152
If we note the assumption on the curvature KN < 0, we can verify, using a similar computation to the proof of Proposition 4.2 (2), (4.29)
8 A - 8t} v=
IV
2
as I
m
- t=1
RN du(et) au 8s
8s ,
du(ei)
Here, RN denotes the curvature tensor of N, and {e;} represents an orthonor m al basis for the tangent space TT M at each x E M. Hence, from the maximal principle for the heat equation, we get for
each aE [0,1]and0
x
v(t2i s) = m v(2, t2, s) <
ti, s) = v(tl, 8).
This implies v(t2) < V(t1)-
LEMMA 4.21. Let N be of nonpositive curvature & < 0, and let
f E C2 (M x [0,1], N). Let f8(x) = f (x, s) denote the homotopy defined by f, and let o,(x,t) be a geodesic joining fo(x) and f, (x) in N. For each s E (0, 11, let u(z, t, s) denote the solution to the initial value problem
(x, t, s) = r(u(x, t, s)), 1u(x, 0, s) = a(x, s).
(x, t) E M x (0, T),
Denote by a geodesic in N joining u(z, t, 0) and u(x, t, 1). Assume that the geodesic is homotopic to a curve given by
1u(x, (1!(X, 8) =
3s)t, 0),
a(x, 3s - 1), u(x, (3s - 2)t,1),
0 < s < 1/3, 1/3 < s < 2/3, 2/3 < s < 1.
Also denote by d(u(z, t, 0), u(x, t, 1)) the length of the geodesic. Then
0(t) = sup d(u(x, t, 0), u(x, t,1)) =EM
is a monotone nonincreasing function in t.
PROOF. Since N is of nonpositive curvature KN < 0, the geodesic a(x, t, s) is uniquely determined. Since the curve s 14 u(x, 0, s) = a(x, s) is a geodesic for each x E Mat t = 0, we have, for any s° E [0,1] , d(u(x, 0, 0), u(x, 0, 1)) =
1 I
JQ
o (x, 0, s) Ids =
1
JO
v(x, 0, s° )ds.
4.4.
EXISTENCE AND UNIQUENESS OF HARMONIC MAPS
Consequently, we get 9(0) = supXEM
v(x, 0, s°) =
153
v(0). On the
other hand, for any t > 0, d(u(x, t, 0), u(x, t, 1))
=
Ji
I
as
(x, t, s) Ida
=
Joo
i
v(x, t, s)ds
holds. Hence, from Lemma 4.20, we see that
9(t) <
j
v(t, s)ds <
u(t) <
v_(0)
= 0(0).
For general 0 < ti < t2 < T, we may apply the same arguments to
0
t - tI.
We are now ready to prove Theorem 4.19. First, we apply Lemma
4.21 to the case in which we set fo = fl, fi = u,,.. Since u,,,, is a harmonic map, 7-(u(x, 0,1)) = r(ua,(x)) = 0. Hence, u(x, t, l) = U00 (x) for any t. On the other hand, u(x, t, 0) = u(x, t) follows from the definition, since u(x, 0, 0) = f (x). Consequently, from Lemma 4.21,
d(ut, u..) = sup j(u(x, t), u.(x)) = sup d(u(x, t, 0), u(x, t,1)) zEM
zEM
is a monotone nonincreasing function. Proposition 4.17 implies that there exists a sequence { t= } with tt -, oc such that
d(u(., t,), u00) - 0 (t, - oc). Hence, we have
d(ut, u ) -* 0 (t When existence of harmonic maps, namely, existence of solutions to the equation for harmonic maps, is shown, the next logical question is to ask if the solutions are unique. However, uniqueness does not
hold when the harmonic maps are constant maps. It is also readily seen that uniqueness does not hold when the images of the harmonic maps are closed geodesic, as indicated in Figure 4.2. Nevertheless, if we exclude the above cases, we can prove the uniqueness of the solutions in the theorem of Eells and Sampson. THEOREM 4.22 (Hartman). Let (M, g) and (N,h) be compact Riemannian manifolds, and assume that N is of nonpositive curvature. Then the following holds: (1) Let uo, ui E CO°(M, N) be free-homotopic harmonic maps. Then uu and uI are free-homotopic to each other through a family of harmonic maps {u, I s E [0, 1J} C C°°(M, N). Consequently, the set of harmonic maps free-homotopic to no is connected.
4. EXISTENCE OF HARMONIC MAPS
164
FIGURE 4.2
(2) In particular, if N is of negative curvature KN < 0, uniqueness of harmonic maps in the following sense holds. Let uo, ul E C' (M, N) be free-homotopic harmonic maps. Then uO, = ul holds wept for the following cases (1), (ii): (1) uQ is a constant map. Then ul is also a constant map.
(ii) The image uo(M) of M under uo coincides with a closed geodesic c. In this case, the image ul (M) also coincides with c. Fur, for any point x E M, ul(x) is obtained by moving uo(x) a constant distance along c in the same direction. OUTLINE OF THE PROOF. (1) The homotopy {u8 I s E [0, 1]} by
a family of harmonic maps is constructed as follows. Given the harmonic maps uo, u1, we denote by f : M x [0,11- N a C°° homotopy between uo and ul. At a fixed s E [0,1], we consider the initial value problem (x, t, s) = T(u(x, t, s)), 1u(x, 0, s) = f (x, s).
J
(x, t) E M x (0, oo),
Since N is of nonpositive curvature KN < 0, Theorem 4.16 implies that this initial value problem posses a global time-dependent solux [0, oo), N) fl (Coo (M x (0, oo)), N). From Theorem tion u E 4.19, the solutions u(x, t, s) converge uniformly to a harmonic map as t .-+ oo. Then u9(z) = u(z, t, s) give rise to the desired homotopy by harmonic maps. In fact, we can even prove the following stronger statement. Namely, by modifying the above homotopy u., we can get a C°° homotopy u: between uo and ul by harmonic maps. Furthermore,
4.5. APPLICATIONS TO RIEMANNIAN GEOMETRY
155
the curve c' : s
us (x) is a geodesic in N for each x E M and its length is constant independent of x E M. (2) Let f (x, s) be a homotopy between uo and u1 such that each f (', a) E C°° (M, N) is a harmonic map and f (x, ) is a geodesic joining uo(x) and u1(x). We apply Lemma 4.20 to this f (x, s). From (4.29) and v(x, t, s) being a constant, we get
/RN d(f(ei). Lf
) 8$f
,
cU(ei)) = 0,
1:5 i < M.
Since KN < 0, we see either that as
0 or that df (TTM) is a Bf =0 subspace of at most one dimension in T f(s) N for each s . If holds, we get uo = ul; hence, it is unique. Otherwise we can readily verify that either (i) or (ii) holds (see Proposition 4.24). Fbr more detail, the reader may consult the original paper of Hartman [5].
4.5. Applications to R.iernannian geometry As we saw in Chapter 3, geodesics, minimal submanifolds, isomet-
ric diffeomorphisms and some holomorphic maps constitute typical examples of harmonic maps. Consequently, we can study structures of Riemannian or Kiihlerian manifolds using existence and properties of harmonic maps. In this section, we take up such applications of harmonic maps. In what follows, unless mentioned specifically, we let (M, g) and (N, h) be Rieinannian manifolds of dimensions m and
n, respectively. We begin with a remark on Weitzenbod formula regarding the energy density of harmonic maps.
PROPOSITION 4.23. Let u : M - be a harmonic map. Then the following holds regarding the energy density e(u) of u:
Le(u) = I VuI2 + Q(du).
(4.30)
Q(du) is given by m
rra
RicM (ei, e,)e,), du(et))
(du (
Q(du) _
t=1
s-1
-
"t
(Rh'(du(es)=du(e))du(e,),du(ej)).
4j-1
Here, A is the Laplace operator of M, and Ricm and RN denote the Ricci tensor of M and the curvature tensor of N, respectively.
156
4. EXISTENCE OF HARMONIC MAPS
Furthermore., {e;} is an orthonormal base for the tangent space T1M at each z E M, and (, ) rep esents the natural fiber metric in u-'TN.
The proof is the same as in Proposition 4.2, and is left to the reader. From this proposition readily follows
PROPOSMON 4.24. Let M be a compact Riemannian manifold iktrtherrrtore, assume that the Rica tensor of M is positive semideftnite, i.e., Rich > 0, and that N is of nonpositive curvature KN < 0. Then, for a harmonic map u : M - N. the following hold: (1) Vdu = 0; namely, u is a totally geodesic map (see Exercise 4.5 at the end of this chapter) and e(u) is a constant. (2) If the Ricci tensor of M is positive definite, i. e., Ric" (x) > 0 at a point z E M, then u is a constant map. (3) If N is of negative curvature KN < 0, u is a constant map or the image of u coincides with the image of a closed geodesic of N. PROOF. (1) From Green's theorem, we have
IM
L e(u)dpg = 0.
Hence, the integral on the right hand side of (4.30) is also 0. On the other hand, each term on the right hand side is nonnegative from the
assumptions .ii"' _ 0 and KN <_ 0. Consequently, each of them must be 0 at each point; implying that Vdu = 0 and Q(du) = 0. Since we get Lie(u) = 0 from theme, we see that e(u) is a harmonic map in a compact R.iemannian manifold M. Hence, e(u) is a constant.
(2) Since (1) implies Q(du) = 0, we get, with respect to an orthanortnal base {e, } for the tangent space TAM at x E M. (cm
` RicM (eiei)el ,du(et) } = 0.
If Rich (x) > 0 holds, the derivative du. of u at x must equal 0. This implies that e(u)(x) = 0. Since e(u) is a constant, we get e(u) . 0. Hence, u is a constant map. (3) Since Q(du) = 0, we have, with respect to an orthonormal base {e{} for the tangent space T,,M at x E M, (RN (du(e;),du(ep))du(e,),du(e;)) = 0,
1 < i, j < m.
On the other hand, since we have KN < 0 from the assumption, the sectional curvature K(o) of any two-dimensional subspace
4.5. APPLICATIONS TO RIEMANNIAN GEOMETRY
157
o C TTM of T,,M is negative. Hence, du(e;) and du(ej) are never linearly independent. This implies at each x E M that d(x) = dim du,, (TTM) < 1.
M, e(u) 0; hence, u is a constant map. Otherwise, d(ay) = 1. Noting that u is totally geodesic, we readily we that the image of u coincides with the image of a closed geodesic in
Now if d(z) = 0 at an x
0
N.
In Chapter 2, we saw that the fundamental group irl (M) of a Riemannian manifold M of positive curvature is a finite group. In other words, if M is of positive curvature Km > 0, its fundamental group is a small group. On the other hand, the fundamental group of a compact Riemannian manifold of negative curvature is known to be a large group. Here, as an application of harmonic maps, we verify that a nontrivial commutative subgroup of the fundamental group 71 (M) of M with negative curvature is an infinite cyclic group. Namely, the following holds. THIEOREM 4.25 (Preissmann). Let (M, g) be a compact and connected Riemannian manifold, and assume that the sectional curvature KM of M always satisfies Ks,t < 0. Then any nontrivial commutative subgroup of the fundamental group ir1(M) of M is an infinite cyclic group.
PROOF. Let Tl (M, xo) be the fundamental group of M with base xo E M. Furthermore, assume that the two elements a, b of 1r1 (M)
are commutative; namely, ab = ba holds. By the definition, a, b are the homotopy classes of loops with xa as the base point. We express loops representing a, b by the same symbols a, b. Since they are commutative, there is a homotopy f : [0,1] x [0,11 -+, M between loops a - b : [0,1) -+ M and b a : [0, 1] -+ M. Here, this homotopy keeps the base point fixed throughout the deformation from a - b = f 0) to b a = f 1). In other words, noting ,f (0, s) = f (1, s),
s E [0,11,
we readily see from Figure 4.3 that f defines a continuous map T2 - M from a two-dimensional torus T2 into M. Applying Corollary 4.18 to j, we see that f can be deformed free-homotopically to a harmonic map u : T2 -- M. In this case, the loops a - b and b - a are also free-homotopically deformed. The base point is not necessarily fixed throughout this deformation, but we
4. EXISTENCE OF HARMONIC MAPS
158 1
h
a
a
A6
0
a
a
b
FIGURE 4.3
point out that the loops corresponding to a b and b a have the same base point in each stage of the deformation. Since M is of negative curvature KN < 0 from the assumption, Proposition 4.24 implies that either u is a constant map or the image u(T2) of u coincides with the image of a closed geodesic c passing through a point x1 E M in M. Consequently, if u is not a constant map, the loops corresponding to a - b and b a both cover c multiple times in the fundamental group iri (M, x1) of M based at x1. Thus, a b and b a are contained in the infinite cyclic subgroup generated by c. This implies that a b and b - a are both contained in an infinite cyclic subgroup of iri (M, xa). From the above arguments, we see readily that any nontrivial commutative
subgroup of the fundamental group rri(M,XO) of M is an infinite cyclic group.
0
In a like manner using the uniqueness of the harmonic maps, one can show that the set of all isometric transformations of a compact Riemannian manifold of negative curvature forms a small group. In general, it is well known that the set of all isometric transformations of a compact Riemannian manifold M of negative curvature is a Lie group. In particular, if KAJ < 0, this group is a finite group. Namely, the following holds. THEOREM 4.26. Let (All, g) be a connected and compact Riemann-
ian manifold. Furthermore, assume that the sectional curvature KM of M always satisfies K11 < 0. Then the group G of the isometric transformations of M is a finite group.
PROOF. First, we note that an isometric transformation of M homotopic to the identity map is the identity map if K11.t < 0. In fact, let f be an isometric transformation homotopic to the identity
4.5. APPLICATIONS TO RIEMANNIAN GEOMETRY
159
map. Since f is a harmonic map, Theorem 4.22 due to Hartman implies that it is the identity map. From this readily follows that G is discrete. Since G is compact, it is finite.
0
If we note that a holomorphic map between Kahlerian manifolds is harmonic, we get the following. TrEoItzM 4.27. A complex submanifold of a KaJelerian manifold is a minimal submanifold. PROOF. Since M is a complex submanifold of a Kahlereian manifold N, there is an analytic imbedding u : M , N. M is a Kihlerian manifold with the induced metric. Then u becomes a harmonic map from M into N. On the other hand, as seen in Example 3.15, an isometric embedding being harmonic is equivalent to M being a minimal submanifold of N. This gives the desired conclusion. 0
The existence problem of analytic maps between complex manifolds is an important research topic in complex analysis. Especially since analytic maps between Kiihlerian manifolds are harmonic, we
can study the existence problem from the viewpoint of harmonic maps. Indeed, much has been done in this respect. Fbr example, as an application of the theorem of Eells and Sampson, Siu [23] proved the following in 1980. THEOREM 4.28 (Siu). Let N be a Keihleri,an manifold obtained as a quotient manifold of an irreducible bounded symmetric domain. Let N be compact and at least two complez dimensional. Furthermore,
assume that M and N are of the same homptopy type. Then M is biholomophic or anti-biholomorphic to N.
This theorem asserts that the complex structure on a compact Kiihlerian manifold N obtained as a quotient manifold of an irreducible bounded symmetric domain is determined by the homotopy type of the manifold, except for the complex one-dimensional case. It is called the strong rigidity theorem of such Kahleriau manifolds as above. The essential parts of the proof consist of an improvement of the Witzenbock formula for Kahlerian manifolds and the existence
theorem of harmonic maps due to Eells and Sampson. If M and N have the same homotopy type, there is a homptopy equivalence map from M into N. The desired biholomorphic or antibiholomorphic map is obtained by deforming this homotopy equivalence map to a harmonic map. This strong rigidity theorem of Siu is one of the most successful applications of the theory of harmonic maps.
4. EXISTENCE OF HARMONIC MAPS
160
Summary 4.1 The heat flow method and its idea to obtain the critical points of the energy functional e. 4.2 Existence of time-dependent local solutions to the initial value problems of the parabolic equation for harmonic maps. The relationships between the growth rate of solutions and the curvature, and the role of the Weitzenock formula for the estimation. 4.3 Existence of time-dependent global solutions to initial value problems of the parabolic equation for harmonic maps and their convergence to harmonic maps. 4.4 Due to the Eells-Sampson theorem, any continuous map from a compact Riemannian manifold into a compact Riemannianl manifold of nonpositive curvature are free-homotopically deformed to harmonic maps. 4.5 A theorem of Hartman regarding the uniqueness of harmonic maps. A theorem of Preissmann regarding the fundamental group of Riemannian manifolds of negative curvature.
Exercises 4.1
Let (M, g) and (N, h) be Riemannian manifolds of dimen-
sion m and n, respectively. Denote by u E CO° (M, N) a COD map from
M into N. Consider the tensor product TM' 0 u-'TN of the cotangent bundle TM* of M and the induced bundle u 1TN by u from the tangent bundle of N. Denote by V the connection in TM' ® u`1TN compatible with its natural fiber metric (, ). Prove the following (1) Given T E r(TM'(&u-1TN), its second covariant differential
VVT E r(TM' 0 TM' 0 TM' 0 u-TN) is a tensor field of type (0,3) which takes its value in the induced bundle u-1 N. Furthermore,
VVT(x, Y, z) = (V x(V yT))(Z)
- (Vv,ryT)(Z)
holds for x, Y, z e r(TM). (2) Denote by V the connection in TM* 0 u-'TN and by 'V the induced connection in u-'TN. Set RV(X, Y) = VxVy - DY©x - V[X,YJ
R'v(X,Y)
QXVY -I V, VX -, ©(X.YJ.
Also denote by RM, R'' the curvature tensor of M, N, respectively.
EXERCISES
161
For T E F(TAI* (& u-1TN) and X, Y, Z E r(TAI ),
(RV (X,Y)T)(Z) = R'V (X,Y)(T(Z)) - T(R I (X.Y)Z) holds.
(3) With respect to local coordinate systems (x'), (y°), express T, VVT by m
n
?,'dx' (9)
T
i=1 a=1
M
o u,
n
vtvjTk dx' ®drj 0 dxk ®aa o u.
VVT = i,j,k=1 a=1
Then
vtv1Tk - vjVi k =
-
aa0au7 T
RA1ijkT1` + 1=1
RNa.-,.6 x= AX-0 v,i',a=1
holds. Here, R11 jk, R' ,.,,6 are the components of RAr, RN with respect to (x'), (y°). This identity is called the Ricci identity. 4.2 Let (M, g) and (N, h) be Riemannian manifolds of dimension m and n, respectively. Let u : M x [0, T) - N be a solution to the parabolic equation for harmonic maps
,jT(x,t)=,r(u(x,t)),
(x, t) E AI x (0, T).
Set ut(S) = u(x, t). Prove the Weitzenbock formula alC ut)
ar
=
at
"'
2
IV2!Lt I
+
/RN (du(e2)t)
,du(et)}
I2 Here, RN is the curvature tensor of N, and {et} represents an orthonormal base for the tangent space TM at each f o r c(ut) = 11
at
x E Al.
4.3 Let Al denote a compact submanifold of a Riemannian manifold N. Denote by U an open subset of the normal bundle TAI1 of AI consisting of all the normal vectors whose magnitudes are less than e. Show that, for sufficiently small f > 0, the map
exp:U -N obtained by restricting the exponential map to U gives rise to a differential diffeomorphism from U onto the submanifold U = exp(U) of N. This U is called a tubular neighborhood of AI in N.
4. EXISTENCE OF HARMONIC MAPS
162
Let M1, M2, M3 be Riemannian manifolds, and let f, : M1 - M2 and f2: M2 M3 be C°° maps. Regarding the second fundamental form Vd(f2 o fl) and the tension field r(f2 o fl) of the composition f2 o fl,
4.4
Vd(f2 o fi) _ Vdf2(df1, df1) + df2(Vdf1), r(f2 o f1) = trace Vdf2(dfi,df1) + df2(r(f1))
Regarding a COO map u : M - N between R.iemannian manifolds M and N, show that the following are equivalent: (i) The second fundamental form of u satisfies Vdu = 0. (ii) u maps a geodesic c in M onto a geodesic u o c. The aline parameter of c also represents an affine parameter of u o c. Based on these, we call u totally geodesic if Vdu = 0.
hold. 4.5
4.6
Let M be a compact Riemannian manifold and let 0 <
a < 1. Prove that the Holder spaces
C11,a12(M X [0, TJ, R4) and
C2+a,1+a/2(M x [0, T], RQ) are Banach spaces with respect to the norm defined in (4.16).
Let M be a compact Riemannian manifold and set Q = M x [0, 11. Assume that 0 < a' < a < 1. Given a positive number e,
4.7
choose a C°° function ( : R - R as being ((t) = 1 (t < e), ((t) =
0(t>2e); 0<((t)<1, I('(t)I<2/e(tER).Prove that, given a w E Ca'a/2(Q, R') with w(x, 0) = 0, there exists a constant C > 0 independent of a and w such that the following holds: IbwIQ',a'/2)
<
Ce(a-a')/2IwIQ,a/2).
4.8 Let M be a compact Riemannian manifold, and let u E CO (M x (0, T)) n C2'1(M x (0, T)). Let C denote a constant and assume that u satisfies the inequality
8 < L u+Cu inMx(0,T).Show that ifu<0inMx{0}, then u<0always holds in M x [0, T). 4.9 Let M, N be compact C°° manifolds, and let f : M -> N
be a continuous map. Show that there is a CO° map f : M - N, free-homotopic to f. 4.10 Let M be compact Riemannian manifold of nonpositive curvature Km < 0. Show that
irk(M)=0, k>2, where lrk(M) denotes the k-dimensional homotopy group of M.
APPENDIX A
Fundamentals of the Theory of Manifolds and Functional Analysis As the requirements for reading this book, we assumed some fun-
damental facts in the theory of manifolds and functional analysis. Regarding manifolds, for example, we assumed the definitions of C40 manifolds and their tangent bundles, the derivative of a map between
manifolds and partition of unity etc. As for functional analysis, we used the inverse function theorem in Banach spaces, fundamental solutions to heat equations and the Schauder estimate regarding solutions to elliptic partial differential equations and linear parabolic partial differential equations. Here we give a brief summary of these topics as appendices for the sake of convenience of the reader. Although the reader should be reasonably well informed about manifolds, details of the part on functional analysis may be skipped.
A.I. Fundamentals of manifolds (a) d' manifolds. Let M be a Hausdorff topological space satisfying the second axiom of countability. Let {(U,,,, c¢a)}aEA be a
given set of pairs consisting of an open subset U,, in M and a map 0,,, : U0 -e Rm from U,, into the in-dimensional number space R", where A indicates an appropriate set of indices. Then M is called an r-dimensional C' manifold or a smooth manifold if the set satisfies the following conditions: (i) {U,l}QEA forms an open cover of M; namely, M = U0EAUa holds.
(ii) For each a E A, the image 0a(U,,) is an open subset of Rm, and 0 is a homeomorphism from U. onto Sao. (U,, ). (iii) For any ct,j3EAwith UaflUO00,the map
o,aoan1 : di.(U0 nU8) i'os(UO nUs) 163
APPENDIX A
164
is a C°° map from 0a(Ua fl U,Q) onto 0p(Ua n Up), both of which are open subsets in JRm. The family {(U0, Oa)}aEA is called a C°° coordinate neighborhood system of M, and each (Ua, a) is called a coordinate neighborhood. Given a point p in a coordinate neighborhood &a(p) is a point in Rm. Setting xa = x` o m) with respect to the coordinate functions (xa, ... , xa ), we can write qSa (p) = (xa (p), ... , xa (p))
(xa(p), ... , x. (p)) are called local coordinates of p in the local coordinate neighborhood Each xa = x4 o 0. is called a coordinate function in U,,,. The m-tuple of coordinate functions (xa,... , xa) is called a local coordinate system in (Ua, 4a). Given p E U. n Ua, we p in two ways: express the local Oa(p) = (xa(E'), ... , xa t!'))+
0,8 (p) = (x01 (p),...
,
x (p))
Then condition (iii) in the definition of a CO° manifold implies that the coordinate transformation 0# o 0a-1(xa(p), ... , xa (p)) = (x,1e(p), ... , xm (p)),
p E Ua n U0,
is a C°° map. Given an open subset U in M, U becomes a COO manifold with respect to the restriction ((U n U0),i0a I (U n Ua)}aEA of the C('O coordinate neighborhood system {(U0, 00)}aEA to U. This U is called an open submanifold of M. Let M and N be C°O manifolds of dimensions m and n, respectively. Denote by {(Ua, Oa)}aEA, {(U0,''0)}$EB the local coordinate neighborhood systems of M and N, respectively. Define a map 0a xz/iQ:Ua x 00 (p, 4') = (-0a (p), 00 (q) ), (p, 4) E U. x V,3. {(U0 X V0, 0a X 10)}(a,p)EAxB give rise to C°O coordinate neighborhood systems in the product topological space M x N; hence, M x N is a COO manifold. This is called the product manifold of M and N.
(b)
Tangent bundles. Let f : U - JR be a function defined in
an open subset U of a COO manifold M. We call f a Cr differentiable
function (1 < r < oo) if, given a coordinate neighborhood (Ua, 0a) with U n Q. 0 0, the function defined in the open subset U n Ua 0 0 of JRm by
f o0a1 : O(UnUa) --R is a C' function.
A.I. FUNDAMENTALS OF MANIFOLDS
165
Given a point p in a C°° manifold M of dimension in., denote by C°°(p) the set of all CO° functions defined in open neighborhoods of
p. Let U1 and U2 be the domains of definition for fl, f2 E C°(p), respectively. The sum and product f1 + f2, f1f2 E C°°(p) can be considered as functions defined in the domain Ul fl U2. Also given a
scalar a E R, the scalar product a f 1 E C(p) is a function defined in U1. A correspondence v : C°'° (p) - lR that assigns a real member v(f) to f E C°° (p) is calm a tangent vector to ?111 at p, if it satisfied the following conditions: (linear property). (i) v(afi + bf2) = av(f1) + bv(f2) (Leibniz rule). (ii) v(f1f2) = v(fi)f2(P) + fl(P)v(f2) Here, fl, f2 E GO° (p) a, b E R. As seen from the definition, the tangent vector of a C°° manifold is a generalization of the directional derivative in the Euclidean space. Let vi, v2 denote tangent vectors to M at p. We define the sum f1 + f2 and the scalar product with a E lR by
(av1)(f) = avi(f), f E C°°(p) (f1 + f2)(f) = vi (f) + v2(f), With these operations, the set of all tangent vectors to M at p naturally becomes a vector space over R. This vector space is denoted by
T1 and is called the tangent vector space of M at p or simply the tangent space.
Let (U0, 0a) be a coordinate neighborhood that contains p, and denote by (x', ... , xa) the local coordinate system in (U., 4a). Since, given a function f E C°° (p) defined about p, f o 0; 1 is a C°° function defined about 0,,,,(p) E It', we may consider the partial derivatives 8(f o 0;1)(0a(p)) with respect to the coordinate functions x1 in JRt. Then the map
a ():C00(P)R, °
l
m,
P
obtained by setting
a(f) cr
p
i _ ax_a
(0, (P) ),
defines a tangent vector of M at the point p. Furthermore, (A.I)
APPENDIX A
166
are linearly independent as elements of TpM, and any element V E TpM can be expressed as a linear combination of these elements as m
vv(xa
x CXa
MP
We can verify this using (i), (ii). Hence, the tangent space TpM of M is an tn-dimensional vector space, and (A.1) forms a base for TFM.
Given a tangent vector v E TpM, we denote ' - u(a'), and call the n-tuple (f 1, . - . , '") the components of v in the local coordinate system Given a C°O function f E C°° (p), we define
dfp(v) = v(f),
v E TPAM.
Then df , bewmes a linear functional fP : Tp M -, R of the tangent space T.M. In other words, df, is an element of the dual space TPM* of TpM. We call this dff the derivative of f at p. Especially, if we consider the derivatives of the coordinate functions x' of a local coordinate neighborhood (U,,, we get s
(d4)P
p
=
ax
< i,1 < M.
('0.(p))
Hence, { (dxa ')p,. .. (d2a )P )
forms a bane for TM* dual to the base (A.1) for TpM.
(c) Derivative of maps. A continuous map co : M -, N from an rn-dimensional C°O manifold M into an n-dimensional C°° manifold N is said to be a C' map (1 < r < oo) if for any Cr function f : V --+ R defined in any open subset V of N, f o p becomes a Cr map in the open subset cp-'(V) in M. Given any point p in M, take coordinate neighborhoods (U., .), (Va, 0.\) in Al, N, respectively, such that p E Ua and So(UQ) C V. Then v being C' is equivalent to that the map 0a O io o 0.-' : 0,(U.)
'0 a(V.\)
is a Cr map from the open subset 0a (Ua) of R' into the open subset a(Va) of R". In particular, if a C°° map Sp : M -, N is one-toone, onto and the inverse map V-1 : N -- M of V is also C°°, cp is cafed a C°° diffeamorphism from Al onto N. When there is a C°° diffeomorphism from Al onto N, M and N are said to be mutually C°°
A.I. FUNDAMENTALS OF MANIFOLDS
167
diffeomorphic or simply difeomorphic. Diffeomorphic COO manifolds have the same dimension.
Let co : M - N be a CO° map from a C°D manifold M into a C°° manifold N. Given p E M, let f be a C°° function defined about W(p) E N. The composition f o cp is a C' function defined around p. Given v E TPM, set
'`'c'P(v)(f) = v(f o'P),
f E C°°(cp(P))
Then pp(v) is a tangent vector to N at the point cp(p), and we obtain a linear map
dcpp:TPM-+T'p{ )N from the tangent space TM into the tangent space TT(p)N. Call this
d'P the derivative of the map cp : M N at p. Let (x...... x T) denote a local coordinate system in a local coordinate neighborhood (Ua, 4Q) about p and (ya, ... , y") a local coordinate system in a local coordinate neighborhood (V,1, ,Ox) about ip(p). From the definition, we have &-PP
EV(P) P-1
ax=
ate=
aY
1
In other words, dcp. is nothing but a linear map given by the Jacobian matrix J(W)(P) =
W
axi
(P),
1 < i < m, 1 < j < n,
with respect to the bases
\"u/ P
( -
a )I, in
P
(ay.} a W(P)
,...'
"A tP)
for TPM and T,p(p) N.
Given a C°O map cp : M N, cp is called an immersion if the derivative dcpP : TPM N of cp is injective at each point p E M. When there is an immersion cp from M into N, we call M an immersed suhmanifold of N. In particular, if the immersion cp is a homeomorphism from M into the subset V(M) with the relative topology in N, we call 'o an imbedding. We call this M an imbedded submanifold of N, or simply a submanifold.
APPENDIX A
168
Let M1i M2, M3 be C°° manifolds, and let Cpl : M1 --> M2, W2 : M2 - M3 be COO maps. Then the composition map 402 o Ipi : Ml M3 is a C°° map from M1 into M3, and the composition rule d((p2 o Spi)p = (dco2),pi(p) o (d401)p
holds regarding the derivative at p E M1. This is nothing but the product formula of the Jacobian matrices J(W2 o SPi)(P) = J(cp2)(41(P)) AVOW-
(d)
Partition of unity. Let M be a C°° manifold. An open
cover
of M is said to be locally finite if for any point p E M,
there is a neighborhood W of p such that W f1 U. 96 0 for only a finite number of a E A. Let {Ua }aEA and {Va }$EB be two open covers of M. {Va}sEB is said to be a refinement of {Ua}aEA if for any ,Q E B, there is an a E A such that VV C Ua. A Hausdorff space is said to be paracompact if any open cover has a locally finite refinement. In particular, a compact Hausdorff space is paracompact. A C°O manifold M is also paracompact if it satisfies the second axiom of countability.
Given a function defined in M, the closure of the subset {p E M I f (p) 54 0} of M is called the support of f. For an open cover {Ua}aEA of M, a family of COO functions { fa}aEA defined in M is called a partition of unity subordinate to the open cover {Ua}aEA if the following conditions hold.
(i)For each a,0< fa(p)<1(pEM). (ii) For each a, the support of fa is contained in Ua. (iii) For each point p E M, Ear--A fa(P) = 1.
Note here that the sum in (iii) is a finite sum from (ii) and local finiteness of {Ua}aEA. The following is well known.
THEOREM A.1 (Existence of partition of unity). Given a locally finite open cover {Ua}aEA of M, if the closure Q. of each U. is compact, there exist a partition of unity { fa}aEA subordinate to {Ua}aEA.
The proof of the theorem is based upon the following fact. Given a compact subset K and an open subset U containing K in M, there
exists a CO° function f defined in M such that (i) 0 < f (p) < 1 (pEM),(ii) f(p) = 1 (p E K) and f (p) = 0 (p E M K). These
-
functions are often used to "cut and paste" geometric structures on COO manifolds.
A.I. FUNDAMENTALS OF MANIFOLDS
169
(e) Tangent bundles. Let Al be an 7n-dimensional COG manifold. Let TAI represent the set of all tangent spaces TpAI. p E Al. Namely, set
TAI = {(p,v) p E Al. v E TTAl}. Denote by r: TAI - Al the standard projection given by 7r(p, v) = p,
(p. v) E TAI.
TAI naturally is a 2m-dimensional COG manifold, and it becomes a
CO° map from TAI onto M. This CG manifold TAI is called the tangent bundle of Al, and zr is called the projection from TM onto M. The C°° structure of TM is defined as follows. Let {(U0, via) }aEA
denote the C°° local coordinate neighborhood systems in M. Denote by (xli, ... , xa) the local coordinate system in (Ua, 0a). Given 7r-1(Ua) _' R" by (p, v) E 7r-'(p), define a map ,ba (p, v) = (ca (p), d0a (v) )
= (xa(p), ... , xa (p), dx'(v), ... , dxa (v))-
Clearly, a is a diffeomorphism from it-1(Ua) onto the open subset (ir-I(U0))
of R2m. Also, given (p,v) E 7r-I(U.)flzr-1(Up), we have
(0Qo0a1(0a(p)),d(0ao0a1)(dO,(v))) Therefore, d(00 o 0;1) is CO°, since 4 o 0. 1 is C'. From these, we can readily verify the following: (i) Given a E A, 1r-1(Ua) is an open subset of TAI.
(ii) There is a uniquely determined topology, with respect to which 4a becomes a homeomorphism from ?r-'(U,,,) onto (7r-1MI)) (iii)
determines COD coordinate neighbor-
hood systems of TM. From the definition, it becomes a C°O map from TM onto Al. If we consider the total space TAP of the dual spaces TpAI' of the tangent space TpAI of Al at all point p E At, we see that TAP naturally becomes a 2m-dimensional manifold. This C°° manifold TAP is called the cotangent bundle of Al. Let Sp : Al -+ N be a C°° map from a COO manifold Al into a C' manifold N. Denote by dtpp the derivative of So at each point p E Al. If we set dtp(p, v) = (tp(p), dyop(v)), then we have a map
dtp:TM -'TN
APPENDIX A
170
from the tangent bundle TM of M into the tangent bundle TN of N. This dsp is a C°° map from TM into TN. Let U be an open subset of a C°° manifold M. A CO° map X : U -' TM of U into the tangent bundle TM of M which satisfies, for
anypEU,
iroX(p)=p is called a C°O vector field of M in U. In particular, if U = M, X is called a C°° vector field of M. Since X (p) E it-1(p) = TpM by definition, a vector field X of M in U is nothing but a correspondence
that assigns to each point p E U a tangent vector X (p) of M at xa) the local coordinate that point. Hence, if we denote by system in a local coordinate neighborhood (Ua, 0.) of M, at each point p in U fl Ua, we can express X (p) as m
(A.2)
(p)
X (P) 1=1
(h-). ° p
The m functions' (1 < i < m) are called the components of X with respect to the local coordinate system (xi, ... , x, m). That X is a C°° means that each component of X is a COO function. Given a C°° curve c : (a, b) -. M of M defined in an open interval (a, b), a C°° map X : (a, b) -+ TM from (a, b) into the tangent bundle TM of M with it o X = c is called a C°O vector field along the curve c. Since X (t) E T(t) M for each t E (a, b), a vector field along a curve c means a correspondence that assigns to each point c(t) a tangent vector X(c(t)) of M at the point. For example, the C°O vector field
)t),
c' (t) = dct (
\
t E (a, b),
gives a vector field c: (a, b) - TM along c. Also given a C°O vector field X; M - TM of M, X o c : (a, b) --i TM gives a vector field along c. This is called the restriction of X to c. Given a C'°° vector field X, if there is a C°° curve c : (a, b) -4 M with c4 = X o c, c is called an integral curve of X. Let (Ua, 0,,) be a local coordinate neighborhood containing p = c(t), and let c' = xi o c (1 < i < rn). Then c'(t) is expressed as
(t)
i
=
14
i=1
( ) (t) (
a
(
A.I. FUNDAMENTALS OF MANIFOLDS
171
Comparing it with (A.2), we see that c$(t) satisfies a system of linear ordinary differential equations given by
de
en).
dt
1 < i < rra,
in order for c' = X o c to hold. From the uniqueness and existence properties of the solutions to this system, we see that for a sufficiently small e, there exists a unique integral curve cp : (-E, E) -- U°, satisfying cp (0) = p. We also see, from differentiability of the solutions in
the parameter, that cp(t) is C°° in both t and p. Now fix t E (-F, E), and set Vt(p) = cp(t). Then cot defines a diffeomorphism from Ua onto an open subset pt (UQ) in M, and apt o job = cpt+$ holds in the intersection of the domains of definitions for both sides. In other words, a C°° vector field X generates a local one parameter transformation group of local diffeomorphisms in M. In particular, if M is compact, (pt becomes a diffeomorphism defined in the entire M for any C°° vector field X, and Wt o p. = Opt+6 holds
for all t, s E R. Hence, X defines a one parameter transformation group {Spt}tEa of diffeomorphisms of M.
The tangent bundle and the cotangent bundle represent typical examples of what is called a vector bundle over a manifold. More generally, given a C'° manifold M, if there are a C°° manifold E and a CO° map r : E - M satisfying the following conditions, E is said to be a (C°°) vector bundle over M, and r is called the projection from E onto M. (i) The inverse image Ep -1(p) for each p E M has the structure of r-dimensional vector space over R. (ii) For each point p E M, there exist an open neighborhood U of p and a C°° diffeomorphism 0 : r-1(U) --' U x Wr satj f3 the following. For each point q E U, ci(Eq) = {q} x R'' holds, and furthermore the restriction of 0 to E.
OIEq:Eq-'{q}xRr=W is a linear isomorphism. Ep = 7r-'(p) is called the fiber over p of the vector bundle E. The dimension r of E is called the rank or dimension. Condition (ii) is called the local triviality condition. It is evident that TM and T M* are vector bundles over M. Let E and F be two vector bundles over M. A C°° map 0: E F satisfying the following two conditions is called a homomorphism from E into F. (i) 9(Ep) C Fp for each p E M.
APPENDIX A
1T2
(ii) The restriction of 0 to E is a linear map from E.y, into Fp. In particular, a homomorphism 0: E -- F which is a C°O diffeomorphienm is called an isomorphism If there is an isomorphism from E onto F, E and F are called isomorphic. As is seen readily, if a homomorphism 0 : E - F is injective, the image 0(E) of 0 obtains a nat-
ural vector bundle structure. 0: E - 8(F) is an isomorphism. This @(E) is called a subbundle of F. Given a vector bundle E over M and an open subset U of M, a C°° map s : U - E satisfying 7r o a(p) = p is called a C°° section of E over U. Since s(p) E EE (p E U) from the definition, a C°° section of E over U is nothing but a correspondence that assigns to each p the element s(p) in the fiber Ep. For example, a cross section over U of the tangent bundle TM of M is precisely a Cc* vector field of M. Denote by r(EIU) the set of all C°O sections of M over U. In par-
ticular, r(E) denotes r(EIU) when U = M. Also denote by C°°(U) the set of C°° real valued functions defined in U. COO (U) forms a comilnutative ring under the sum and product of functions. Given
sl:$2 Er(EJU) anda,bER,set (a81 + b62) (P) = a81(p) + bs2(p)
Then asl + bs2 E r(E}U), and r(EJU) becomes a vector space over R under this operation. If we, for s E r(E(U) and f E C°O(U), set (fs)(P) = f (p)e(P), p E U, clearly, f a E r(EIU) and
f(81+82)=f81+fs2, (fl + f2)8 = f is + f2s, (flf2)s = fl(f28)
hold for 8,81,82 E r(EJU) and 1,11,12 E COD(U). Hence, r(EI U) f naps a module over C°°(U). For instance, the set r(TM) of the C00 vector fields over M is a module over C°°(M). Assume that each fiber Ep = yr-I (p) of a vector bundle E over M is equipped with an inner product gp. If, for given al, 82 E r(E), the correspondence
M
ER
gives rise to a C°° function defined in M, (gp) p,E A! is called a fiber metric of E. In particular, a fiber metric g in the tangent bundle TM of M is called a Riemannian metric of Af.
A.2. FUNDAMENTALS OF FUNCTIONAL ANALYSIS
1T3
Let F be a vector bundle over a C°" manifold N and ar : F --+ N denotes its projection. Given a C°° map cp : M -- N, we define a
subset c-1F in the product manifold M x F of M and F by
-1F = {(p, v) I V(p) = 7r(v)}. Let w denote the restriction to p-' F of the standard projection from M x F onto M. We easily see that cp-'F is a vector bundle over M with a fiber t' 1(p) FFlpl over p E M. This tp' 1 F is called the induced vector bundle over M from F by Sp. A C°° section
s of cp-1F over M, by definition, is nothing but a correspondence smoothly assigning to each p E M a point s(p) in the fiber Fw(p) of F. For example, a C°O vector field on M along a C°° curve c : (a, b) -- M
is a C°° section of the vector bundle c 1TM over the open interval (a, b) induced by c from the tangent bundle of M. Let E and F be vector bundles over a C°° manifold M. We can naturally define the dual bundle E* of E, the direct sum EEF of E and F, the tensor product EOF of E and F, the bundle of homomorphisms Hom (E, F) from E into F, etc. Denote by Ep and F. the fibers of E and F over p, respectively. The above vector bundles have the dual space Ep of E, the direct sum EpEFp of Ep and Fp, the tensor product
E® 0 Fp of Ep and Fp, the space of all linear maps Hom (Ep, Fp) from Ep into Fp, etc., as their fibers over p, respectively. The bundle of homomorphisms Hom (E, F) is isomorphic to the tensor product E* 0 F. A fiber metric g is nothing but a symmetric and positive definite CO* section of E" (9 E*.
A.2. Fundamentals of functional analysis (a)
Inverse function theorem. Let V be a vector space over
R, assigning to each z E V a real number R. A function II II : V (1x11, is called a norm in V if it satisfies the following conditions:
IIxII >- o and IIx11= o = x = o;
(i) (ii)
IlAxll = INIIxII,
()
A E R;
Ilx + vll -< Ilxll + IIvII
is given to V, the pair (V, 11 11) of V and II IIr When a norm 11 or simply V, is called a normed space. A normed space V becomes a metric space under the distance p(x, y) = I}x - yll. Hence, we can naturally introduce the notion of convergence in V by defining xn -- x if Hxn - xll = Q. A sequence {x,, } of elements in V is called a Cauchy 11
sequence if Ilxn - Xm 11
0 as rn, n -' ©a. If any Cauchy sequence
APPENDIX A
174
{x,a} converges, i.e., there is an element x E V such that x,, - x, we call V complete. A complete nonmed space V is called a Banach space. A Banach space V is called a Hilbert space, if, to each pair
x, y E V and a real A E R. there corresponds a real number (x, y) satisfying the following conditions: (i)
(x, y) = (y, x);
(ii)
(A1x1 + A2x2, y) = al (X 1, Y) + 1\2(x2. V);
(iii)
IIXI12 = (x, x).
In a Hilbert space, the Schwarz inequality I (x, y) < IIx1111 yll holds.
Let V and W be Banach spaces. A linear operator is a linear map T from a subspace D(T) of V, called the domain of T, into W. A linear operator T is called a continuous operator if Tx. -* Tx as x,, --+ x. T is called a bounded operator if IITxll is bounded for 11x!I < 1. When D(T) = V, being continuous is equivalent to being bounded. In this case, T is called a bounded linear operator. Given a bounded linear operator T, the (operator) norm II T II of T is a real number IITII defined by 11T11 = sup{IITxII Ix E V, IIx1l <- 1}. The set
G(V, W) of all bounded linear operators from V into W becomes a Banach space with this norm. Let U be an open subset of V, and let f : U - W be a map from U into W. Given x0 E U, f is called Fr6chet differentiable at x0 if there exists T E G(V, W) such that lim 11f (x) - f(xo)-T(x-xo)II/hIx-xoll = 0.
sXa
We denote this T by f'(xo) and call it the Frechet derivative or simply the derivative of f at x = xo. f is called Flrechet differentiable in U if f is FWchet differentiable at each point in U. On the other hand, given x0 E U, if there exists li Q(f (xo + ty)
- f (xo))/t = df (xo, y)
for an arbitrary y E V, f is called Gateaux differentiable at xo df (xo, y) is called the Gateaux derivative of f at xo in the direction of y. f is called Gateaux differentiable in U if f is Gateaux differentiable at each point in U. A necessary and sufficient condition for f to be Frechet differentiable is that f is Gateaux differentiable in U, Gateaux derivative df (zo, h) is linear in h, and the correspondence xo i- 4f(xo) is a continuous map from U into £(V, W). Here, we write df (xo, h) = clf (xo)h. Then the Frcchet derivative f'(xo) of f coincides with Of (zo).
A.2. FUNDAMENTALS OF FUNCTIONAL ANALYSIS
175
The following theorem is a generalization of the inverse function theorem in the finite-dimensional vector spaces to Banach spaces. The theorem is very useful. THEOREM A.2 (Inverse function theorem). Let V and W be Banach spaces. Given an open neighborhood U of 0 E V, let f : U - W be a map from U into W satisfying the following conditions: (i) f (O) = 0 and f is Fhchet differentiable in U. (ii) f is C1; namely, for any x E U, the correspondence x'-4 f'(x) is a continuous map from U into £(V, W), where f'(x) is the Fechet
derivative of f at x. (iii) f'(0) : V - W is a homeomorphism; namely, f'(0) is bijective and f'(0) and its inverse map f'(0)-1 are both bounded linear operators.
Then there exists a an open neighborhood V C U of 0 E U such that f is a homeomorphism from V onto an open neighborhood f (V)
of0EW. The reader may consult Lang [101 or Schwarz [22] regarding the inverse function theorem for Banach spaces.
(b) Function spaces and differential operators. Let f) C Rt be a bounded and connected open set, and let 0 < a < I. Given a function u : Si -- R, if
(n)' = =,yEf1 sup
lU(x)
- uWI
Ix - y]
< o©
X3
holds, it is called a Holder continuous function of index a or an aHolder continuous function. A typical example of an a-Holder continuous function is the function u(x) = lxIa defined in a bounded set St containing the origin of R1. Given nonnegative integers c and a, the set of all C' functions u : S3 --+ R whose rc-th partial derivative functions are a-Holder con-
tinuous is denoted by C"+a(Sl) and is called a Holder space. The Holder space C"+' (SZ) becomes a Banach space under the norm ]ulna
=P I ICI
sa
emu) + ICI=
(u)i
Here, /3 = ((31, ... , (1m) denotes a multi-index consisting of m nonnegative integers 's, and 101 = 01 + . + 0,,, denotes its length.
APPENDIX A
176
Furthermore, Dozy represents
D;=
s,
1
Set C"+°(fl) = {u E Ck(11) I Iu.I'.+a < oo} for all open sets W such
that IF' c t. Also set Ca(O) = C°+"(1l) and Ca(Q) = C+`(Sl). Note that we can define the Holder space C"+* (M) over an entire C°O manifold M using partition of unity. Then 0 < a < a' < 1 imply that there is an inclusion C"+12(M) -- C-+*' (M). The Ascoli-Arzela theorem shows that this inclusion is a compact map. Namely, any bounded subset A in C-+" (M) is relatively compact in C +a (M) with respect to the inclusion. In general, a differential operator acting on vector valued functions u : f) -> R9 defined in an open subset of R' is called a linear partial differential operator if it can be expressed as
as(x)D'.
P(x, D) =
(A.3)
101<"
The maximum integer 101 with as # 0 is called the order of P(x, D).
Given ath order P(x, D), a,6(x)Ds
P,1(x, D) IRI=k
is called the principal part of P(x, D). The polynomial in (t1, - - ,m) given by a#(x) o,
P"(X= ) =
d3,
_
d
I$I=k
is called the characteristic polynomial. If the characteristic polynomial P,. (z, C) has no zero except at t = 0 for each x E 12, the partial differential operator P(x, D) is called elliptic. Given an elliptic partial differential operator P(x, D), we call a differential operator of the following form
P(x, D) -- 5i ,
(x, t) E Sl x (0, T),
a parabolic partial differential operator. A typical example of the elliptic differential operator is the Laplace operator A = Di + + D. The heat operator A - 8/8t is a good example of the parabolic differential operator. Let E and F be vector bundles over a C°° manifold M. A linear
map P : r(E) -, r(F) from the space r(E) of C°° sections in E
A.2. FUNDAMENTALS OF FUNCTIONAL ANALYSIS
177
into the space 1'(F) of C° sections in F is called a rc-th order linear partial differential operator if P can be locally expressed by the form of (A.3) as a rcrth order partial differential operator. (c) Heat equation and fundamental solutions. Let (M, g) be an m-dimensional Riemannian manifold, and let i be the Laplace operator on M. A partial differential operator acting on the C2,1 functions u(x, t) defined in M x (0, oo), namely, those which are C2 in X E M and C1 in (0, oo), given by
is called the heat operator in M. The parabolic partial differential equation
Lu=0 is called the heat equation in M. It is called a heat equation due to the following fact. Let 1 C M
be a region with a smooth boundary OP in M. Let v denote the outward unit normal vector of Oil. Assume that St consists of a homogeneous and equidirectional medium. Denote by u(x, t) the temperature at x E S2 at time t. Under the assumption that there is no external application or absorption of heat, the law of preservation of heat implies that the rate of change of the total amount of heat in SZ at t equals the total amount of heat that outflows (or inflows) through the boundary O. In other words,
i (I
u(x, t)dµ9(x)) =
,
8v (w, t)do9(w)
holds. Here, a9 is the naturally induced measure in Oil from pg. This equation, by Green's theorem, equals
j5
OIL
(x, t)d g(x) = J Lu(x, t)dp9(x).
Hence, if this relation holds in eachregion S2, we get Lu = 0 as an equation representing the heat conduction. Given the heat equation Lu = 0, a function H (x, y, t) defined in M x M x (0, co) is called the fundamental solution to the heat equation in M if it satisfies the following conditions. First, it is C2 at x E M and C1 in t E (0, oo). Second, for any bounded continuous function f defined in M, u(x, t) =
fRI
H(x, y, t)f (y)du9(y)
APPENDIX A
178
satisfies Lu = 0 and li o fm H(x, y, t)f (y)dp9(y) = f (s) M
holds.
The significance of the fundamental solution is easier to understand if it is interpreted as follows. H(x, y, t) represents the volume
of heat that transmits from y to x E M after time t when the unit volume of heat is concentrated at the point y E M at time t = 0. If the initial temperature distribution at t = 0 is f (y), the volume of heat transmitted from y to x after t seconds equals H(x, y, t) f (y). Hence, the temperature u(x, t) at x after t seconds is given by u(x, t) =
IM
H(x, y, t)f (y)dp9(y);
consequently, the heat equation Lu = 0 is obtained. When M is the Euclidean space Rtm, it is readily verified that the fundamental solution H(x, y, t) is given by (4irt)-m/2e-Ix-vl'/4t.
H(x, y, t) =
In the case where M is a compact Riemannian manifold, choose a complete orthonormal basis {0s} for the Hilbert space L2(M) con sisting of all L2 functions in M as the eigenfunctions
i4:+a:c: =0, 0=Ao < Al < a2 <... T oo of the Laplace operator A of M. Then the fundamental solution H(x, y, t) is given by 00
H(x, y, t) = E e-`tOa(x)(ris(y)i=O
For given arbitrary x, y E M and t, s > 0, the following are known to be satisfied: (i) H(x, y, t) = H(y, x, t) (symmetric property);
(ii) fM H(x, z, t)H(y, z, s)dµ9(z) = H(x, y, t + s) (semi group property); (iii) fm H(x, y, t)dp9(y) = 1 (conservation property). LEMMA A.3. Let (M, g) be a compact Riemannian manifold and let H(x, y, t) be the fundamental solution of the heat equation in M. Then given x E M and t > 0, we have
H(x, x, 2t) < 1/V + c(M)t-m/2,
A.2. FUNDAMENTALS OF FUNCTIONAL ANALYSIS
179
where C(M) > 0 is a constant dependent on only M, and V represents the volume of M. PROOF. From the above properties (i), (ii) and (iii) regarding the fundamental solution H(x, y, t), we get
H(x, x, 2t) - 1/V =
(A.4)
J (H(x,y, t) -1/V)2dµ9(y)-
Since f,1(H(x,y,t) - 1/V)dµg(y) = 0, from the Sobolev inequality [3], there is a constant ci (111) > 0 dependent solely on M such that m-2
(J(H(xvt) - 1/V) < c1(M)
M
dµ9 (y)
I DYH(x, y, t)I2dµ9(y)
JM
holds. Applying the Holder inequality [31 here, we have m+2
(fMf2g)
mm=2
`
(JMIfIdpg)m
(JM If I
In-2dµ9
Furthermore, noting that the following inequality JM IH(x, y, t) - 1/V I dµ9(y) < fm IH(x, y, t)I dti9(y) + 1 = 2 holds, differentiating (A.4) gives
opt
JM
y, t) - 1/V)2dµ9(y)
= 2 fM(H(x, y, t) - 1/V)
H(x, y, t)dug(y)
= 2 fm (H(x, y, t) - 1/V)D,,H(x, y, t)dµ9(y)
_ -2 fM I DyH(x, y, t)I2dµ9(y) <_ -2c1(M)-1
Im
(H(x, y, t) - I/V)
j
-2c1(M)-1 .2-4/,n J (H(x,y,t) M
' dµ9(y))
m-2 m
m+4
- 1/V)2dµ9(y)
APPENDIX A
180
Integrating the above inequality and noting that H(x, x, 2t) -> oo as t - 0, we get the desired conclusion: -m/2 4c1(M)-1t
f
M
(H(x, y, t)
- I/V)2dµg(y) <_ 4
'm
O
COROLLARY A.4. Let (M, g) be a compact Riemannian manifold and let H(x, y, t) denote the fundamental solution to the heat equation
on M. Then for any x, y E M and t > 0, we have H(x, y, t) < c(M, t).
Here, c(M, t) > 0 is a constant dependent on M and t alone. PRooF. By Lemma A.3 and Schwarz's inequality, we get
H(x, y, 2t) -1/V =
J(H(xi z, t) -1/V)(H(y, z, t) -1/V)dpg(z) 1/2
!f -1/V)2dpg(z) JM(H(x,z,t) 1/2
/r'
JM(H(y, z, t) -1/V) 2 dµg(z)
x
<
4(4c1(M)-1t/m)-m/2.
0 (d) Differentiability of solutions. It is known that the solutions to elliptic and parabolic partial differential equations, in general, reach the maximum possible differentiability under the given data and coefficients of the equations. Let f) C Rm be a bounded and connected open set, and let P be a linear elliptic partial differential operator given by
P=
a=1(x)
i, j,1
8x axj
+ E b (x) z=1
ax=
+ d(x).
The the following holds.
THEOREM A.5. (1) Given 0 < a < 1, assume that a'3, bi, d, f E Ca (SZ). Then U E C2+a (1) if u E C2 (12) satisfies a linear elliptic partial differential equation Pu(x) = f (x).
A.2. FUNDAMENTALS OF FUNCTIONAL ANALYSIS
181
(2) Furthermore, if aU, b°, d, f E C"+a (!1) forgiven K > 1, then C"+2+a. In particular, if a:,, b, d, f E Coo (fl), the solution u to (1) is then u E COO (M).
We may state the case of linear parabolic partial differential equa-
tions as follows. Given T > 0, set Q = SZ x (0, T). For a function set a =u )x
{
sup
Iu(x, t) - u(x', t) I I t - t' I
(x,t),(i ,t)EQ x,6x'
sup (u} = (x,t),(x,t')EQ
(u (z, t) - u(x, t`)I
It - t/Ia
t#t' +a'1+a/2) are defined as (4.16) in ChapThe norms IUIQ'a/2) and IUI( Q ter 4. If we define function spaces Ca,a/2(Q), C2+a,1+a/2(Q) C2+a,l+a/2(Q) and Ca'a/2(Q),
as given in §A.2(b) with respect to these norms, the following holds.
THEOREM A.6. (1) Given 0 < a < 1, assume as', b', d E Ca(t ) C2+a,1+a/2 holds, if u E C2"l satisfies and f E Ca'e/2(Q). Then u E the following linear parabolic partial differential equations
(
-
) u(x, t) = .f (x, t).
(2) Let p, q be nonnegativve integers. Given 0, n with I!01 < p,
q, assume Dxaii, Djbt, Dgd E Ca(fl) and DpDt f E Ca,a/2(Q). Then the solutions u to (1) satisfy DgDt u E Ca'0/2 for any 0, n with 101 + 2K < p + 2, K < q + 1. In particular, ail, b, d E C°°(0) and f E C°°(Q) imply that u E C°°(Q). IQI+2K C K, K
Regarding the results mentioned above, see Murata-Kurata [29] and Gilbarg Trudinger [3].
(e)
Schauder estimate. We give here a quick review on the
Schauder estimate of solutions to linear elliptic and parabolic partial differential equations. The estimate was used in §4.3.
Given r>0,setB(0,r)={xERtI IxI
1 < i, j < m,
APPENDIX A
182
and that
m
a`1(x)ttt3 <
AItI2 <
holds for some constants 0 < A < A < ooA and for any x E B(O, r) and t E 1R1. Then Theorem A.7 holds regarding the linear elliptic partial differential operator 2
P=
a=' (x)
is=1
8xt0x, + i=I
b` (x) axs + d(x)
and the linear parabolic differential operator
L=P --
.
THEOREM A.7. (1) If f E Ca(B(O,r)) and 2a E C2(B(O,r)) satisfy
Pu(x) = f (x),
then u E C2+0 (B(0, r)) and IUIC2+a(B(O,r/2)) :5 C(I f I L-(B(O,r)) + IuIL°°(B(0,r))),
ItIC2+° (B(0,r/2)) !5 C(I f I C° (B(O,r)) + I UI L-(B(O,r)) )
hold. Here, C is a constant determined only by m, a, A/A, IaE"IC'(B(O,r)), Ib`Ic.(B(o,r)), IdfcQ(B(0,r)) (2) Let 0:5 t < T. If f t) E Ca(B(0, r)) and u(-, t) E C2(B(0, r))
satisfy then
Lu(x, t) = f (x, t), t) E C2+a(B(O, r)) and
Iu(',t) I C° (B(O,r/2) ) !5
C( SUP IK, t)IL-(B(O,r)) + SUP Iu(', t)IL-(B(0,r))) I tE[0,T1
tE[O,T1
Iu(',t)IC2+a(B(O,r/2)) +
cat
(.'t)IC
(B(O,r/a))
< C( SUP I f(, t)ILO(B(O,r)) + SUP IU(', t)I L°0(B(O.r))) tE[O,TJ
hold.
tE (0,T]
Here, C is a constant determined only by m, a, A/a, Ia''IC-(B(O,r)),
Ib'ICa(B(0,r)), Idlca(B(O,r))
Regarding the results mentioned above, see Murata-Kurata [29] and Gilbarg-Trudinger [3].
Prospects for Contemporary Mathematics The objective of this book is to guide the reader to the threshold of general geometric variational problems, through an in-depth discussion of variational problems that arise in the study of the length and energy of curves and the energy of maps. The reader obtains basic knowledge of Riemannian geometry in the the process of learning properties of geodesics and harmonic maps as solutions to the variational problems. In this aspect, this book is also designed to play a role as an introduction to differential geometry. The comparison theorem of Jacobi fields and the Morse index theorem were originally planned to be taken up in the book. They,
however, are not included due to space limitations. The reader is advised to consult Milnor [12] for these topics. As stated in the Preface, geometric variational problems define an area of mathematics where variational problems in the study of the geometry of manifolds are investigated from the viewpoint of global analysis. Geodesics and harmonic maps are typical of subjects of investigation in this area. As discussed in this book, the solutions to variational problems that take place in manifolds, in general, are characterized as solutions to nonlinear partial differential equations. Consequently, one may
also state that geometric variational problems represent the study of geometric phenomena described by nonlinear partial differential equations in manifolds. It is relatively recent, after the 1950's, that the study of geometric
variational problems as above became active. Among them are the Yamabe problem regarding the deformation of the scalar curvature of Riemannian manifolds and the Calabi conjecture regarding existence of Einstein-Kiihler metrics. They were formulated, respectively, as
problems in nonlinear elliptic partial differential equations and the complex Monge-Ampere equation, and have since become constant sources of stimulation and motivation in the area. On these topics, 183
184
PROSPECTS FOR CONTEMPORARY MATHEMATICS
the reader is urged to read the series, Reports on Global Analysis, Proceedings of the Workshop "Surveys in Geometry" organized by T. Ochiai and M. Obata (see Books, 7-9). Among this series, the lecture notes by Kazdan [7] are most suitable as an introduction to this area. The Yamabe problem was solved by R. Schoen in 1984, and the Calabi conjecture by T. Aubin and S: T. Yau in 1977. In addition to the Yamabe problem and the Calabi conjecture, there are a number of major problems which currently are the targets of very vigorous research activities. They include the Plateau problem on the existence of minimal surfaces and Hamilton's problem regarding deformations of Ricci curvature of manifolds. In addition to the more traditional research on the structures and properties of manifolds, differential geometry has recently expanded its research activities into studies of the set consisting of all spaces. For example, the theory of connections on fiber bundles has found a natural tie with gauge theory in mathematical physics. As a consequence, geometry of gauge fields has been and is being very actively investigated from the viewpoint of geometric variational problems. In the case of harmonic maps, the existence theorem of Eells and Sampson was extended in 1975 to compact manifolds with boundary by Hamilton [4]. He showed existence of harmonic maps under the same curvature condition as the Eells-Sampson theorem by applying the heat flow method to the Dirichlet and Neumann boundary value problems and by using analysis in the Sobolev space Wk,p. Through the study of the existence problem of closed geodesics, Palais [17] and Smale [21] developed Morse theory of infinite-dimensional manifolds in the 1960s. However, this theory is not applicable to the existence problem of harmonic maps due to the fact that the energy functional, in general, does not satisfy Palais-Smale condition (C). Eells and Sampson used the heat flow method in order to circumvent this difficulty. Uhlenbeck [20] reformulated the definition of energy functional in such a way that Palais-Smale condition (C) is satisfied, and investigated the existence problem of harmonic maps from the viewpoint of Morse theory in infinite-dimensional manifolds. In collaboration with Sacks, she proved a representation theorem for the two-dimensional homotopy group ir2(R7) by harmonic spheres. They also discussed there a phenomenon now known as bubbling up. It is a cohesive phenomenon of energy that is characteristic in the existence problem of two-dimensional harmonic maps. The phenomenon
PR.OSPE.AL" TS FOR CONTEMPORARY MATHEMATICS
168
of cohesion of energy also occurs in geometric variational problems regarding gauge fields. The reader is advised to see the original papers
about this important phenomenon. Chapter 4 ends with a discussion of the strong rigidity theorem of Kahler manifolds. In 1960, Calabi and Vesentini, from the viewpoint of the deformation theory of complex structures, showed that any compact quotient manifolds of irreducible symmetric bounded domains of dimension greater than two poi a local rig ty prop. erty. In other words, complex structures in these manifolds allow no nontrivial infinitesimal deformations. The strong rigidity theorem of Siu [23} corresponds to the strong rigidity theorem of locally syunmetric Riemannian manifolds proved by Mostow in 1970, and shows that the theory of harmonic maps is fundamentally useful in complex geometry. Along the same line in use of harmonic maps, Nishiicawa and Shiga [161 took up an equivalence problem among the bounded domains in Kahler manifolds of negative curvature. Since the publication of the paper by Eells and Sampson in 1964, there have been numerous results on the existence problem of harmonic maps between compact Riemannian manifolds. The existence problem between noncompact manifolds has not yet seen systematic research activities. It may be considered as one of the most important research topics for the present and future in the field of geometric variational problems. Complete, simply connected Riemannian manifolds of nonpositive curvature are called Hadamard manifolds. One can define an ideal boundary for a Hadamard manifold by considering that all mutually asymptotic geodesics give rise to the same point of infinity. The set of all points of infinity is defined to be the ideal boundary. The boundary problem at infinity for harmonic maps between Hadamard manifolds, namely, whether or not a map between two ideal boundaries is extendible as a harmonic map between Hadamard manifolds, is an intriguing problem. This problem was solved over the period from 1990 to 1992 by P. Li and L. F. Tam, M. Gromov and K. Akutagawa for the spaces of negative constant curvature (real hyperbolic space forms). In this case, uniqueness of solutions does not generally hold; therefore, this is quite different from the Dirichlet boundary value problem for compact Riemannian manifolds with boundary. Later in 1993, H. Donnelly studied the same problem for noncompact symmetric spaces of rank I including complex hyperbolic space forms. Recently, Nishilmwa and
188
PROSPECTS FOR. CONTEMPORARY MATHEMATICS
K. Uerno extended the above results to harmonic maps between more general homogeneous spaces of negative curvature. Through these research efforts, it has become more apparent that harmonic maps play an essential role in relating geometry and analysis of homogeneous spacm (domains) of negative curvature to those in the ideal boundaries (points of infinity). As described above, the ideal boundary value problem of harmonic maps is an extremely interesting problem, in which some mutual relationships between geometry and
analysis are being revealed. The research on this topic is still in its infancy, and much is expected to be done in the future.
Solutions to Exercise Problems Chapter 1 1.1 Easily follows from the definition. 1.2 Let {Vol be an open cover of M such that each V,l is a local
coordinate neighborhood of M and V3 is compact. Since M with the second axiom of countability is paracompact, there are a locally finite refinement {Ua} and a partition of unity {f} subordinate to {U,,}. Since Ua can be identified with an open subset of R' through the local coordinate system, it has a Riemannian metric gQ. At each point x E M, define an inner product gr in TxM by f,(x)gz (v, w),
gs(v, w) _
v, w E T1M.
Here, F,` represents the sum over a with x E U. Then g = {gam} is the desired Riemannian metric. There is another way to prove it. Note the theorem of Whitney that states that any m-dimensional C° manifold can be imbedded in 2m+1-dimensional Euclidean space. Then follow Example 1.2. 1.3 Do in the same manner as deriving (1.14) from (1.13). 1.4 (1) The transformation formula of the natural frame fields nt UXr m
a
a
a.j = L axj Ox axq
a
a
,
8xk
axk axr
and (1.23) yield a_
°
in
m
axq
E E -W
axk = p=1
q=1
o2 82xp
axgax
in
p axr
E rgra-k + q=1
a aaxp.
Substitute the following transformation formula in the above equation
a axp
'nax3a i=1
and compare it with (1.22). 187
axpaP
SOLUTIONS TO EXERCISE PROBLEMS
188
(2) Define V Y by (1.23) in each coordinate neighborhood. Then we can patch them together using the transformation formulas in (1); consequently, a linear connection V is obtained in M. 1.5 (1) For the second equation, we replace X, Y, Z in [X, [Y, Z]] [X, YZ - ZY] = XYZ - XZY - YZX + ZYX in the order X -+
Y -,Z - X.
(2) From the first equation in (1), we we that the Lie derivative LXY satisfies the following rules:
(i) Lx(Y + Z) = LxY +LxZ, (ii) Lx(fY) = X(f)Y+ fLxY. (iii) Lx+yZ = LxZ + LyZ, (iv) L fxY = f LxY - (Y f )X. Among them, (iv) is different from the rule V fxY = f VXY for covariant differentiation. This implies that the value LXY(x) of the Lie derivative is essentially determined by the behavior of both X and Y around x, unlike the covariant derivative, where the value VxY is determined by the behavior of Y around x and the value X (x) of X at x. In other words, Lie differentiation is not an operator determined by the direction X (x) of differentiation alone. Furthermore, the second equation in (1) yields a relation Lx Ly LyLX = L[x,y), but the relation VX Vy - Vy Vx = V 1x,yj does not generally hold for covariant differentiation. This last relation holds only when the Riemannian manifolds are flat (curvature tensor R = 0) (see §2.1).
1.6 Let lei, ... , e,n} be a base for T1M. Set EE(t) = Pte (1 < i < m). Since {Ei(t), ... , E1 (t)} is a base for T ,(t) .M for each t, we can write as m
Yati = EYi(t)Es(t),
Y ` E C°°([0.1]).
i-1 Then, from the properties of linear connections and Ei being parallel, we get
V'Y = a.1
dY9 (0)e. dt
d
= 3i E Y`(t)e;(t) i=1
)I
= dtPt 'Y(t) t=o
t=o
1.7 If we assume that there is a vector field fi as stated in the problem, we see that its integral curves Sp(t) = (c(t), c'(t)) are solutions to (1.30) and, hence, unique. As for its existence, we may locally define the vector field -iP by (1.30). From the uniqueness, 4 is defined globally.
SOLUTIONS TO EXERCISE PROBLEMS
189
1.8 (1) Necessity is clear from the definition. Sufficiency follows from the uniqueness of geodesics regarding the initial condition. (2) follows readily from (1). 1.9 As is well known, a connected C°° manifold is arc-wise connected. Hence, there is a continuous curve c : [a, b] --> M joining p and q. Cover the compact set c([a, b]) C M by a finite number of coordinate neighborhoods U,,, then replace c by a COO curve in each Ua-
1.10 Given P E M, for sufficiently small E, expp BE(0) = {q E M I d(p, q) < E} holds, where BE(0) = {v E TpM I 'vi < E}. expp is a local homeomorphism and expp BE (0)'s form a base for the local neighborhood system at p. Consequently, the topology defined by d coincides with the topology of the manifold.
Chapter 2 2.1 This follows readily from the definition. 2.2 A type (1, s) tensor field
T=
a
T,._j,(d ") 0 ... g (dr -)x 0
a
on TM corresponds to an s linear map \\
)')
x
`
),z
from T¢M X ... X T2M into T=M. 2.3 This follows because, in general, V(X, f Y) = (X f )Y +
foxY 0 f4(X, Y).
2.4 Since ]8/8x',818z3] = 0, from the definition, we have
a
,
s
a
aa
axe
=
.
t l
2.5
_
0
a -V Sxk
V
a-
9rik+E(`a`'jk-ayrr'i )
{
T
-
VF
It can be done readily by noting the properties of R in
Proposition 2.11 and that the orthonormal basis {v', w'} for or is given by the Gram-Schmidt orthonormalization as
= v lv)
w' = w - 9x(V, w)v'
1w - 9.:(v, w)v'I.
SOLUTIONS TO EXERCISE PROBLEMS
190
2.6 It suffices to be able to determine R(u, v, w, t), du, v, w, t E TIM, given R(u, v, v, u), Vu, v E TIM. First of all, from the relation R(u + t, v, v, u + t) = R(u,v,v,u) + R(t, v, v, t) + 2R(u, v, v, t) follows that R(u, v, v, t) is also determined. From this and the relation R(u, v + w, v + w, t) = R(u, v, v, t) + R(u, w, w, t) + R(u, v, w, t) + R(u, w, v, t), we have
R(u, v, to, t) + R(u, to, v, t) = (*),
where (*) is the sum of known terms. By applying the first Bianchi identity to the second term, we get 2R(u, v, w, t) - R(w, v, u, t) = (*). By exchanging u and to here, we also get 2R(w, v, u, t) - R(u, v, w. t) = (*). These two equations imply that R(u, v, w, t) _ (*); hence, R(u, v, w, t) is determined.
2.7 Consider a C°° map u : 0 -+ M with u(0, s) = u(0,0) (-E < s < E), where 0 is an open subset of R2 given by
O={(t,s)I -E0). Given v E TM, define a C°° vector field V along u to satisfy that V (0, s) = v and V (t, s) is the parallel transport of v along each C°° curve t +-- u(t, s) for t # 0. Then by Lemma 2.15, we get
D DV=O= as at
D
V + R (au u D atas as° t
V.
Since the parallel transport does not depend on the choice of the curves from the assumption, V(1, s) equals the parallel transport of V(1, 0) along the C°° curve s +- u(1, s); hence, sequently, from the above equation, we get R
(
(1, 0),
(1, O)
D V (l' 0) = 0. Conus
V(1,0) = 0.
Since u and v are arbitrarily chosen, we get the desired conclusion.
2.8 Set'VxY = dV-1(V'd,(x)dcp(Y)). Then 'V defines a linear connection in M and satisfies conditions (i) and (ii) in Theorem 1.12. From the uniqueness of the Levi-Civita connections, we see that 'V = V. This implies (1). (2) readily follows from (1).
SOLUTIONS TO EXERCISE PROBLEMS
2.9
191
For (1), define g = u+*g. (2) readily follows from (2) of
Problem 2.8 above..
2.10 If AI is orientable, Theorem 2.26 implies that Al is sinlply connected. If Al is not orientable, apply Theorem 2.26 to the orientable double covering space M.
Chapter 3 3.1 Since the support of f is compact. the right. hand side is a finite sum. Let {Vj, +0dEB be another coordinate neighborhood system and denote by {O3}3EB a partition of unity. Then at p E U, n V,3, we have det(9k )(p) _ I det J(00 o V', 1)(+l',s(p))I det(9,)(P)
The change of variable formula for the integral in R"' then readily implies
EL t3
_
o u'3 Idx3
0 "Of (
E
f
... d J
(ci8nf Vdet(Yii) o ii
... d
13
(aapaf det(gj1) -1I det .1(p,, o t'' 1)Id.r 3 ... dx"'
o VJ
=E a.ti
_
(ydp f Jdet(g1)) oo 'd4 ...dx"' J46,. (U,+ntr., )
J
f
1 ... d.r It (pt Jdet(g ,)) o 0 1dr0 n
.
That `!9 is positive definite. namely. /!q(f) 0 for nonnegative functions f, is clear from the definition. In order to see 11g being a Radon measure on Al, namely, being a bounded linear functional in Co (AI ). we must show the following. Given an arbitrary continuous function f whose support is a compact subset K. there is a constant CK such that.
Ii1 (f)I <_ cti Sul) If(P)I PE AI
holds. This is also clear from the definition.
SOLUTIONS TO EXERCISE PROBLEMS
192
3.2 Choose a coordinate neighborhood U of x E M and a coordinate neighborhood V of u(x) E N so that u(U) C V holds. Denote by (ya) a local coordinate system in V. Since { (8/eya) o u} (1 < a _< n)
forms a basis for the fiber T ()N over x in u-1TN, a section rJ of u-1T N over U is expressed as tl = E. rf (8/8ya) o u. Assume that a linear connection IV in the problem exists. Then, from the properties of a connection, we must have
V
8ya
{v(i) 8ya o u +
r)
and, hence, is unique.
As for the existence, we may define 'V by the above equation. It is readily verified that 'V, independent of the choice of coordinate systems, defines a linear connection. 3.3 Each can be verified directly from the definitions. For (2), see the solution to Problem 3.4 below. 3.4 Let {ipt } be the one-parameter group of transformations
generated by X. Denote by (xi) a local coordinate system in the coordinate neighborhood (U, 0). The measure determined by the pullback Spt g of g under yet is given as k dµ,pt*9
=
det
kj
gki 0 sat ate;
sat
axJ
dx1... dx'
.
Hence, we have d I t=0 dµ4g = divX dug. dt
In fact, when (ai?) (t) is a differentiable nonsingular matrix with respect to t, the derivative of the determinant is given by
d det(ai2(t)) = det(aij(t)) 1:
akt(t)dtakt(t),
k,1
where (ai3(t)) denotes the inverse matrix of (ai3(t)). Furthermore,
since d/dtlt=o (8¢i /8xi) = 0Xk/8c= and aho/exa = aik
for
SOLUTIONS TO EXERCISE PROBLEMS
193
X = Et X to/axt, we have
L.H.S. =
X det(gt1) +
det (9z
det
2
1
{',fagk'\ 2
k, 1
)
d
Jdet(gj) dx1
VAX'
det
+[ dX
8x4
i
I
dt t=o
o
dx1... dxm
d.r'
t
L 8.x'
det (gti) dxl ... dxm
dxm
.
= R.H.S. Since each Spt is a diffeomorphism, we see readily from definition that
f
dµg.
At
t
Consequently, by adding the local forms of the above result using the partition of unity, we get
0=
I'=0
dt
J dµ4; g= f div X dµg I
.
AI
3.5 (1) follows readily from the definition. To see that VT becomes an (r, s + 1) tensor, it, for instance, suffices to note that VT becomes C°°-linear with respect to C°` modules C(TAI), I'(TAI*); namely, the following holds
VT(fX, f1X,,... ,f.X.,hjwj.... ,hr'Wr) = f ft ... f8h, ... hrVT(X, X1. .. . X8, u71, ... , Wr).
(2) follows from a similar computation to Lemma 3.4, noting (3.16), (3.26).
3.6 From Problem 3.5 above, VR is defined by
VR(X,Y,Z,V,w) = X R(Y,Z,V.w)-R(VxY,Z.V,w) -R(Y,VxZ,V,w)-R(Y.Z,VxV.w)
-R(Y,Z,V.V w). Noting that R(Y, Z, V, w) = w(R(Y, Z)V) and V.V w(Y) = X w(Y) w(VXY), we get
VR(X, Y, Z, V, w) = w(Vx R(Y, Z)V - R(V Y. Z)V - R(Y, VxZ)V - R(Y, Z)VxV).
SOLUTIONS TO EXERCISE PROBLEMS
194
Hence, it suffices to see that the symmetric sum over X, Y, Z of the expression
Vx R(Y, Z)V - R(VxY, Z)V - R(Y, VxZ)V - R(Y, Z)VxV = [Vx, R(Y, Z)]V - R(VxY, Z)V - R(Y, VxZ)V equals 0. Here, noting that V XY - V y X = [X, Y] and R(X, Y) Z = [V x, V y] - V [x,y], we get
[Vx, R(Y, Z)]V - R(VxY, Z)V - R(Y,VXZ)V + [Vy, R(Z, X)]V - R(VyZ, X)V - R(Z, VyX)V + [Vz, R(X, Y)]V - R(VzX, Y)V - R(X, VzY) V = [V x, R(Y, Z)]V + [Vy, R(Z, X )]V + [Vz, R(X, Y)]V R([X, Y], Z)V - R([Y, Z], X )V - R([Z, X], Y)V
-
= ([Vx, [Vy, Vzl] + [Vy, [Vz, VxJJ + [Vz, [Vx, Vyll)V + (V I[x,yi,z] + V ((y,z],x] + V [iz,xl,yj )V.
Hence, we get the desired conclusion from the Jacobi identity for the operator and the vector fields. 3.7 (1), (2) follow readily from the definitions. (3) Apply the definition (1.26) of the Levi-Civita connection together with (1), (2). 3.8 One may verify that 0 satisfies the equation for harmonic maps by a direct computation. One can also show that 0 is a harmonic map from Proposition 3.17 by noting that S3 -> S2 is a Riemannian
submersion and that 0-1(y) is a geodesic (a great circle) of S3 for each y r S2. 3.9
Since cp' = e2Pg,
!
Egki (P) 8i
= e2P9:j
kj
and
0k
1
gj(W) 0xi (P axi = e-2P9kj(W) :J
SOLUTIONS TO EXERCISE PROBLEMS
195
hold for the components of g with respect to a local coordinate system. From this we get
e(u o cp) =
8ua 8ipk au.3 8p1 (u o'P) -5X--k 8xi 8x1 8xj
EE'j 9
i, j,k,l a, O
=
cp*(dug) _
e-2p
!:kJ Q"3
o V)
ekal
jdet
8i
k,1
&
3 = e-2pe(u)(p),
dxldx2 = e2Adp9.
er3E(a-a')/2IWI(a,a/2) Q
From this it follows that follows ((w) ,°'2) <
C3e(a-a')/zjwjQ'si2)
through a simple computation for X, Y E r(TAI). Consequently, from
the definition of tension field and Lemma 3.3, we get that. r(cp) = (2 - m) grad p, where m is the dimension of Al. In particular, when m = 2, cp is a harmonic map. 3.10 Let h = u*h denote the induced metric from h by u and denote by hij the components of h. Let h denote the components of the type (1,1) tensor field corresponding to the type (0,2) tensor field h of Al under the ispomorphism TAI * ® TAI * . Then we get h = Ek hjkgki. On the other hand, we see that
/det(ii) <
2 trace (h )
holds for the (2,2) matrix (h'), and that the equality holds when there is a positive number A such that hl = A6 : namely, only when h? = Agi j . Exercise 3.10 follows readily from this. On the other hand, noting that the induced connection '0 on the vector bundle tp-ITAI is compatible with the fiber metric yp* together with cp*g = e 2Pg and (1.26), we can readily verify that
'VxdV(Y) =
(Xp)Y + (Yp) X - 9(X, Y) grad p
through a simple computation for X, Y E t (TAI ). Consequently, we get that r(ip) = (2 - m) grad p, where in is the dimension of AI. In particular, when m = 2, p is a harmonic map.
SOLUTIONS TO EXERCISE PROBLEMS
196
Chapter 4 4.1 (1) Given T E I'(TM* ®u-ITN) and X, Y, Z E r(TM), we have
(VVT)(X, Y, Z) = (VxVT)(Y, Z) = Vx(VT(Y, Z)) - VT(VxY, Z) - VT(Y, VXZ)
= Vx((VyT)(Z)) - (VvxyT)(Z) - (VyT)(VxZ) = (Vx(VyT))(Z) - (Vv,ryT)(Z) (2) From the definition of the connection in TM* ®u-ITN, we have
'Vy(T(Z)) = (Vy)(Z) +T(VYZ), 'Vx'Vy(T(Z)) =(VxVyT)(Z) + (VyT)(VxZ), + (VxT)(VYZ) + (VxT)(VyZ) Similarly, computing -'V' Vx (T(Z)) and -'V (x,yj (T(Z)) and adding them together, we get
R V (X,Y)(T(Z)) = (RV (X, Y)T)(Z) +T(RM(X,Y)Z) (3) From (2) and the definition of the induced connection 'V, we get
a a T) a ax'' axi a2k a s
Rv
=R'v
_
ax1 ° axi
R a
T
Na
axk
Ou0 8u
a,7 ,a
p,y,S
a s a az= aW 57k
T (RM
Tax' axj
R11li t3kTia
a
aya
0 U.
!
On the other hand, since we have
VVT
ate=
aZ
,
aaxk
ou,
- VEV?T'°`k a
the conclusion follows from the definition of Rv. 4.2 We may use an idea similar the one used in the proof of the Weitzenbock formula for e(ut). First, from the definition of ro(ut), we have
{ )_
at
p
afi
haQ(ut)Vt
at &
SOLUTIONS TO EXERCISE PROBLEMS
197
On the other hand, noting that VI att = VtV1u' . we get EE9kthaj3(ut)VkVtV1 W 0.13
3
2
N + IC at
I
*
From the Ricci identity regarding the connection V in T (Alx(0, T)) * 0
u-1TN, we then get
VkVtVlut - VtVkVkul RA1x(0.T)r auf + kt!jX
ER
x act dut out ,1Eaxkataa1
Noting that Al x (0, T) is a product of Riemannian manifolds, we can readily verify that Riutx(0.T)ktr = 0 from the definition of the curvature tensor. Consequently, we get the desired result by substituting
the Ricci identity in the above equation and by noting that u is a solution to the equation for harmonic maps. 4.3 If we consider the derivative d exp(p,o) of the map exp : U N at (p, 0) E TAI -L following the line along the proofs of Theorems 1.24 and 1.25, we see that its matrix representation with respect to the canonical coordinate system gives rise to
1 I0 0I Hence, noting that AI is compact, the existence of the desired c follows from the inverse function theorem.
4.4 Noting Lemma 3.3 and the definition of the induced connection, we get, for X. Y E F(TAII ),
Vd(f2 o fl)(X,Y) = Vx(df2 odfl(Y)) - d(f2 o df1)(VxY) (Vdf1(x)df2)(df1(Y)) +df2(Vx(dfl(Y))) - df2 o dfl(VxY) = Vdf2(df1(X),df1(Y)) +df2(Vdf1(X.Y)).
From this follows the first equation. The second equation readily follows from the first equation.
4.5 If Vdu = 0, the formula for the second fundamental form of composition maps implies that d(u o c) _ du dt
-
de\
d ] + Vdu
de do dt dt
_ =0
SOLUTIONS TO EXERCISE PROBLEMS
198
for any geodesic c : I -> M. From this follows (i) (ii). Conversely, if (ii) holds, Vdu(dc/dt, do/dt) = 0 holds for the tangent vector do/dt of the geodesic. Vdu = 0 holds, since any tangent vector v E TM at each point x E M can be given in the form of do/dt for some geodesic. C2+a,1+a/2(Q, RQ) 4.6 Set Q= Mx [0, TJ. For example, regarding being a Banach space, a proof goes as follows. Let { Uk } be a Cauchy C2+a,1+a/2(Q, Re). sequence in Since (Uk) is a uniformly bounded and equicontinuous sequence in C2,1 (Q, Re), the Ascoli-Arzela theo-
rem implies that there is a subsequence {uk'} of {uk} such that it converges to some u in C2,1 (Q, Re). Then it suffices to show that u C2+a,1+a/2(Q,RQ) is an element of and that {uk} converges to u in C2+a,1+a/2(Q, Re). These can be directly verified from the definition of the norm IuI(+a,1+a/2) ((w)xa"/2)
4.7
Since I(wI( ''a'/2) = I(WIQ + ((w)(a') + from the definition of the norm, it suffices to estimate individually I(wIq, ((w)za') and ((w)xa'/2) From the assumption, w E C2"1(Q,RQ) and w(x, 0) = 0, we we that I(WIQ < Similarly, we see C1Ea/2I(wI(,a/2).
that ((w)(a') <
C2E(a-a')/21(wI(,,a/2). (For
example, we may treat the two cases where d(x, x') > E1/2 and d(x, x') < E1/2.) On the other hand, regarding ((w)Xa'/2), for 0 < t < t' < 2E, we have 1((t)w(x, t)
- ((t')w(x, t') I
I ((t)w(x, t) - w(x, t')I + I (((t) - ((t'))(w(x, t) - w(x, 0))1 I((t)IIwI(,a/2)It
- t'Ia/2 +2E-1It -
t'IiwI(a,a/2)It'Ia/2
Dividing both sides of the above inequalities by It
I((t)w(x, t)
- ((t')w(x, t')IIt -
- t'/2, we get
t'1-a/2
< IwI( ,a/2)(2E)(a-a')/2 + 2e-1(2E)1_Q'/21w1 (a,a/2)(2e)a/2It'Ia/2 :5 C3E(a-a')/2IWI( ,a/2).
From this follows ((w)Xa'/2) <
C3E(a-a')/21
wI ( 4.8 Given a constant C and an e > 0, set
fi=e-(C+1)tu,
.a/2).
Q=Mx[0,T-e].
SOLUTIONS TO EXERCISE PROBLEMS
199
From the definition, the signs of u and it are the same. They satisfy, in M x (O.T), the inequality
atu 0. Since 0) < 0 from the assumption of u, we must have t° > 0. We readily see that this contradicts the above
inequality regarding u, applying a similar argument to the proof of Lemma 4.11. Since E is arbitrary, we get the desired conclusion.
4.9 Consider M to be a Riemannian manifold with any Riemannian metric. Since M is compact, we can choose a finite cover of M consisting of geodesic spheres Br(x1), ... , Br(xk) such that f (B3r(xi)) is contained in a coordinate neighborhood of N for each i L... , k. In B3r(x1) and for sufficiently small f, define
MX) _ J0 J B2r(Xi) (47rt)
-m/2 exp(-d(x, y)2/4t)f (y)dµ9(y)dt.
Choose a C°° function Vi : M --> R such that
0 < Bpi <_ 1, gi(x) = 1 (x E Br(xi)), pi(x) = 0 (x V B3r/2(Xi)) Set 11 = cpl f 1 + (1- Cpl) f. Then 11 becomes a function defined in the
entire M. We inductively define fi by
fi(x) =
(4irt)-m/2
fO'
exp(-d(x, y)2/4t)fi-1(y)dµs(y)dt,
B2r(zi)
and set fi = cpifi + (1 - cpi)-f i_1.7k gives the desired C°° map f. It is readily verified that f is free-homotopic to f as s - 0. 4.10 An element a in wk(M) is nothing but the homotopy class
of a continuous map f : Sk - M from the k-dimensional sphere Sk into M. Since K%(< 0, f is free-homotopic to a harmonic map u : Sk - M by Corollary 4.18. Proposition 4.24, then, implies that u is a constant map for k > 2; hence, the desired conclusion is obtained.
Bibliography
[2]
J. Eells, Jr. and J. H. Sampson, Harmonic mappings of Riemannian manifolds, Amer. J. Math., 86 (1964), 109-160. J. Eells, Jr. and J. C. Wood, Restrictions of harmonic amps, Topology,
[3]
15(1976), 263-266. D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of sec-
[1]
[4] [5]
ond order, Grundlehren der mathematischen Wissenschaften 224, 2nd edition, Springer-Verlag, 1983. R. S. Hamilton, Harmonic maps of manifolds with boundary, Lecture Notes in Mathematics 471, Springer-Verlag, 1975. P. Hartman, On hamotopic harmonic maps, Canad. J. Math.. 19(1967). 673687.
H. Hopf and W. Rinow, Ueber den Begriff der vollstindigen differentialgeometrischen Flichen, Comm. Math. Helv., 3(1931), 209-225. [7] J. Kazdan, Some applications of partial differential equations to problems in geometry, Reports on Global Analysis. VI, 1-130, Seminar Kankokai, 1984. [8] W. Klingenberg, Lectures on closed geodesics, Grundlehlen der mathematischen Wissenschaften 230, Springer-Verlag, 1978. [9] O. A. Ladytenskaya, V. A. Solonnikov and N. H. Ural'ceva, Linear and quasilinear equations of parabolic type. Translations of Mathematical Monographs 23, Amer. Math.Soc., 1968. [10] S. Lang, Introduction to differentiable manifolds, Interscience. 1962. [11] T. Levi-Civita, Notione di parallelisimo in una varieth qualunque e consequente specificazione geometrica dells curvatura Riemanniana. Rend. Circ. Mat. Palermo, 42(1917), 173-204. [12] J. Milnor, Morse Theory, Annals of Mathematics Studies 51, Princeton Univ. Press, 1963. [13] M. J. Micallef and j. D. Moore, Minimal two-spheres and the topology of [6]
manifolds with positive curvature on totally isotropic two planes, Ann. of Math., 127(1988), 199-227.
[14] S. B. Myers, Riemannian manifolds with positive mean curvature, Duke Math., J., 8(1941), 401-404. [15] J. Nash, The imbedding problem for Riemannian manifolds, Ann. of Math.. 63(1956), 20-63. [16] S. Nishikawa and K. Shiga, On the holomorphic equivalence of bounded domains in complete Kiibler manifolds of nonpositive curvature, J. Math. Soc. Japan, 35(1983) 273-278. [17] R. S. Palais, Morse Theory on Hilbert manifolds, Topology, 2(1963), 299-340. 201
202
BIBLIOGRAPHY
[18] R. S. Palais and S. Smale, A generalized Morse Theory, Bull Amer. Math. Soc., 70(1964), 165-172. [19] A Preissmann, Quelques proprieth globales des espaces de R.iemann, Comm. Math. Helv., 15(1943), 175-216 [20] J. Sacks and K. Uhlenbeck, The existence of minimal immersions of 2-spheres, Ann. of Math., 113(1981), 1-24. [21] S. Smale, Morse Theory and a non-linear generalization of the Dirichlet problem, Ann. of Math., 80(1964), 382-292. [22] J. T. Schwarz, Nonlinear functional analysis, Gordon and Breach, 1969. [23) Y: T. Siu, The complex-analyticity of harmonic maps and the strong rigidity
of compact Kahler manifolds, Ann. of Math., 112(1980), 73-111: Strong rigidity of compact quotients of exceptional bounded symmetric domains, Duke Math. J., 48(1981), 857-871. [24) Y.-T. Siu and S: T. Yau, Compact Kibler manifolds of positive bisectional curvature, Invent. Math., 59(1980), 857-871. [25] J. L. Synge, On the connectivity of spaces of positive curvature, Quart. J. Math., 7(1936), 316-320. [26) A. Treibergs, Lectures on the Bells-Sampson theorem on parabolic deformation of maps to harmonic maps, University of Utah notes, 1987
[27] S. Kobayashi, Complex geometry 1, Iwanami koza "Gendai sugaku no kiso", Iwanami Shoten, 1997. [28] Y. Matsumoto, Foundations of Morse Theory, Iwanami koza "Gendai sugaku no kilo", Iwanami Shoten, 1997.
[29] M. Murata and K. Kurata, Partial differential equations 1, Iwanami koza "Gendai sugaku no kiso", Iwanami Shoten, 1997. [30] T. Nagano, The global variational method, Kyoritsu koza, "Gendai sugaku no kiso", Kyoritsu Shuppen Sha, 1991. [31] H. Urakawa, The variational method and harmonic maps, Shokabo, 1990.
BIBLIOGRAPHY
203
Books 1.
T. Aubin, Nonlinear analysis on manifolds. Monge-Ampere
equations, Springer-Verlag, 1982.
This is a serious introductory book to the nonlinear problems that appear in differential geometry. This book may be regarded as a treatise on nonlinear analysis of Riemannian manifolds suitable for the reader who has learned analysis. In particular, the "Yamabe problem" and the "Calabi conjecture" are discussed in detail. 2. J. Eells and L. Lamaire, A report on harmonic maps, Bull. London Math. Soc., 10(1978), 1-68; Selected topics in harmonic maps, C. B. M. S. Regional Conference Series 50, Amer. Math. Soc., 1983; Another report on harmonic maps, Bull. London Math. Soc., 20(1988), 385-524. These three may be regarded as a comprehensive report on harmonic maps. They are best suited to survey the history and the latest developments in the study of harmonic maps. 3. R. S. Hamilton, Harmonic maps of manifolds with boundary, Lecture Notes in Mathematics 471, Springer-Verlag, 1975. This presents a solution to the Dirichlet boundary value problem and the Neumann boundary value problem under the same curvature condition as the theorem of Eells and Sampson. 4. L. Jost, Harmonic mappings between Riemannian manifolds, Proceedings of the Centre for Mathematical Analysis 4, Australian National Univ., 1983; Nonlinear method in Riemannian and Kahlerian geometry, DMV Seminar 10 Birkhauser, 1986. These articles discuss the differentiability of weak solutions to the equation of harmonic maps and the heat flow method regarding the Yang Mills connection. 5. R. Schoen and S: T. Yau, Lectures on harmonic maps, International Press, 1997.
This contains a series of lectures on the existence problem of harmonic maps between Riemann surfaces, the applications of the theory of harmonic maps to the "topological sphere theorem" and the "Frankel conjecture", and the existence of harmonic maps into spaces with singular points, etc. 6. R. Schoen and S: T. Yau, Lectures on differential geometry, International Press, 1994.
BIBLIOGRAPHY
204
This is a serious introductory book to the study of differential geometry using analytic methods- This consists of a series of lectures. The authors, from their own point of view, discuss the nonlinear analysis on Riemannian manifolds, using the nonlinear partial differential equations that appear in problems in differential geometry. There is a collection of "unsolved problems in differential geometry" at the end of the book. 7. T. Ochiai, Editor, Nonlinear problems in differential geometry, Reports on Global Analysis I, Seminar Kankokai, 1979; S. Nishikawa and T. Ochiai, Existence and applications of harmonic maps, Reports on Global Analysis II, Seminar Kankokai, 1980. Discussed in I are the "Kazdan-Warner problem", the "Yamabe problem", the "Minkowski problem", the " Calabi conjecture", etc.
II contains a treatise on the "Plateau problem" regarding minimal surfaces, and a proof for the theorem of Eells and Sampson using estimates in the Sovolev space Wk,P. 8. T. Ochiai, Editor, Minimal surfaces, Reports on Global Analysis IV, Seminar Kankokai, 1982. This consists of the differential geometric aspects of minimal surfaces (Volume 1), applications of minimal surfaces (Volume 2) and the analytics aspects of minimal surfaces (Volume 3). Volume 3 contains a detailed discussion of the solution of Douglas-Rado-Morrey regarding the existence of minimal surfaces and its properties. 9. T. Ochiai, Editor, Minimal surfaces, Reports on Global Analysis VI, XII, Seminar Kankokai, 1984, 1989.
These reports contain papers by J. L. Kazdan, Some applications of partial differential equations to problems in geometry; S. Nishikawa, On continuity of weak solutions to non-linear elliptic partial differential equations, I, II; A. Tachikawa, On differentiability of the solutions to variational problems, etc. The lecture notes of Kazdan are best suited as an introduction to these topics. 10. T. Sakai, Riemannian Geometry, Shokabo, 1992. This book contains detailed accounts on the relationship between the curvature and topology of Riemannian manifolds. Special atten-
tion is paid to the method called the comparison theorem and its application. 11.
H. Urakawa, The variational method and harmonic maps,
Shokabo, 1990.
BIBLIOGRAPHY
205
This well written book explains the general theory of the variational method and the theory of harmonic maps starting from a very basic level. In particular, it contains Uhlenbeck's proof of the theorem of Eells and Sampson from the point of view of Morse theory on infinite dimensional manifolds.
Index complete. 70 component, 3, 4, 17, 54, 166, 170 conformal. 118 connection coefficient, 20 coordinate function, 164 coordinate neighborhood, 164 cotangent bundle, 64, 169 covariant derivative, 21, 91, 111 covariant derivative in the direction of v, 21 covariant differential, 91 covariant differentiation along c, 23 critical point, 13 critical value, 13 curvature tensor, 58
a Holder continuous function, 175 cOO curve, 6
C°° vector field, 17 C°O coordinate neighborhood system, 164 COO diffeomorphic, 167 COO diffeomorphism, 166 C°O function, 162 C°O manifolds, 162
C°° map, 165 C°° section, 64, 172 COO tensor field, 55
COO variation, 48, 99 C°O vector field, 170 (r, s) tensor, 54
(r, s) tensor bundle, 64 (r, s) tensor field, 55
derivative, 166 diffeomorphism, 166 dimension, 171 direct sum, 173 Dirichlet integral, 106 Dirichlet principle, 106 distance, 42 divergence, 115 dual vector bundle, 173
action integral, 8, 89 affine connection, 19 afine parameter, 28 antiholomorphic function, 110 arc length, 13 area, 118 Banach space, 163 Bianchi first identity, 58 Bianchi second identity, 117 bounded linear operator, 174 bracket product, 18 bundle of homomorphisms, 173
elliptic, 176 energy, 8, 89 energy density, 89 equation for harmonic maps, 104 Euler's equation, 13 exponential map, 32
characteristic polynomial, 176 Christoffel symbols, 15 closed geodesic, 74 commutator product, 18 compatible, 25
fiber, 30, 64, 108, 171 fiber metric, 172 first variation formula, 13 207
208
Frechet differentiable, 174 free homotopy, 74 free homotopy class, 74 functional, 13 fundamental lemma in the theory of variation, 44 fundamental solution, 177
Gateaux differentiable, 174 Gauss lemma, 40 geodesic, 15, 27 geodesic ball, 39 geodesic flow, 45 geodesic loop, 74 geodesic sphere, 39 geodesic spray, 45 geodesically complete, 70 gradient, 115 Green's theorem, 116
Holder continuous function, 175 Holder space, 136 harmonic function, 106 harmonic map, 104 heat equation, 177 heat flow method, 119 heat operator, 177 Hilbert space, 174 holomorphic map, 110 homomorphism, 171 Hopf map, 117 horizontal component, 108 horizontal lift, 108 imbedded submanifold, 167 imbedding, 167 immersed submanifold, 167 immersion, 167 induced connection, 111 induced metric, 5 induced vector bundle, 173 integral curve, 170 isometric, 83 isometric immersion, 107 isomorphism, 172
Laplace operator, 115 Laplacian, 115 length, 7
INDEX
Levi-Civita connection, 25 Lie derivative, 45 linear connection, 19
linear partial differential operator, 176
local coordinate system, 164 local coordinates, 164 local one parameter transformation group, 171 local triviality, 171 locally finite, 168 loop, 74 maximal principle, 142 minimal submanifold, 107 minimizing sequence, 76
natural frame, 17 norm, 173 normal coordinate neighborhood, 36 normal coordinate system, 37 normed space, 173 one parameter transformation group, 171 open submanifold, 164
parabolic, 176 parabolic equation for harmonic maps, 123 paracompact, 168 parallel, 23, 92 parallel displacement, 24 parameter, 6 partition of unity, 168 piecewise smooth curve, 6 piecewise smooth variation, 47 Poincar4 upper half-space, 34 product manifold, 164 projection, 30, 169, 171 rank, 171
real hyperbolic space, 34 refinement, 168 regular, 7
Ricci curvature, 62 R.icci identity, 124, 161 Ricci tensor, 62 Riemannian connection, 25 Riemannian curvature tensor, 59
Riemannian metric, 4, 172 Riemannian submersions, 107 scalar curvature, 63 Schauder estimate, 181 second fundamental form, 96 sectional curvature, 61 smooth curve, 6 smooth manifold, 163 smooth variation, 48 space of constant curvature, 62 standard connection, 19 standard measure, 114 submanifold, 167 submersion, 107 support, 168 symmetric, 25 tangent bundle, 30, 164, 169 tangent space, 165 tangent vector, 165 tangent vector field, 22 tension field, 98 tensor field, 97 tensor product, 4, 173 totally geodesic, 162 trivial, 74 tubular neighborhood, 161 unit speed geodesic, 28
variational vector field, 48, 99 vector bundle, 171 vector field, 17 vector field along u, 38 vector subbundle, 172 vertical component, 109 Weitzenb6ck formula, 124
Variational Problems in Geometry Seiki Nishikav%wa
This volume is an outgrowth of the lectures delivered
at Tohoku University and at the Summer Graduate Program held at the Institute for Mathematics and its Applications at the University of Minnesota. It would make a suitable textbook for advanced undergraduates and graduate students.
ISBN 0-6214-1356-0
ll
9H78082111813560 MMONO/20.5
ww v.ams.or