Introduction To Riemannian Manifolds
Introduction To Riemannian Manifolds
Introduction To Riemannian Manifolds
John M. Lee
Introduction to
Riemannian
Manifolds
Second Edition
Graduate Texts in Mathematics 176
Graduate Texts in Mathematics
Series Editors:
Sheldon Axler
San Francisco State University, San Francisco, CA, USA
Kenneth Ribet
University of California, Berkeley, CA, USA
Advisory Board:
Graduate Texts in Mathematics bridge the gap between passive study and
creative understanding, offering graduate-level introductions to advanced topics in
mathematics. The volumes are carefully written as teaching aids and highlight
characteristic features of the theory. Although these books are frequently used as
textbooks in graduate courses, they are also suitable for individual study.
Introduction to
Riemannian Manifolds
Second Edition
123
John M. Lee
Department of Mathematics
University of Washington
Seattle, WA, USA
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
v
vi Preface
Many other results and techniques might reasonably claim a place in an introduc-
tory Riemannian geometry book, but they would not fit in this book without dras-
tically broadening its scope. In particular, I do not treat the Morse index theorem,
Toponogov’s theorem, or their important applications such as the sphere theorem;
Hodge theory, gauge theory, minimal surface theory, or other applications of elliptic
partial differential equations to Riemannian geometry; or evolution equations such as
the Ricci flow or the mean curvature flow. These important topics are for other, more
advanced, books.
When I wrote the first edition of this book twenty years ago, a number of superb
reference books on Riemannian geometry were already available; in the intervening
years, many more have appeared. I invite the interested reader, after reading this
book, to consult some of those for a deeper treatment of some of the topics introduced
here, or to explore the more esoteric aspects of the subject. Some of my favorites are
Peter Petersen’s admirably comprehensive introductory text [Pet16]; the elegant
introduction to comparison theory by Jeff Cheeger and David Ebin [CE08] (which
was out of print for a number of years, but happily has been reprinted by the American
Mathematical Society); Manfredo do Carmo’s much more leisurely treatment of the
same material and more [dC92]; Barrett O’Neill’s beautifully integrated introduction
to pseudo-Riemannian and Riemannian geometry [O’N83]; Michael Spivak’s classic
multivolume tome [Spi79], which can be used as a textbook if plenty of time is
available, or can provide enjoyable bedtime reading; the breathtaking survey by
Marcel Berger [Ber03], which richly earns the word “panoramic” in its title; and
the “Encyclopaedia Britannica” of differential geometry books, Foundations of
Differential Geometry by Shoshichi Kobayashi and Katsumi Nomizu [KN96]. At the
other end of the spectrum, Frank Morgan’s delightful little book [Mor98] touches on
most of the important ideas in an intuitive and informal way with lots of pictures—I
enthusiastically recommend it as a prelude to this book. And there are many more to
recommend: for example, the books by Chavel [Cha06], Gallot/Hulin/Lafontaine
[GHL04], Jost [Jos17], Klingenberg [Kli95], and Jeffrey Lee [LeeJeff09] are all
excellent in different ways.
It is not my purpose to replace any of these. Instead, I hope this book fills a niche
in the literature by presenting a selective introduction to the main ideas of the
subject in an easily accessible way. The selection is small enough to fit (with some
judicious cutting) into a single quarter or semester course, but broad enough, I hope,
to provide any novice with a firm foundation from which to pursue research or
develop applications in Riemannian geometry and other fields that use its tools.
This book is written under the assumption that the student already knows the
fundamentals of the theory of topological and smooth manifolds, as treated, for
example, in my two other graduate texts [LeeTM, LeeSM]. In particular, the
student should be conversant with general topology, the fundamental group, covering
spaces, the classification of compact surfaces, topological and smooth manifolds,
immersions and submersions, submanifolds, vector fields and flows, Lie brackets and
Lie derivatives, tensors, differential forms, Stokes’s theorem, and the basic theory of
Lie groups. On the other hand, I do not assume any previous acquaintance with
Riemannian metrics, or even with the classical theory of curves and surfaces in R3 .
(In this subject, anything proved before 1950 can be considered “classical”!)
Preface vii
Although at one time it might have been reasonable to expect most mathematics
students to have studied surface theory as undergraduates, many current North
American undergraduate math majors never see any differential geometry. Thus the
fundamentals of the geometry of surfaces, including a proof of the Gauss–Bonnet
theorem, are worked out from scratch here.
The book begins with a nonrigorous overview of the subject in Chapter 1,
designed to introduce some of the intuitions underlying the notion of curvature and
to link them with elementary geometric ideas the student has seen before. Chapter 2
begins the course proper, with definitions of Riemannian metrics and some of their
attendant flora and fauna. Here I also introduce pseudo-Riemannian metrics, which
play a central role in Einstein’s general theory of relativity. Although I do not
attempt to provide a comprehensive introduction to pseudo-Riemannian geometry,
throughout the book I do point out which of the constructions and theorems of
Riemannian geometry carry over easily to the pseudo-Riemannian case and which
do not.
Chapter 3 describes some of the most important “model spaces” of Riemannian
and pseudo-Riemannian geometry—those with lots of symmetry—with a great deal
of detailed computation. These models form a sort of leitmotif throughout the text,
serving as illustrations and testbeds for the abstract theory as it is developed.
Chapter 4 introduces connections, together with some fundamental constructions
associated with them such as geodesics and parallel transport. In order to isolate the
important properties of connections that are independent of the metric, as well as to
lay the groundwork for their further study in arenas that are beyond the scope of this
book, such as the Chern–Weil theory of characteristic classes and the Donaldson and
Seiberg–Witten theories of gauge fields, connections are defined first on arbitrary
vector bundles. This has the further advantage of making it easy to define the induced
connections on tensor bundles. Chapter 5 investigates connections in the context of
Riemannian (and pseudo-Riemannian) manifolds, developing the Levi-Civita con-
nection, its geodesics, the exponential map, and normal coordinates. Chapter 6
continues the study of geodesics, focusing on their distance-minimizing properties.
First, some elementary ideas from the calculus of variations are introduced to prove
that every distance-minimizing curve is a geodesic. Then the Gauss lemma is used to
prove the (partial) converse—that every geodesic is locally minimizing.
Chapter 7 unveils the first fully general definition of curvature. The curvature
tensor is motivated initially by the question whether all Riemannian metrics are
“flat” (that is, locally isometric to the Euclidean metric). It turns out that the failure
of parallel transport to be path-independent is the primary obstruction to the
existence of a local isometry. This leads naturally to a qualitative interpretation of
curvature as the obstruction to flatness. Chapter 8 is an investigation of submanifold
theory, leading to the definition of sectional curvatures, which give curvature a
more quantitative geometric interpretation.
The last four chapters are devoted to the development of some of the most
important global theorems relating geometry to topology. Chapter 9 gives a simple
moving-frames proof of the Gauss–Bonnet theorem, based on a careful treatment of
Hopf’s rotation index theorem (often known by its German name, the Umlaufsatz).
Chapter 10 has a largely technical nature, covering Jacobi fields, conjugate points,
viii Preface
the second variation formula, and the index form for later use in comparison
theorems. Chapter 11 introduces comparison theory, using a simple comparison
theorem for matrix Riccati equations to prove the fundamental fact that bounds on
curvature lead to bounds (in the opposite direction) on the size of Jacobi fields,
which in turn lead to bounds on many fundamental geometric quantities, such as
distances, diameters, and volumes. Finally, in Chapter 12 comes the denouement:
proofs of some of the most important local-to-global theorems illustrating the ways
in which curvature and topology affect each other.
This book contains many questions for the reader that deserve special mention.
They fall into two categories: “exercises,” which are integrated into the text, and
“problems,” grouped at the end of each chapter. Both are essential to a full
understanding of the material, but they are of somewhat different characters and
serve different purposes.
The exercises include some background material that the student should have
seen already in an earlier course, some proofs that fill in the gaps from the text,
some simple but illuminating examples, and some intermediate results that are used
in the text or the problems. They are, in general, elementary, but they are not
optional—indeed, they are integral to the continuity of the text. They are chosen
and timed so as to give the reader opportunities to pause and think over the material
that has just been introduced, to practice working with the definitions, and to
develop skills that are used later in the book. I recommend that students stop and do
each exercise as it occurs in the text, or at least convince themselves that they know
what is involved in the solution of each one, before going any further.
The problems that conclude the chapters are generally more difficult than the
exercises, some of them considerably so, and should be considered a central part
of the book by any student who is serious about learning the subject. They not only
introduce new material not covered in the body of the text, but they also provide the
student with indispensable practice in using the techniques explained in the text,
both for doing computations and for proving theorems. If the result of a problem is
used in an essential way in the text, or in a later problem, the page where it is used is
noted at the end of the problem statement. Instructors might want to present some
of these problems in class if more than a semester is available.
At the end of the book there are three appendices that contain brief reviews of
background material on smooth manifolds, tensors, and Lie groups. I have omitted
most of the proofs, but included references to other books where they may be
found. The results are collected here in order to clarify what results from topology
and smooth manifold theory this book will draw on, and also to establish definitions
and conventions that are used throughout the book. I recommend that most readers
at least glance through the appendices before reading the rest of the book, and
consider consulting the indicated references for any topics that are unfamiliar.
Preface ix
This second edition, titled Introduction to Riemannian Manifolds, has been adapted
from my earlier book Riemannian Manifolds: An Introduction to Curvature,
Graduate Texts in Mathematics 176, Springer 1997.
For those familiar with the first edition, the first difference you will notice about
this edition is that it is considerably longer than the first. To some extent, this is due
to the addition of more thorough explanations of some of the concepts. But a much
more significant reason for the increased length is the addition of many topics that
were not covered in the first edition. Here are some of the most important ones: a
somewhat expanded treatment of pseudo-Riemannian metrics, together with more
consistent explanations of which parts of the theory apply to them; a more detailed
treatment of which homogeneous spaces admit invariant metrics; a new treatment of
general distance functions and semigeodesic coordinates; introduction of the Weyl
tensor and the transformation laws for various curvatures under conformal changes
of metric; derivation of the variational equations for hypersurfaces that minimize
area with fixed boundary or fixed enclosed volume; an introduction to symmetric
spaces; and a treatment of the basic properties of the cut locus. Most importantly,
the entire treatment of comparison theory has been revamped and expanded based
on Riccati equations, and a handful of local-to-global theorems have been added
that were not present in the first edition: Cartan’s torsion theorem, Preissman’s
theorem, Cheng’s maximal diameter theorem, Milnor’s theorem on polynomial
growth of the fundamental group, and Synge’s theorem. I hope these will make the
book much more useful.
I am aware, though, that one of the attractions of the first edition for some
readers was its brevity. For those who would prefer a more streamlined path toward
the main local-to-global theorems in Chapter 12, here are topics that can be omitted
on a first pass through the book without essential loss of continuity.
• Chapter 2: Other generalizations of Riemannian metrics
• Chapter 3: Other homogeneous Riemannian manifolds and model pseudo-
Riemannian manifolds
• Chapter 5: Tubular neighborhoods, Fermi coordinates, and Euclidean and
non-Euclidean geometries
• Chapter 6: Distance functions and semigeodesic coordinates
• Chapter 7: The Weyl tensor and curvatures of conformally related metrics
• Chapter 8: Computations in semigeodesic coordinates, minimal hypersur-
faces, and constant-mean-curvature hypersurfaces
• Chapter 9: The entire chapter
• Chapter 10: Locally symmetric spaces and cut points
• Chapter 11: Günther’s volume comparison theorem and the Bishop–Gromov
volume comparison theorem
• Chapter 12: All but the theorems of Killing–Hopf, Cartan–Hadamard, and
Myers
x Preface
In addition to the major changes listed above, there are thousands of minor ones
throughout the book. Of course, I have attempted to correct all of the mistakes that I
became aware of in the first edition. Unfortunately, I surely have not been able to
avoid introducing new ones, so if you find anything that seems amiss, please let me
know by contacting me through the website listed below. I will keep an updated list
of corrections on that website.
I have also adjusted my notation and terminology to be consistent with my two
other graduate texts [LeeSM, LeeTM] and hopefully to be more consistent with
commonly accepted usage. Like those books, this one now has a notation index just
before the subject index, and it uses the same typographical conventions: mathema-
tical terms are typeset in bold italics when they are officially defined; exercises in
the text are indented, numbered consecutively with the theorems, and marked with
the special symbol I to make them easier to find; the ends of numbered examples
are marked with the symbol //; and the entire book is now set in Times Roman,
supplemented by the MathTime Professional II mathematics fonts created by
Personal TEX, Inc.
Acknowledgements
I owe an unpayable debt to the authors of the many Riemannian geometry books I
have used and cherished over the years, especially the ones mentioned above—I
have done little more than rearrange their ideas into a form that seems handy for
teaching. Beyond that, I would like to thank my Ph.D. advisor, Richard Melrose,
who many years ago introduced me to differential geometry in his eccentric but
thoroughly enlightening way; my colleagues Judith Arms, Yu Yuan, and Jim
Isenberg, who have provided great help in sorting out what topics should be
included; and all of the graduate students at the University of Washington who have
suffered with amazing grace through the many flawed drafts of both editions of this
book and have provided invaluable feedback, especially Jed Mihalisin, David
Sprehn, Collin Litterell, and Maddie Burkhart. And my deepest gratitude goes to
Ina Mette of Springer-Verlag (now at the AMS), who first convinced me to turn my
lecture notes into a book; without her encouragement, I would never have become a
textbook author.
Finally, I would like to dedicate this book to the memory of my late colleague
Steve Mitchell, who by his sparkling and joyful example taught me more about
teaching and writing than anyone.
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
1 What Is Curvature? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
The Euclidean Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Surfaces in Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Curvature in Higher Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2 Riemannian Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Methods for Constructing Riemannian Metrics . . . . . . . . . . . . . . . . . . 15
Basic Constructions on Riemannian Manifolds . . . . . . . . . . . . . . . . . . 25
Lengths and Distances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Pseudo-Riemannian Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Other Generalizations of Riemannian Metrics . . . . . . . . . . . . . . . . . . . 46
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4 Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
The Problem of Differentiating Vector Fields . . . . . . . . . . . . . . . . . . . 85
Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Covariant Derivatives of Tensor Fields . . . . . . . . . . . . . . . . . . . . . . . 95
Vector and Tensor Fields Along Curves . . . . . . . . . . . . . . . . . . . . . . . 100
xi
xii Contents
Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Parallel Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Pullback Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
7 Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
Local Invariants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
The Curvature Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
Flat Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
Symmetries of the Curvature Tensor . . . . . . . . . . . . . . . . . . . . . . . . . 202
The Ricci Identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
Ricci and Scalar Curvatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
The Weyl Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
Curvatures of Conformally Related Metrics . . . . . . . . . . . . . . . . . . . . 216
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
If you have spent some time studying modern differential geometry, with its intricate
web of manifolds, submanifolds, vector fields, Lie derivatives, tensor fields, differ-
ential forms, orientations, and foliations, you might be forgiven for wondering what
it all has to do with geometry. In most people’s experience, geometry is concerned
with properties such as distances, lengths, angles, areas, volumes, and curvature.
These concepts, however, are often barely mentioned in typical beginning graduate
courses in smooth manifold theory.
The purpose of this book is to introduce the theory of Riemannian manifolds:
these are smooth manifolds equipped with Riemannian metrics (smoothly varying
choices of inner products on tangent spaces), which allow one to measure geometric
quantities such as distances and angles. This is the branch of differential geometry in
which “geometric” ideas, in the familiar sense of the word, come to the fore. It is the
direct descendant of Euclid’s plane and solid geometry, by way of Gauss’s theory of
curved surfaces in space, and it is a dynamic subject of contemporary research.
The central unifying theme in current Riemannian geometry research is the
notion of curvature and its relation to topology. This book is designed to help you
develop both the tools and the intuition you will need for an in-depth exploration
of curvature in the Riemannian setting. Unfortunately, as you will soon discover, an
adequate development of curvature in an arbitrary number of dimensions requires
a great deal of technical machinery, making it easy to lose sight of the underlying
geometric content. To put the subject in perspective, therefore, let us begin by asking
some very basic questions: What is curvature? What are some important theorems
about it? In this chapter, we explore these and related questions in an informal way,
without proofs. The “official” treatment of the subject begins in Chapter 2.
Theorem 1.1 (Side-Side-Side). Two Euclidean triangles are congruent if and only
if the lengths of their corresponding sides are equal.
Theorem 1.2 (Angle-Sum Theorem). The sum of the interior angles of a Euclidean
triangle is .
As trivial as they may seem, these theorems serve to illustrate two major types of
results that permeate the study of geometry; in this book, we call them “classification
theorems” and “local-to-global theorems.”
The side-side-side (SSS) theorem is a classification theorem. Such a theorem
tells us how to determine whether two mathematical objects are equivalent (under
some appropriate equivalence relation). An ideal classification theorem lists a small
number of computable invariants (whatever “small” may mean in a given context),
and says that two objects are equivalent if and only if all of these invariants match.
In this case the equivalence relation is congruence, and the invariants are the three
side lengths.
The angle-sum theorem is of a different sort. It relates a local geometric property
(angle measure) to a global property (that of being a three-sided polygon or triangle).
Most of the theorems we study in this book are of this type, which, for lack of a
better name, we call local-to-global theorems.
After proving the basic facts about points and lines and the figures constructed
directly from them, one can go on to study other figures derived from the basic
elements, such as circles. Two typical results about circles are given below; the first
is a classification theorem, while the second is a local-to-global theorem. (It may not
be obvious at this point why we consider the second to be a local-to-global theorem,
but it will become clearer soon.)
Theorem 1.3 (Circle Classification Theorem). Two circles in the Euclidean plane
are congruent if and only if they have the same radius.
(Here and throughout this book, the word “curve” refers to a parametrized curve,
not a set of points. Typically, a curve will be defined as a smooth function of a real
variable t , with a prime representing an ordinary derivative with respect to t .)
Geometrically, the curvature has the following interpretation. Given a point
p D .t /, there are many circles tangent to at p—namely, those circles whose
velocity vector at p is the same as that of when both are given unit-speed
parametrizations; these are the circles whose centers lie on the line that passes
through p and is orthogonal to 0 .p/. Among these circles, there is exactly one
unit-speed parametrized circle whose acceleration vector at p is the same as that of
; it is called the osculating circle (Fig. 1.1). (If the acceleration of is zero, replace
the osculating circle by a straight line, thought of as a “circle with infinite radius.”)
The curvature is then .t / D 1=R, where R is the radius of the osculating circle.
The larger the curvature, the greater the acceleration and the smaller the osculating
circle, and therefore the faster the curve is turning. A circle of radius R has constant
curvature 1=R, while a straight line has curvature zero.
It is often convenient for some purposes to extend the definition of the curvature
of a plane curve, allowing it to take on both positive and negative values. This is done
by choosing a continuous unit normal vector field N along the curve, and assigning
the curvature a positive sign if the curve is turning toward the chosen normal or a
negative sign if it is turning away from it. The resulting function N along the curve
is then called the signed curvature.
Here are two typical theorems about plane curves.
The first of these is a classification theorem, as its name suggests. The second
is a local-to-global theorem, since it relates the local property of curvature to the
global (topological) property of being a simple closed curve. We will prove both of
these theorems later in the book: the second will be derived as a consequence of a
more general result in Chapter 9 (see Corollary 9.6); the proof of the first is left to
Problem 9-12.
It is interesting to note that when we specialize to circles, these theorems reduce
to the two theorems about circles above: Theorem 1.5 says that two circles are con-
gruent if and only if they have the same curvature, while Theorem 1.6 says that if
a circle has curvature and circumference C , then C D 2. It is easy to see that
these two results are equivalent to Theorems 1.3 and 1.4. This is why it makes sense
to regard the circumference theorem as a local-to-global theorem.
Surfaces in Space
The next step in generalizing Euclidean geometry is to start working in three dimen-
sions. After investigating the basic elements of “solid geometry”—points, lines,
planes, polyhedra, spheres, distances, angles, surface areas, volumes—one is led
to study more general curved surfaces in space (2-dimensional embedded subman-
ifolds of R3 , in the language of differential geometry). The basic invariant in this
setting is again curvature, but it is a bit more complicated than for plane curves,
because a surface can curve differently in different directions.
The curvature of a surface in space is described by two numbers at each point,
called the principal curvatures. We will define them formally in Chapter 8, but here
is an informal recipe for computing them. Suppose S is a surface in R3 , p is a point
in S , and N is a unit normal vector to S at p.
1. Choose a plane ˘ passing through p and parallel to N . The intersection of ˘
with a neighborhood of p in S is a plane curve ˘ containing p (Fig. 1.2).
2. Compute the signed curvature N of at p with respect to the chosen unit
normal N .
3. Repeat this for all normal planes ˘ . The principal curvatures of S at p,
denoted by 1 and 2 , are the minimum and maximum signed curvatures so
obtained.
Although the principal curvatures give us a lot of information about the geome-
try of S , they do not directly address a question that turns out to be of paramount
importance in Riemannian geometry: Which properties of a surface are intrinsic?
Roughly speaking, intrinsic properties are those that could in principle be measured
or computed by a 2-dimensional being living entirely within the surface. More pre-
cisely, a property of surfaces in R3 is called intrinsic if it is preserved by isometries
(maps from one surface to another that preserve lengths of curves).
To see that the principal curvatures are not intrinsic, consider the following two
embedded surfaces S1 and S2 in R3 (Figs. 1.3 and 1.4): S1 is the square in the xy-
Surfaces in Space 5
plane where 0 < x < and 0 < y < , and S2 is the half-cylinder f.x; y; z/ W z D
p
1 y 2 ; 0 < x < ; jyj < 1g. If we follow the recipe above for computing principal
curvatures (using, say, the downward-pointing unit normal), we find that, since all
planes intersect S1 in straight lines, the principal curvatures of S1 are 1 D 2 D 0.
On the other hand, it is not hard to see that the principal curvatures of S2 are 1 D 0
and 2 D 1. However, the map taking .x; y; 0/ to .x; cos y; sin y/ is a diffeomorphism
from S1 to S2 that preserves lengths of curves, and is thus an isometry.
Even though the principal curvatures are not intrinsic, the great German mathe-
matician Carl Friedrich Gauss made the surprising discovery in 1827 [Gau65] that
a particular combination of them is intrinsic. (See also [Spi79, Vol. 2] for an excel-
lent discussion of the details of Gauss’s paper.) He found a proof that the product
K D 1 2 , now called the Gaussian curvature, is intrinsic. He thought this result
was so amazing that he named it Theorema Egregium. (This does not mean “totally
awful theorem” as its English cognate egregious might suggest; a better translation
into modern colloquial English might be “totally awesome theorem.”)
6 1 What Is Curvature?
To get a feeling for what Gaussian curvature tells us about surfaces, let us look at
a few examples. Simplest of all is any surface that is an open subset of a plane: as we
have seen, such a surface has both principal curvatures equal to zero and therefore
has constant Gaussian curvature equal to zero. The half-cylinder described above
also has K D 1 2 D 0 1 D 0, as the Theorema Egregium tells us it must, being iso-
metric to a square. Another simple example is a sphere of radius R. Every normal
plane intersects the sphere in a great circle, which has radius R and therefore curva-
ture ˙1=R (with the sign depending on whether we choose the outward-pointing or
inward-pointing normal). Thus the principal curvatures are both equal to ˙1=R, and
the Gaussian curvature is 1 2 D 1=R2 . Note that while the signs of the principal
curvatures depend on the choice of unit normal, the Gaussian curvature does not: it
is always positive on the sphere.
Similarly, any surface that is “bowl-shaped” or “dome-shaped” has positive
Gaussian curvature (Fig. 1.5), because the two principal curvatures always have the
same sign, regardless of which normal is chosen. On the other hand, the Gaussian
curvature of any surface that is “saddle-shaped” (Fig. 1.6) is negative, because the
principal curvatures are of opposite signs.
The model spaces of surface theory are the surfaces with constant Gaussian cur-
vature. We have already seen two of them: the Euclidean plane R2 (K D 0), and the
sphere of radius R (K D 1=R2 ). The most important model surface with constant
negative Gaussian curvature is called the hyperbolic plane, and will be defined in
Chapter 3. It is not so easy to visualize because it cannot be realized globally as a
smoothly embedded surface in R3 (see [Spi79, Vol. 3, pp. 373–385] for a proof).
Surface theory is a highly developed branch of geometry. Of all its results, two—
a classification theorem and a local-to-global theorem—are generally acknowledged
as the most important.
Theorem 1.7 (Uniformization Theorem). Every connected 2-manifold is diffeo-
morphic to a quotient of one of the constant-curvature model surfaces described
above by a discrete group of isometries without fixed points. Thus every connected
2-manifold has a complete Riemannian metric with constant Gaussian curvature.
Theorem 1.8 (Gauss–Bonnet Theorem). Suppose S is a compact Riemannian 2-
manifold. Then
Z
K dA D 2.S /;
S
We end our survey of the basic ideas of Riemannian geometry by mentioning briefly
how curvature appears in higher dimensions. Suppose M is an n-dimensional Rie-
mannian manifold. As with surfaces, the basic geometric invariant is curvature, but
curvature becomes a much more complicated quantity in higher dimensions because
a manifold may curve in so many different directions.
The first problem we must contend with is that, in general, Riemannian mani-
folds are not presented to us as embedded submanifolds of Euclidean space. There-
fore, we must abandon the idea of cutting out curves by intersecting our manifold
with planes, as we did when defining the principal curvatures of a surface in R3 .
Instead, we need a more intrinsic way of sweeping out submanifolds. Fortunately,
geodesics—curves that are the shortest paths between nearby points—are ready-
made tools for this and many other purposes in Riemannian geometry. Examples
are straight lines in Euclidean space and great circles on a sphere.
The most fundamental fact about geodesics, which we prove in Chapter 4, is
that given any point p 2 M and any vector v tangent to M at p, there is a unique
geodesic starting at p with initial velocity v.
Here is a brief recipe for computing some curvatures at a point p 2 M .
1. Choose a 2-dimensional subspace ˘ of the tangent space to M at p.
8 1 What Is Curvature?
2. Look at all the geodesics through p whose initial velocities lie in the selected
plane ˘ . It turns out that near p these sweep out a certain 2-dimensional sub-
manifold S˘ of M , which inherits a Riemannian metric from M .
3. Compute the Gaussian curvature of S˘ at p, which the Theorema Egregium
tells us can be computed from the Riemannian metric that S˘ inherits from M .
This gives a number, denoted by sec.˘ /, called the sectional curvature of M
at p associated with the plane ˘ .
Thus the “curvature” of M at p has to be interpreted as a map
˚
sec W 2-planes in Tp M ! R:
Looking back at the remarks concluding the section on surfaces above, you can
see that these last three theorems generalize some of the consequences of the uni-
formization and Gauss–Bonnet theorems, although not their full strength. It is the
primary goal of this book to prove Theorems 1.9, 1.10, and 1.11, among others; it is
a primary goal of current research in Riemannian geometry to improve upon them
and further generalize the results of surface theory to higher dimensions.
Chapter 2
Riemannian Metrics
In this chapter we officially define Riemannian metrics, and discuss some of the
basic computational techniques associated with them. After the definitions, we
describe a few standard methods for constructing Riemannian manifolds as subman-
ifolds, products, and quotients of other Riemannian manifolds. Then we introduce
some of the elementary geometric constructions provided by Riemannian metrics,
the most important of which is the Riemannian distance function, which turns every
connected Riemannian manifold into a metric space.
At the end of the chapter, we discuss some important generalizations of Rie-
mannian metrics—most importantly, the pseudo-Riemannian metrics, followed by
brief mentions of sub-Riemannian and Finsler metrics.
Before you read this chapter, it would be a good idea to skim through the three
appendices after Chapter 12 to get an idea of the prerequisite material that will be
assumed throughout this book.
Definitions
Everything we know about the Euclidean geometry of Rn can be derived from its
dot product, which is defined for v D .v 1 ; : : : ; v n / and w D .w 1 ; : : : ; w n / by
X
n
vw D vi wi :
iD1
The dot product has a natural generalization to arbitrary vector spaces. Given a
vector space V (which we always assume to be real), an inner product on V is
a map V V ! R, typically written .v; w/ 7! hv; wi, that satisfies the following
properties for all v; w; x 2 V and a; b 2 R:
(i) S YMMETRY: hv; wi D hw; vi.
(ii) B ILINEARITY: hav C bw; xi D ahv; xi C bhw; xi D hx; av C bwi.
If two vector spaces V and W are both equipped with inner products, de-
noted by h; iV and h; iW , respectively, then a map F W V ! W is called a lin-
ear isometry if it is a vector space isomorphism that preserves inner products:
hF .v/; F .v 0 /iW D hv; v 0 iV . If V and W are inner product spaces of dimension n,
then given any choices of orthonormal bases .v1 ; : : : ; vn / for V and .w1 ; : : : ; wn / for
W , the linear map F W V ! W determined by F .vi / D wi is easily seen to be a lin-
ear isometry. Thus all inner product spaces of the same finite dimension are linearly
isometric to each other.
Riemannian Metrics
Using this inner product, we can define lengths of tangent vectors, angles between
nonzero tangent vectors, and orthogonality of tangent vectors as described above.
The length of a vector v 2 Tp M is denoted by jvjg D hv; vi1=2 g . If the metric is
understood, we sometimes omit it from the notation, and write hv; wi and jvj in
place of hv; wig and jvjg , respectively.
The starting point for Riemannian geometry is the following fundamental exam-
ple.
Example 2.6 (The Euclidean Metric). The Euclidean metric is the Riemannian
metric gx on Rn whose value at each x 2 Rn is just the usual dot product on Tx Rn
1Tx R Š R . This P w 2 Tx Rn writ-
n n
under the natural identification means that for v;P
ten in standard coordinates x ; : : : ; x as v D i v @i jx , w D j w j @j jx , we
n i
have
Xn
hv; wigx D vi wi :
iD1
Isometries
Suppose .M; g/ and M ; gz are Riemannian manifolds with or without boundary.
An isometry from .M ;g/ to M f ,e
g is a diffeomorphism ' W M ! M such that
' gz D g. Unwinding the definitions shows that this is equivalent to the requirement
that ' be a smooth bijection and each d'p W Tp M ! T'.p/ M be a lin-
differential
ear isometry. We say .M; g/ and M ; gz are isometric if there exists an isometry
between them.
A composition of isometries and the inverse of an isometry are again isometries,
so being isometric is an equivalence relation on the class of Riemannian manifolds
with or without boundary. Our subject, Riemannian geometry, is concerned primar-
ily with properties of Riemannian
manifolds that are preserved by isometries.
If .M; g/ and M ; gz are Riemannian manifolds, a map ' W M ! M is a local
isometry if each point p 2 M has a neighborhood U such that 'jU is an isometry
onto an open subset of M .
;g
I Exercise 2.7. Prove that if .M; g/ and M z are Riemannian manifolds of the same
is a local isometry if and only if ' g
dimension, a smooth map ' W M ! M z D g.
Rn with its Euclidean metric. Problem 2-1 shows that all Riemannian 1-manifolds
are flat; but we will see later that this is far from the case in higher dimensions.
An isometry from .M; g/ to itself is called an isometry of .M ,g/. The set of
all isometries of .M; g/ is a group under composition, called the isometry group
of .M ,g/; it is denoted by Iso.M; g/, or sometimes just Iso.M / if the metric is
understood.
A deep theorem of Sumner B. Myers and Norman E. Steenrod [MS39] shows
that if M has finitely many components, then Iso.M; g/ has a topology and smooth
structure making it into a finite-dimensional Lie group acting smoothly on M . We
will neither prove nor use the Myers–Steenrod theorem, but if you are interested, a
good source for the proof is [Kob72].
g D gij dx i ˝ dx j
D 12 gij dx i ˝ dx j C gj i dx i ˝ dx j
D 12 gij dx i ˝ dx j C gij dx j ˝ dx i
D gij dx i dx j :
hEi ; Ej i D ıij ;
g D ."1 /2 C C ."n /2 ;
where ."i /2 denotes the symmetric product "i "i D "i ˝ "i .
Proposition 2.8 (Existence of Orthonormal Frames). Let .M; g/ be a Riemannian
n-manifold with or without boundary. If .Xj / is any smooth local frame for TM
over an open subset U M , thenthere is a smooth orthonormal
frame .Ej / over
U such that span E1 jp ; : : : ; Ek jp D span X1 jp ; : : : ; Xk jp for each k D 1; : : : ; n
and each p 2 U . In particular, for every p 2 M , there is a smooth orthonormal
frame .Ej / defined on some neighborhood of p.
Proof. Applying the Gram–Schmidt algorithm to the vectors X1 jp ; : : : ; Xn jp at
each p 2 U , we obtain an ordered n-tuple of rough orthonormal vector fields
.E1 ; : : : ; En / over U satisfying the span conditions. Because the vectors whose
norms appear in the denominators of (2.5)–(2.6) are nowhere vanishing, those for-
mulas show that each vector field Ej is smooth. The last statement of the proposition
follows by applying this construction to any smooth local frame in a neighborhood
of p.
Warning: A common mistake made by beginners is to assume that one can find
coordinates near p such that the coordinate frame .@i / is orthonormal. Proposition
2.8 does not show this. In fact, as we will see in Chapter 7, this is possible only
when the metric is flat, that is, locally isometric to the Euclidean metric.
For a Riemannian manifold .M; g/ with or without boundary, we define the unit
tangent bundle to be the subset U TM TM consisting of unit vectors:
˚
U TM D .p; v/ 2 TM W jvjg D 1 : (2.10)
I Exercise 2.10. Use local orthonormal frames to prove the preceding proposition.
Riemannian Submanifolds
Example 2.13 (Spheres). For each positive integer n, the unit n-sphere Sn RnC1
is an embedded n-dimensional submanifold. The Riemannian metric induced on Sn
by the Euclidean metric is denoted by gV and known as the round metric or standard
metric on Sn . //
I Exercise 2.15. Prove the preceding proposition. [Hint: Apply the Gram–Schmidt algo-
rithm to a coordinate frame in slice coordinates (see Prop. A.22).]
Suppose M ; gz is a Riemannian manifold and M M is a smooth submanifold
with or without boundary in M . Given p 2 M , a vector 2 Tp M is said to be
normal to M if h; wi D 0 for every w 2 Tp M . The space of all vectors normal
to M at p is a subspace of Tp M , called the normal space at p and denoted by
?
Np M D .Tp M / . At each p 2 M , the ambient tangent space Tp M splits as an
orthogonal direct sum Tp M D Tp M ˚ Np M . A section N of the ambient tangent
jM is called a normal vector field along M if Np 2 Np M for each
bundle T M
p 2 M . The set a
NM D Np M
p2M
called the tangential and normal projections, that for each p 2 M restrict to
to Tp M and Np M , respectively.
orthogonal projections from Tp M
Proof. Given any point p 2 M , Theorem A.16 shows that there is a neighborhood
U of p in M that is embedded in M , and then Proposition 2.14 shows that there is a
smooth orthonormal frame .E1 ; : : : ; Em / that is adapted to U on some neighborhood
Uz of p in M
. This means that the restrictions of .E1 ; : : : ; En / to Uz \ U form a local
orthonormal frame for M . Given such an adapted frame, the restrictions of the last
m n vector fields .EnC1 ; : : : ; Em / to M form a smooth local frame for NM , so it
follows from Lemma A.34 that NM is a smooth subbundle.
The bundle homomorphisms > and ? are defined pointwise as orthogonal
projections onto the tangent and normal spaces, respectively, which shows that they
are uniquely defined. In terms of an adapted orthonormal frame, they can be written
> X 1 E1 C C X m Em D X 1 E1 C C X n En ;
? X 1 E1 C C X m Em D X nC1 EnC1 C C X m Em ;
I Exercise 2.18. Prove this proposition. [Hint: Use the paragraph preceding Prop. B.17
as a starting point.]
Since ıX is just the map X itself, regarded as a map into M , this is really just X gz.
The simplicity of the formula for the pullback of a tensor field makes this expression
exceedingly easy to compute, once a coordinate expression for gz is known. For
example, if M is an immersed n-dimensional Riemannian submanifold of Rm and
X W U ! Rm is a smooth local parametrization of M , the induced metric on U is
just
0 12
Xm Xm Xn
@X i Xm Xn
@X i @X i j k
g D X gx D .dX i /2 D @ dujA
D du du :
@uj @uj @uk
iD1 iD1 j D1 iD1 j;kD1
The set C is called its generating curve (see Fig. 2.1). Every smooth local parametri-
zation .t / D .a.t /; b.t // for C yields a smooth local parametrization for SC of the
form
X.t; / D .a.t / cos ; a.t / sin ; b.t //; (2.11)
provided that .t; / is restricted to a sufficiently small open set in the plane. The
t -coordinate curves t 7! X.t; 0 / are called meridians, and the -coordinate curves
7! X.t0 ; / are called latitude circles. The induced metric on SC is
Methods for Constructing Riemannian Metrics 19
Riemannian Products
Next we consider products. If .M1 ; g1 / and .M2 ; g2 / are Riemannian manifolds, the
product manifold M1 M2 has a natural Riemannian metric g D g1 ˚ g2 , called the
product metric, defined by
ˇ ˇ
g.p1 ;p2 / .v1 ; v2 /; .w1 ; w2 / D g1 ˇp .v1 ; w1 / C g2 ˇp .v2 ; w2 /; (2.12)
1 2
where .v1 ; v2 / and .w1 ; w2 / are elements of Tp1 M1 ˚ Tp2 M2 , which is natu-
rally identified with T.p1 ;p2 / .M1 M2 /. Smooth local coordinates .x 1 ; : : : ; x n / for
M1 and .x nC1 ; : : : ; x nCm / for M2 give coordinates .x 1 ; : : : ; x nCm / for M1 M2 .
In terms of these coordinates, the product metric has the local expression g D
gij dx i dx j , where .gij / is the block diagonal matrix
.g1 /ab 0
.gij / D I
0 .g2 /cd
I Exercise 2.23. Show that the induced metric on T n described in Exercise 2.21 is equal
to the product metric obtained from the usual induced metric on S1 R2 .
where .v1 ; v2 /; .w1 ; w2 / 2 Tp1 M1 ˚ Tp2 M2 as before. (Despite the similarity with
the notation for product metrics, g1 ˚ f 2 g2 is generally not a product metric unless
f is constant.) A wide variety of metrics can be constructed in this way; here are
just a few examples.
(c) If we let denote the standard coordinate function on RC R, then the map
˚. ; !/ D ! gives an isometry from the warped product RC Sn1 to
Rn X f0g with its Euclidean metric (see Problem 2-4). //
Riemannian Submersions
(that is, the tangent space to the fiber containing x), and the horizontal tangent
space at x is its orthogonal complement:
Hx D .Vx /? :
(a) Every smooth vector field W on M can be expressed uniquely in the form
W D W H C W V , where W H is horizontal, W V is vertical, and both W H
and W V are smooth.
.
(b) Every smooth vector field on M has a unique smooth horizontal lift to M
(c) For every x 2 M and v 2 Hx , there is a vector field X 2 X.M / whose hori-
zontal lift Xz satisfies Xz x D v.
22 2 Riemannian Metrics
I Exercise 2.26. Let W R2 ! R be the projection map .x; y/ D x, and let W be the
smooth vector field y@x on R2 . Show that W is horizontal, but there is no vector field on
R whose horizontal lift is equal to W .
Now we can identify some quotients of Riemannian manifolds that inherit met-
rics of their own.
Let
us begin by describing what such a metric should look like.
Suppose M ; gz and .M; g/ are Riemannian manifolds, and W M ! M is a
smooth submersion. Then is said to be a Riemannian submersion if for each
x2M , the differential d x restricts to a linear isometry from Hx onto T.x/ M . In
other words, gzx .v; w/ D g.x/ .d x .v/; d x .w// whenever v; w 2 Hx .
Proof. Under the given hypotheses, the quotient manifold theorem (Thm. C.17)
shows that M has a unique smooth manifold structure such that the quotient map
W M ! M is a smooth submersion. It follows easily from the definitions in that
case that the given action of G on M is vertical and transitive on fibers. Since the
action is also isometric, Theorem 2.28 shows that M inherits a unique Riemannian
metric making into a Riemannian submersion.
Here is an important example of a Riemannian metric defined in this way. A
larger class of such metrics is described in Problem 2-7.
24 2 Riemannian Metrics
Example 2.30 (The Fubini–Study Metric). Let n be a positive integer, and con-
sider the complex projective space CP n defined in Example C.19. That example
shows that the map W C nC1 X f0g ! CP n sending each point in C nC1 X f0g to
its span is a surjective smooth submersion. Identifying C nC1 with R2nC2 endowed
with its Euclidean metric, we can view the unit sphere S2nC1 with its round metric gV
as an embedded Riemannian submanifold of C nC1 X f0g. Let p W S2nC1 ! CP n de-
note the restriction of the map . Then p is smooth, and it is surjective, because every
1-dimensional complex subspace contains elements of unit norm. We need to show
that it is a submersion. Let z0 2 S2nC1 and set
0 D p.z0 / 2 CP n . Since is a smooth
submersion, it has a smooth local section W U ! C nC1 defined on a neighborhood
U of
0 and satisfying .
0 / D z0 (Thm. A.17). Let W C nC1 X f0g ! S2nC1 be the
radial projection onto the sphere:
z
.z/ D :
jzj
Since dividing an element of C nC1 by a nonzero scalar does not change its span, it
follows that p ı D . Therefore, if we set z D ı , we have p ı z D p ı ı D
ı D IdU , so z is a local section of p. By Theorem A.17, this shows that p is a
submersion.
Define an action of S1 on S2nC1 by complex multiplication:
z 1 ; : : : ; z nC1 D z 1 ; : : : ; z nC1 ;
for 2 S1 (viewed as a complex number of norm 1) and z D z 1 ; : : : ; z nC1 2
S2nC1 . This is easily seen to be isometric, vertical, and transitive on fibers of p.
By Theorem 2.28, therefore, there is a unique metric on CP n such that the map
p W S2nC1 ! CP n is a Riemannian submersion. This metric is called the Fubini–
Study metric; you will have a chance to study its geometric properties in Problems
3-19 and 8-13. //
Riemannian Coverings
Another important special case of
Riemannian submersions occurs in the context of
covering maps. Suppose M ; gz and .M; g/ are Riemannian manifolds. A smooth
! M is called a Riemannian covering if it is a local isometry.
covering map W M
Proof. Proposition A.49 shows that is a surjective smooth submersion. The au-
tomorphism group acts vertically by definition, and Proposition C.21 shows that it
acts transitively on fibers when the covering is normal. It then follows from Theorem
2.28 that there is a unique metric g on M such that is a Riemannian submersion.
Basic Constructions on Riemannian Manifolds 25
One elementary but important property of Riemannian metrics is that they allow us
to convert vectors to covectors and vice versa. Given a Riemannian metric g on M ,
we define a bundle homomorphism gy W TM ! T M by setting
gy.v/.w/ D gp .v; w/
Thus the matrix of gy in any local frame is the same as the matrix of g itself.
Given a vector field X , it is standard practice to denote the components of the
covector field gy.X / by
Xj D gij X i ;
so that
gy.X / D Xj "j ;
and we say that gy.X / is obtained from X by lowering an index. With this in mind,
the covector field gy.X / is denoted by X [ and called X flat, borrowing from the
musical notation for lowering a tone.
Because the matrix .gij / is nonsingular at each point, the map gy is invertible, and
1
of gy is just the inverse matrix of .gij /. We denote this inverse matrix
the matrix
by g ij , so that g ij gj k D gkj g j i D ıki . The symmetry of gij easily implies that
ij
g is also symmetric in i and j . In terms of a local frame, the inverse map gy1 is
given by
gy1 .!/ D ! i Ei ;
where
! i D g ij !j : (2.13)
If ! is a covector field, the vector field gy1 .!/ is called (what else?) ! sharp and
denoted by ! ] , and we say that it is obtained from ! by raising an index. The two
inverse isomorphisms [ and ] are known as the musical isomorphisms.
Probably the most important application of the sharp operator is to extend the
classical gradient operator to Riemannian manifolds. If g is a Riemannian metric
Basic Constructions on Riemannian Manifolds 27
Thus if .Ei / is an orthonormal frame, then grad f is the vector field whose
components are the same as the components of df ; but in other frames, this will
not be the case.
The next proposition shows that the gradient has the same geometric interpre-
tation on a Riemannian manifold as it does in Euclidean space. If f is a smooth
real-valued function on a smooth manifold M , recall that a point p 2 M is called a
regular point of f if dfp ¤ 0, and a critical point of f otherwise; and a level set
f 1 .c/ is called a regular level set if every point of f 1 .c/ is a regular point of f
(see Appendix A). Corollary A.26 shows that each regular level set is an embedded
smooth hypersurface in M .
A D Ai j k "i ˝ Ej ˝ "k
28 2 Riemannian Metrics
(see (B.6)), we can lower its middle index to obtain a covariant 3-tensor A[ with
components
Aij k D gj l Ai l k :
To avoid overly cumbersome notation, we use the symbols F ] and F [ without
explicitly specifying which index position the sharp or flat operator is to be applied
to; when there is more than one choice, we will always stipulate in words what is
meant.
Another important application of the flat and sharp operators is to extend the
trace operator introduced in Appendix B to covariant tensors. If h is any covariant
k-tensor field on a Riemannian manifold with k 2, we can raise one of its indices
(say the last one for definiteness) and obtain a .1; k 1/-tensor h] . The trace of h] is
thus a well-defined covariant .k 2/-tensor field (see Exercise B.3). We define the
trace of h with respect tog as
trg h D tr h] :
Sometimes we may wish to raise an index other than the last, or to take the trace on
a pair of indices other than the last covariant and contravariant ones. In each such
case, we will say in words what is meant.
The most important case is that of a covariant 2-tensor field. In this case, h] is
a .1; 1/-tensor field, which can equivalently be regarded as an endomorphism field,
and trg h is just the ordinary trace of this endomorphism field. In terms of a basis,
this is
trg h D hi i D g ij hij :
In particular, in an orthonormal frame this is the ordinary trace of the matrix .hij /
(the sum of its diagonal entries); but if the frame is not orthonormal, then this trace
is different from the ordinary trace.
h!; ig D h! ] ; ] ig :
Basic Constructions on Riemannian Manifolds 29
(Just as with inner products of vectors, we might sometimes omit g from the notation
when the metric is understood.) To see how to compute this, we just use the basis
formula (2.13) for the sharp operator, together with the relation gkl g ki D glk g ki D
ıli , to obtain
h!; i D gkl g ki !i g lj j
D ıli g lj !i j
D g ij !i j :
In other
words, the inner product on covectors is represented by the inverse ma-
trix g ij . Using our conventions for raising and lowering indices, this can also be
written
h!; i D !i i D ! j j :
I Exercise 2.39. Let .M; g/ be a Riemannian manifold with or without boundary, let
.Ei / be a local frame for M , and let ."i / be its dual coframe. Show that the following are
equivalent:
(a) .Ei / is orthonormal.
(b) ."i / is orthonormal.
(c) ."i /] D Ei for each i.
This construction can be extended to tensor bundles of any rank, as the following
proposition shows. First a bit of terminology: if E ! M is a smooth vector bundle, a
smooth fiber metric on E is an inner product on each fiber Ep that varies smoothly,
in the sense that for any (local) smooth sections ; of E, the inner product h; i
is a smooth function.
With this inner product, if .E1 ; : : : ; En / is a local orthonormal frame for TM and
."1 ; : : : ; "n / is the corresponding dual coframe, then the collection of tensor fields
Ei1 ˝ ˝ Eik ˝ "j1 ˝ ˝ "jl as all the indices range from 1 to n forms a local
orthonormal frame for T .k;l/ .Tp M /. In terms of any (not necessarily orthonormal)
frame, this fiber metric satisfies
i :::i
hF; Gi D gi1 r1 gik rk g j1 s1 g jl sl Fj11:::jkl Gsr11:::s
:::rk
l
: (2.16)
where the last factor on the right represents the components of G with all of its
indices raised:
G j1 :::jl D g j1 s1 g jl sl Gs1 :::sl :
Proposition 2.41 (The Riemannian Volume Form). Let .M; g/ be an oriented Rie-
mannian n-manifold with or without boundary. There is a unique n-form d Vg on
M , called the Riemannian volume form, characterized by any one of the following
three equivalent properties:
(a) If "1 ; : : : ; "n is any local oriented orthonormal coframe for T M , then
d Vg D "1 ^ ^ "n :
d Vg .E1 ; : : : ; En / D 1:
1
(c) If x ; : : : ; x n are any oriented local coordinates, then
q
d Vg D det.gij / dx 1 ^ ^ dx n :
without boundary, then d Vg is never exact, because its integral over M is positive,
and exact forms integrate to zero by Stokes’s theorem.
Because there are two conventions in common use for the wedge product (see
p. 401), it should be noted that properties (a) and (c) of Proposition 2.41 are the
same regardless of which convention is used; but property (b) holds only for the
determinant convention that we use. If the Alt convention is used, the number 1
should be replaced by 1=nŠ in that formula.
I Exercise 2.42. Suppose .M; g/ and M;g
z are oriented Riemannian manifolds, and
is an orientation-preserving isometry. Prove that ' d Vgz D d Vg .
'W M ! M
.E1 ; : : : ; En / D 1 (2.18)
(Note that many books, including [LeeSM] and the first edition of this book, define
the Laplacian as div.grad u/. The main reason for choosing the negative sign is
so that the operator will have nonnegative eigenvalues; see Problem 2-24. But the
definition we give here is much more common in Riemannian geometry.)
The next proposition gives alternative formulas for these operators.
where det g D det.gkl / is the determinant of the component matrix of g in these co-
ordinates. On Rn with the Euclidean metric and standard coordinates, these reduce
to
X n
@ @X i
div X i i D ;
@x @x i
iD1
X
n
@2 u
u D :
.@x i /2
iD1
Perhaps the most important tool that a Riemannian metric gives us is the ability to
measure lengths of curves and distances between points. Throughout this section,
.M; g/ denotes a Riemannian manifold with or without boundary.
Without further qualification, a curve in M always means a parametrized curve,
that is, a continuous map W I ! M , where I R is some interval. Unless
otherwise specified, we will not worry about whether the interval is bounded or
unbounded, or whether it includes endpoints or not. To say that is a smooth curve
is to say that it is smooth as a map from the manifold (with boundary) I to M . If
I has one or two endpoints and M has empty boundary, then is smooth if and
only if it extends to a smooth curve defined on some open interval containing I . (If
@M ¤ ¿, then smoothness of has to be interpreted as meaning that each coor-
dinate representation of has a smooth extension to an open interval.) A curve
segment is a curve whose domain is a compact interval.
A smooth curve W I ! M has a well-defined velocity 0 .t / 2 T.t/ M for each
t 2 I . We say that is a regular curve if 0 .t / ¤ 0 for t 2 I . This implies that is
an immersion, so its image has no “corners” or “kinks.”
We wish to use curve segments as “measuring tapes” to define distances between
points in a connected Riemannian manifold. Many aspects of the theory become
technically much simpler if we work with a slightly larger class of curve segments
instead of just the regular ones. We now describe the appropriate class of curves.
If Œa; b R is a closed bounded interval, a partition of Œa; b is a finite sequence
.a0 ; : : : ; ak / of real numbers such that a D a0 < a1 < < ak D b. Each interval
Œai1 ; ai is called a subinterval of the partition. If M is a smooth manifold with
or without boundary, a (continuous) curve segment W Œa; b ! M is said to be
34 2 Riemannian Metrics
piecewise regular if there exists a partition .a0 ; : : : ; ak / of Œa; b such that jŒai 1 ;ai
is a regular curve segment (meaning it is smooth with nonvanishing velocity) for
i D 1; : : : ; k. For brevity, we refer to a piecewise regular curve segment as an admis-
sible curve, and any partition .a0 ; : : : ; ak / such that jŒai 1 ;ai is smooth for each
i an admissible partition for . (There are many admissible partitions for a given
admissible curve, because we can always add more points to the partition.) It is also
convenient to consider any map W fag ! M whose domain is a single real number
to be an admissible curve.
Suppose is an admissible curve and .a0 ; : : : ; ak / is an admissible partition for
it. At each of the intermediate partition points a1 ; : : : ; ak1 , there are two one-sided
velocity vectors, which we denote by
0 ai D lim 0 .t /;
t%ai
0 C
ai D lim 0 .t /:
t&ai
(b) PARAMETER
I NDEPENDENCE : If z is a reparametrization of , then Lg . / D
Lg z .
(c) I SOMETRY I NVARIANCE : If .M; g/ and M ; gz are Riemannian manifolds
(with or without boundary) and ' W M ! M is a local isometry, then Lg . / D
Lgz .' ı /.
Lengths and Distances 35
has an admissible partition .a0 ; : : : ; akC1 /. The inductive hypothesis gives piece-
wise regular homeomorphisms ' W Œ0; c ! Œa; ak and W Œ0; d ! Œak ; b such
that ı ' is an arc-length reparametrization of jŒa;ak and ı is an arc-length
reparametrization of jŒak ;b . If we define 'z W Œ0; c C d ! Œa; b by
(
'.s/; s 2 Œ0; c;
z D
'.s/
.s c/; s 2 Œc; c C d ;
Since is continuous and .0/ D 0, it follows that .s/ D s for all s 2 Œ0; c, and
thus z D y.
We are now in a position to introduce one of the most important concepts from
classical geometry into the Riemannian setting: distances between points.
Suppose .M; g/ is a connected Riemannian manifold with or without boundary.
For each pair of points p; q 2 M , we define the Riemannian distance from p to
q, denoted by dg .p; q/, to be the infimum of the lengths of all admissible curves
from p to q. The following proposition guarantees that dg .p; q/ is a well-defined
nonnegative real number for each p; q 2 M .
Proposition 2.50. If M is a connected smooth manifold (with or without boundary),
then any two points of M can be joined by an admissible curve.
Proof. Let p; q 2 M be arbitrary. Since a connected manifold is path-connected,
p and q can be connected by a continuous path c W Œa; b ! M . By compactness,
there is a partition of Œa; b such that c Œai1 ; ai is contained in a single smooth
coordinate ball (or half-ball in case @M ¤ ¿) for each i . Then we may replace each
such curve segment by a straight-line path in coordinates, yielding an admissible
curve between the same points (Fig. 2.2).
For convenience, if .M; g/ is a disconnected Riemannian manifold, we also let
dg .p; q/ denote the Riemannian distance from p to q, provided that p and q lie in
the same connected component of M . (See also Problem 2-30.)
Lengths and Distances 37
(Note that unlike lengths of curves, Riemannian distances are not necessarily pre-
served by local isometries; see Problem 2-31.)
We wish to show that the Riemannian distance function turns M into a metric
space, whose metric topology is the same as its original manifold topology. To do
so, we need the following lemmas.
The same inequality holds trivially when v D 0. Arguing similarly in the other direc-
tion, we conclude that (2.21) holds for all x 2 K and v 2 Tx Rn .
38 2 Riemannian Metrics
The next lemma shows how to transfer this result to Riemannian manifolds.
Lemma 2.54. Let .M; g/ be a Riemannian manifold with or without boundary and
let dg be its Riemannian distance function. Suppose U is an open subset of M and
p 2 U . Then p has a coordinate neighborhood V U with the property that there
are positive constants C; D satisfying the following inequalities:
(a) If q 2 V , then dg .p; q/ Cdgx .p; q/, where gx is the Euclidean metric in the
given coordinates on V .
(b) If q 2 M X V , then dg .p; q/ D.
Proof. Let W be any neighborhood of p contained in U on which there exist smooth
coordinates x i . Using these coordinates, we can identify W with an open subset
of Rn or RnC . Let K be a compact subset of W containing a neighborhood V of
p, chosen as follows: If p is an interior point of M , let K be the closed Euclidean
ball Bx" .0/, with " > 0 chosen small enough that K W , and let V D B" .0/. If
p 2 @M , let K be a closedP half-ball Bx" .0/\RnC with " chosen similarly, and let V D
B" .0/ \ RC . Let gx D i .dx / denote the Euclidean metric in these coordinates,
n i 2
and let c; C be constants satisfying (2.21). If is any admissible curve whose image
lies entirely in V , it follows that
Taking the infimum over all such curves , we obtain (b) with D D c".
Since every compact metric space is bounded, every compact connected Riemannian
manifold with or without boundary has finite diameter. (Note that the unit sphere
with the Riemannian distance determined by the round metric has diameter , not
2, since the Riemannian distance between antipodal points is . See Problem 6-2.)
40 2 Riemannian Metrics
Pseudo-Riemannian Metrics
From the point of view of geometry, Riemannian metrics are by far the most impor-
tant structures that manifolds carry. However, there is a generalization of Rieman-
nian metrics that has become especially important because of its application to
physics.
Before defining this generalization, we begin with some linear-algebraic pre-
liminaries. Suppose V is a finite-dimensional vector space, and q is a symmetric
covariant 2-tensor on V (also called a symmetric bilinear form). Just as for an inner
product, there is a linear map qy W V ! V defined by
Proof. As in the positive definite case, the basis .bi / is constructed recursively,
starting with b1 D v1 =jv1 j and noting that the assumption that v1 spans a nonde-
generate subspace ensures that jv1 j ¤ 0. At the inductive step, assuming we have
constructed .b1 ; : : : ; bk /, we first set
X
k
hvkC1 ; bi i
z D vkC1 bi :
hbi ; bi i
iD1
Corollary
2.64.
Suppose .V; q/ is an n-dimensional scalar product space. There is
a basis ˇ i for V with respect to which q has the expression
2 2 2 2
q D ˇ 1 C C ˇ r ˇ rC1 ˇ rCs ; (2.23)
By far the most important pseudo-Riemannian metrics (other than the Riemann-
ian ones) are the Lorentz metrics, which are pseudo-Riemannian metrics of index 1,
and thus signature .r; 1/. (Some authors, especially in the physics literature, prefer
to use signature .1; r/; either one can be converted to the other by multiplying the
metric by 1, so there is no significant difference.)
The pseudo-Euclidean metric qx.r;1/ is a Lorentz metric called the Minkowski
metric, and the Lorentz manifold Rr;1 is called .r C 1/-dimensional Minkowski
space. If we denote standard coordinates on Rr;1 by 1 ; : : : ; r ; , then the Min-
kowski metric is given by
2 2
qx.r;1/ D d 1 C C d r .d /2 : (2.25)
In the special case of R3;1 , the Minkowski metric is the fundamental invariant of
Albert Einstein’s special theory of relativity, which can be expressed succinctly by
saying that if gravity is ignored, then spacetime is accurately modeled by .3 C 1/-
dimensional Minkowski space, and the laws of physics have the same form in every
coordinate system in which the Minkowski metric has the expression (2.25). The
differing physical characteristics of “space” (the i directions) and “time” (the
direction) arise from the fact that they are subspaces on which the metric is positive
definite and negative definite, respectively. The general theory of relativity includes
gravitational effects by allowing the Lorentz metric to vary from point to point.
Many, but not all, results from the theory of Riemannian metrics apply equally
well to pseudo-Riemannian metrics. Throughout this book, we will attempt to point
out which results carry over directly to pseudo-Riemannian metrics, which ones
can be adapted with minor modifications, and which ones do not carry over at all.
As a rule of thumb, proofs that depend only on the nondegeneracy of the metric
tensor, such as properties of the musical isomorphisms and existence and uniqueness
of geodesics, work fine in the pseudo-Riemannian setting, while proofs that use
positivity in an essential way, such as those involving lengths of curves, do not.
44 2 Riemannian Metrics
One notable result that does not carry over to the pseudo-Riemannian case is
Proposition 2.4, about the existence of metrics. For example, the following result
characterizes those manifolds that admit Lorentz metrics.
Pseudo-Riemannian Submanifolds
The theory of submanifolds
is only slightly more complicated in the pseudo-
Riemannian case. If M ; gz is a pseudo-Riemannian manifold, a smooth submani-
fold W M ,! M is called a pseudo-Riemannian submanifold of M if gz is non-
degenerate with constant signature. If this is the case, we always consider M to be
endowed with the induced pseudo-Riemannian metric gz. In the special case in
which gz is positive definite, M is called a Riemannian submanifold.
The nondegeneracy hypothesis is not automatically satisfied: for example, if
M R1;1 is the submanifold f.; / W D g and W M ! R1;1 is inclusion, then
the pullback tensor qx.1;1/ is identically zero on M .
For hypersurfaces (submanifolds of codimension 1), the nondegeneracy condi-
tion is easy to check. If M M is a smooth submanifold and p 2 M , then a vector
2 Tp M is said to be normal to M if h; xi D 0 for all x 2 Tp M , just as in the
Riemannian case. The space of all normal vectors at p is a subspace of Tp M de-
noted by Np M .
Proposition 2.70. Suppose M ; gz is a pseudo-Riemannian manifold of signature
.r; s/. Let M be a smooth hypersurface in M , and let W M ,! M
be the inclusion
map. Then the pullback tensor field gz is nondegenerate if and only if gz.; / ¤ 0
for every p 2 M and every nonzero normal vector 2 Np M . If gz.; / > 0 for
every nonzero normal vector to M , then M is a pseudo-Riemannian submanifold of
signature .r 1; s/; and if gz.; / < 0 for every such vector, then M is a pseudo-
Riemannian submanifold of signature .r; s 1/.
Now suppose gz.; / > 0 for every nonzero normal vector . Let p 2 M be
arbitrary, and let be a nonzero element of Np M . Writing n D dim M , we can
complete to a nondegenerate basis .; w2 ; : : : ; wn / for Tp M by Lemma 2.62, and
then use the Gram–Schmidt algorithm to find an orthonormal basis .b1 ; : : : ; bn/ for
Tp M such that span.b1 / D Np M . It follows that span.b2 ; : : : ; bn / D Tp M . If ˇ j
is the dual basis to .bi /, then gzp has a basis representation of the form .ˇ 1 /2 ˙
.ˇ 2 /2 ˙ ˙ .ˇ n /2 , with a total of r positive terms and s negative ones, and with
a positive sign on the first term .ˇ 1 /2 . Therefore, gzp D ˙.ˇ 2 /2 ˙ ˙ .ˇ n /2 has
signature .r 1; s/. The argument for the case gz.; / < 0 is similar.
If M ; gz is a pseudo-Riemannian manifold and f 2 C 1 .M /, then the gradient of
f is defined as the smooth vector field grad f D .df /] just as in the Riemannian
case.
Corollary 2.71. Suppose M ; gz is a pseudo-Riemannian manifold of signature
.r; s/, f 2 C 1 M , and M D f 1 .c/ for some c 2 R. If gz.grad f; grad f / > 0
everywhere on M , then M is an embedded pseudo-Riemannian submanifold of M
of signature .r 1; s/; and if gz.grad f; grad f / < 0 everywhere on M , then M is an
embedded pseudo-Riemannian submanifold of M of signature .r; s 1/. In either
case, grad f is everywhere normal to M .
Proposition
2.72 (Pseudo-Riemannian Adapted Orthonormal Frames). Suppose
M ; gz is a pseudo-Riemannian manifold, and M M is an embedded pseudo-
Riemannian or Riemannian submanifold. For each p 2 M , there exists a smooth
orthonormal frame on a neighborhood of p in M that is adapted to M .
Corollary
2.73
(Normal Bundle to a Pseudo-Riemannian Submanifold). Sup-
pose M ; gz is a pseudo-Riemannian manifold, and M M is an embedded
pseudo-Riemannian or Riemannian submanifold. The set of vectors normal to M
jM , called the normal bundle to M .
is a smooth vector subbundle of T M
46 2 Riemannian Metrics
Sub-Riemannian Metrics
The first generalization arises when we relax the requirement that the metric be
defined on the whole tangent space.
A sub-Riemannian metric (also sometimes known as a singular Riemannian
metric or Carnot–Carathéodory metric) on a smooth manifold M is a smooth fiber
metric on a smooth tangent distribution S TM (i.e., a vector subbundle of TM ).
Since lengths make sense only for vectors in S , the only curves whose lengths can
be measured are those whose velocity vectors lie everywhere in S . Therefore, one
usually imposes some condition on S that guarantees that any two nearby points
can be connected by such a curve. This is, in a sense, the opposite of the Frobenius
integrability condition, which would restrict every such curve to lie in a single leaf
of a foliation.
Sub-Riemannian metrics arise naturally in the study of the abstract models of
real submanifolds of complex space C n , called CR manifolds (short for Cauchy–
Riemann manifolds). CR manifolds are real manifolds endowed with a tangent dis-
tribution S TM whose fibers carry the structure of complex vector spaces (with
an additional integrability condition that need not concern us here). In the model
case of a submanifold M C n , Spis the set of vectors tangent to M that remain
tangent after multiplication by i D 1 in the ambient complex coordinates. If S is
sufficiently far from being integrable, choosing a fiber metric on S results in a sub-
Riemannian metric whose geometric properties closely reflect the complex-analytic
properties of M as a subset of C n .
Another motivation for studying sub-Riemannian metrics arises from control the-
ory. In this subject, one is given a manifold with a vector field depending on parame-
ters called controls, with the goal being to vary the controls so as to obtain a solution
curve with desired properties, often one that minimizes some function such as arc
length. If the vector field is constrained to be everywhere tangent to a distribution S
on the manifold (for example, in the case of a robot arm whose motion is restricted
by the orientations of its hinges), then the function can often be modeled as a sub-
Riemannian metric and optimal solutions modeled as sub-Riemannian geodesics.
A useful introduction to the geometry of sub-Riemannian metrics is provided in
the article [Str86].
Problems 47
Finsler Metrics
Another important generalization arises from relaxing the requirement that norms
of vectors be defined in terms of an inner product on each tangent space.
In general, a norm on a vector space V is a real-valued function on V , usually
written v 7! jvj, that satisfies
(i) H OMOGENEITY: jcvj D jcj jvj for v 2 V and c 2 R;
(ii) P OSITIVITY: jvj 0 for v 2 V , with equality if and only if v D 0;
(iii) T RIANGLE I NEQUALITY: jv C wj jvj C jwj for v; w 2 V .
For example, the length function associated with an inner product is a norm.
Now suppose M is a smooth manifold. A Finsler metric on M is a continuous
function F W TM ! R, smooth on the set of nonzero vectors, whose restriction to
each tangent space Tp M is a norm. Again, the norm function associated with a
Riemannian metric is a special case.
The inventor of Riemannian geometry himself, Bernhard Riemann, clearly envis-
aged an important role in n-dimensional geometry for what we now call Finsler met-
rics; he restricted his investigations to the “Riemannian” case purely for simplicity
(see [Spi79, Vol. 2]). However, it was not until the late twentieth century that Finsler
metrics began to be studied seriously from a geometric point of view.
The recent upsurge of interest in Finsler metrics has been motivated in part by
the fact that two different Finsler metrics appear very naturally in the theory of
several complex variables. For certain bounded open sets in C n (the ones with
smooth, strictly convex boundaries, for example), the Kobayashi metric and the
Carathéodory metric are intrinsically defined, biholomorphically invariant Finsler
metrics. Combining differential-geometric and complex-analytic methods has led
to striking new insights into both the function theory and the geometry of such
domains. We do not treat Finsler metrics further in this book, but you can consult
one of the recent books on the subject [AP94, BCS00, JP13].
Problems
2-1. Show that every Riemannian 1-manifold is flat. (Used on pp. 13, 193.)
2-2. Suppose V and W are finite-dimensional real inner product spaces of the
same dimension, and F W V ! W is any map (not assumed to be linear or
even continuous) that preserves the origin and all distances: F .0/ D 0 and
jF .x/ F .y/j D jx yj for all x; y 2 V . Prove that F is a linear isometry.
[Hint: First show that F preserves inner products, and then show that it is linear.]
(Used on p. 187.)
2-3. Given a smooth embedded 1-dimensional submanifold C H as in Exam-
ple 2.24(b), show that the surface of revolution SC R3 with its induced
metric is isometric to the warped product C a S1 , where a W C ! R is the
distance to the z-axis.
48 2 Riemannian Metrics
2-12. Prove Proposition 2.41 (existence and uniqueness of the Riemannian volume
form).
2-13. Prove Proposition 2.43 (characterizing the volume form of a Riemannian
hypersurface). [Hint: To prove (2.17), use an adapted orthonormal frame.]
2-14. Suppose M ; gz and .M; g/ are compact connected Riemannian manifolds,
and W M ! M is a k-sheeted Riemannian covering. Prove that Vol M D
k Vol.M /. (Used on p. 363.)
2-15. Suppose .M1 ; g1 / and .M2 ; g2 / are oriented Riemannian manifolds of di-
mensions k1 and k2 , respectively. Let f W M1 ! RC be a smooth func-
tion, and let g D g1 ˚ f 2 g2 be the corresponding warped product metric
on M1 f M2 . Prove that the Riemannian volume form of g is given by
d Vg D f k2 d Vg1 ^ d Vg2 ;
(b) Show that the divergence operator satisfies the following product rule for
u 2 C 1 .M / and X 2 X.M /:
grad
- X.M / curl
- X.M / div
- C 1 .M /
C 1 .M /
Id [ ˇ
(2.27)
? ? ? ?
0 .M / - 1 .M / - 2 .M / - 3 .M /;
d d d
where
.f / D f d Vg , and use this to prove that curl.grad f / D 0 for
every f 2 C 1 .M /, and div.curl X / D 0 for every X 2 X.M /.
(c) Compute the formula for the curl in standard coordinates on R3 with
the Euclidean metric.
2-28. Let .M; g/ be an oriented Riemannian manifold and let
denote its Hodge
star operator (Problem 2-18). Show that for every X 2 X.M /,
X ³ d Vg D
X [ ;
div X D
d
X [ ;
where
is the Hodge star operator defined in Problem 2-18. Extend this
definition to 0-forms by defining d ! D 0 for ! 2 0 .M /.
(a) Show that d ı d D 0.
(b) Show that the formula
Z
.!; / D h!; ig d Vg
M
defines an inner product on k .M / for each k, where h; ig is the inner
product on forms defined in Problem 2-16.
(c) Show that .d !; / D .!; d/ for all ! 2 k .M / and 2 k1 .M /.
2-30. Suppose .M; g/ is a (not necessarily connected) Riemannian manifold.
Show that there is a distance function d on M that induces the given topol-
ogy and restricts to the Riemannian distance on each component of M .
(Used on p. 187.)
2-31. Suppose .M; g/ and M ; gz are connected Riemannian manifolds, and
'W M ! M is a local isometry. Show that dgz .'.x/; '.y// dg .x; y/ for
all x; y 2 M . Give an example to show that equality need not hold. (Used
on p. 37.)
2-32. Let .M; g/ be a Riemannian manifold and W Œa; b ! M a smooth curve
segment. For each continuous function f W Œa;
R b ! R, we define the integral
of f with respect to arc length, denoted by f ds, by
Z Z b
f ds D f .t / j 0 .t /jg dt:
a
R
(a) Show that f ds is independent of parametrization in the following
sense: if ' W Œc; d ! Œa; b is a diffeomorphism, then
Z Z
b d ˇ ˇ
0
f .t / j .t /jg dt D fz.u/ ˇz0 .u/ˇg du;
a c
Before we delve into the general theory of Riemannian manifolds, we pause to give
it some substance by introducing a variety of “model Riemannian manifolds” that
should help to motivate the general theory. These manifolds are distinguished by
having a high degree of symmetry.
We begin by describing the most symmetric model spaces of all—Euclidean
spaces, spheres, and hyperbolic spaces. We analyze these in detail, and prove that
each one has a very large isometry group: not only is there an isometry taking any
point to any other point, but in fact one can find an isometry taking any
orthonormal basis at one point to any orthonormal basis at any other point. As we
will see in Chapter 8, this has strong consequences for the curvatures of these man-
ifolds.
After introducing these very special models, we explore some more general
classes of Riemannian manifolds with symmetry—the invariant metrics on Lie
groups, homogeneous spaces, and symmetric spaces.
At the end of the chapter, we give a brief introduction to some analogous models
in the pseudo-Riemannian case. For the particular case of Lorentz manifolds, these
are the Minkowski spaces, de Sitter spaces, and anti-de Sitter spaces, which are
important model spaces in general relativity.
The isometry group does more than just act on M itself. For every ' 2 Iso.M; g/,
the global differential d' maps TM to itself and restricts to a linear isometry
d'p W Tp M ! T'.p/ M for each p 2 M .
Given a point p 2 M , let Isop .M; g/ denote the isotropy subgroup at p, that
is, the subgroup of Iso.M; g/ consisting of isometries that fix p. For each ' 2
Isop .M; g/, the linear map d'p takes Tp M to itself, and the map Ip W Isop .M; g/ !
GL.Tp M / given by Ip .'/ D d'p is a representation of Isop .M; g/, called the
isotropy representation. We say that M is isotropic at p if the isotropy represen-
tation of Isop .M; g/ acts transitively on the set of unit vectors in Tp M . If M is
isotropic at every point, we say simply that M is isotropic.
There is an even stronger kind of symmetry than isotropy. Let O.M / denote the
set of all orthonormal bases for all tangent spaces of M :
a˚
O.M / D orthonormal bases for Tp M :
p2M
some of the isometry group, and in certain cases we will be able to prove that it is
the entire isometry group.
Euclidean Spaces
The simplest and most important model Riemannian manifold is of course
n-dimensional Euclidean space, which is just Rn with the Euclidean metric gx given
by (2.8).
Somewhat more generally, if V is any n-dimensional real vector space
endowed with an inner product, we can set g.v; w/ D hv; wi for any p 2 V and
any v; w 2 Tp V Š V . Choosing an orthonormal basis .b1 ; : : : ; bn / for V defines a
basis isomorphism from Rn to V that sends .x 1 ; : : : ; x n / to x i bi ; this is easily seen
to be an isometry of .V; g/ with R ; gx , so all n-dimensional inner product spaces
n
Problem 3-1 shows that when Rn is endowed n with the Euclidean metric, this action
is isometric and the induced action on O R is transitive. (Later, we will see that
this is in fact the full isometry group of R ; gx —see Problem 5-11—but we do not
n
Spheres
Our second class of model Riemannian manifolds comes in a family, with one for
each positive real number. Given R > 0, let Sn .R/ denote the sphere of radius R
centered at the origin in RnC1 , endowed with the metric gV R (called the round metric
of radius R) induced from the Euclidean metric on RnC1 . When R D 1, it is the
round metric on Sn introduced in Example 2.13, and we use the notation gV D gV 1 .
One of the first things one notices about the spheres is that like Euclidean spaces,
they are highly symmetric. We can immediately write down a large group of isome-
tries of Sn .R/ by observing that the linear action of the orthogonal group O.n C 1/
on RnC1 preserves Sn .R/ and the Euclidean metric, so its restriction to Sn .R/ acts
isometrically on the sphere. (Problem 5-11 will show that this is the full isometry
group.)
Proposition 3.2. The group O.n C 1/ acts transitively on O .Sn .R//, and thus each
round sphere is frame-homogeneous.
Proof. It suffices to show that given any p 2 Sn .R/ and any orthonormal ba-
sis .bi / for Tp Sn .R/, there is an orthogonal map that takes the “north pole”
N D .0; : : : ; 0; R/ to p and the basis .@1 ; : : : ; @n / for TN Sn .R/ to .bi /.
To do so, think of p as a vector of length R in RnC1 , and let py D p=R denote
the unit vector in the same direction (Fig. 3.1). Since the basis vectors .bi / are
tangent to the sphere, they are orthogonal to p, y is an orthonormal
y so .b1 ; : : : ; bn ; p/
Spheres 59
basis for RnC1 . Let ˛ be the matrix whose columns are these basis vectors. Then
˛ 2 O.n C 1/, and by elementary linear algebra, ˛ takes the standard basis vectors
y It follows that ˛.N / D p. Moreover, since ˛ acts
.@1 ; : : : ; @nC1 / to .b1 ; : : : ; bn ; p/.
linearly on RnC1 , its differential d˛N W TN RnC1 ! Tp RnC1 is represented in stan-
dard coordinates by the same matrix as ˛ itself, so d˛N .@i / D bi for i D 1; : : : ; n,
and ˛ is the desired orthogonal map. t
u
Another important feature of the round metrics—one that is much less evident
than their symmetry—is that they bear a certain close relationship to the Euclidean
metrics, which we now describe. Two metrics g1 and g2 on a manifold M are said
to be conformally related (or pointwise conformal or just conformal) to each other
if there is a positive function f 2 C 1 .M / such that g2 D fg1 . Given two Rie-
mannian manifolds .M; g/ and M ; gz , a diffeomorphism ' W M ! M is called a
conformal diffeomorphism (or a conformal transformation) if it pulls gz back to a
metric that is conformal to g:
Problem 3-6 shows that conformal diffeomorphisms are the same as angle-pre-
serving diffeomorphisms. Two Riemannian manifolds are said to be conformally
equivalent if there is a conformal diffeomorphism between them.
A Riemannian manifold .M; g/ is said to be locally conformally flat if every
point of M has a neighborhood that is conformally equivalent to an open set in
.Rn ; gx/.
I Exercise 3.3. (a) Show that for every smooth manifold M , conformality is an equiva-
lence relation on the set of all Riemannian metrics on M .
(b) Show that conformal equivalence is an equivalence relation on the class of all Rie-
mannian manifolds.
I Exercise 3.4. Suppose g1 and g2 D fg1 are conformally related metrics on an ori-
ented n-manifold. Show that their volume forms are related by d Vg2 D f n=2 d Vg1 .
R
.; / D u D : (3.4)
R
It follows from this formula that is defined and smooth on all of Sn .R/ X fN g.
The easiest way to see that it is a diffeomorphism is to compute its inverse. Solving
the two equations of (3.3) for and i gives
ui 1
i D ; DR : (3.5)
The point P D 1 .u/ is characterized by these equations and the fact that P is on
the sphere. Thus, substituting (3.5) into jj2 C 2 D R2 gives
juj2 2 . 1/
2
C R D R2 ;
2 2
from which we conclude
juj2 C R2
D :
2R2
Inserting this back into (3.5) gives the formula
2R2 u juj2 R2
1 .u/ D .; / D ; R ; (3.6)
juj2 C R2 juj2 C R2
In other words,
4R4
. 1 / gV R D gx; (3.7)
.juj2 C R2 /2
where gx now represents the Euclidean metric on Rn , and so is a conformal
diffeomorphism. t
u
Corollary 3.6. Each sphere with a round metric is locally conformally flat.
Hyperbolic Spaces
Our third class of model Riemannian manifolds is perhaps less familiar than the
other two. For each n 1 and each R > 0 we will define a frame-homogeneous
Riemannian manifold Hn .R/, called hyperbolic space of radius R. There are four
equivalent models of the hyperbolic spaces, each of which is useful in certain con-
texts. In the next theorem, we introduce all of them and show that they are isometric.
Theorem 3.7. Let n be an integer greater than 1. For each fixed R > 0, the following
Riemannian manifolds are all mutually isometric.
(a) (H YPERBOLOID MODEL ) Hn .R/ 1is the submanifold
of Minkowski space Rn;1
defined in standard coordinates ; : : : ; ; as the “upper sheet” f > 0g of
n
2 2
the two-sheeted hyperboloid 1 C C n 2 D R2 , with the induced
metric
1
gM R D qx;
where W Hn .R/ ! Rn;1 is inclusion, and qx D qx.n;1/ is the Minkowski metric:
2 2
qx D d 1 C C d n .d /2 : (3.8)
(b) (B ELTRAMI –K LEIN MODEL ) Kn .R/ is the ball of radius R centered at the
origin in Rn , with the metric given in coordinates .w 1 ; : : : ; w n / by
1
n 2
2 w dw C C w dw
1 n
2 .dw / C C .dw n /2
1 2
2
gM R DR CR : (3.9)
R2 jwj2 R2 jwj2 2
(c) (P OINCAR É BALL MODEL ) Bn .R/ is the ball of radius R centered at the
origin in Rn , with the metric given in coordinates .u1 ; : : : ; un / by
3 .du1 /2 C C .dun /2
gM R D 4R4 :
.R2 juj2 /2
Proof. Let R > 0 be given. We need to verify that Hn .R/ is actually a Riemannian
submanifold of Rn;1 , or in other words that gM R 1
is positive definite. One way to do
this is to show, as we will below, that it is the pullback of gM R
2
or gM R
3
(both of which
are manifestly positive definite) by a diffeomorphism. Alternatively, here is a direct
proof using some of the theory of submanifolds of pseudo-Riemannian manifolds
developed in Chapter 1.
Hyperbolic Spaces 63
Note that Hn .R/ is an open subset of a level set of the smooth function
2 2
f W Rn;1 ! R given by f .; / D 1 C C n 2 . We have
df D 2 1 d 1 C C 2 n d n 2 d ;
@ @ @
grad f D 2 1 C C 2 n n C 2 : (3.10)
@ 1 @ @
Direct computation shows that
X
i 2
qx.grad f; grad f / D 4 2 ;
i
which is equal to 4R2 at points of Hn .R/. Thus it follows from Corollary 2.71 that
Hn .R/ is a pseudo-Riemannian submanifold of signature .n; 0/, which is to say it
is Riemannian.
We will show that all four Riemannian manifolds are mutually isometric by defin-
ing isometries c W Hn .R/ ! Kn .R/, W Hn .R/ ! Bn .R/, and
W Bn .R/ ! U n .R/
(shown schematically in Fig. 3.3).
We begin with a geometric construction of a diffeomorphism called central pro-
jection from the hyperboloid to the ball,
c W Hn .R/ ! Kn .R/;
Fig. 3.4: Central projection from the hyperboloid to the Beltrami–Klein model
R
c.; / D : (3.11)
The relation jj2 2 D R2 guarantees that jc.; /j2 D R2 .1 R2 = 2 / < R2 , so
c maps Hn .R/ into Kn .R/. To show that c is a diffeomorphism, we determine its
inverse map. Let w 2 Kn .R/ be arbitrary. The unique positive scalar such that the
point .; / D .w; R/ lies on Hn .R/ is characterized by 2 jwj2 2 R2 D R2 ,
and therefore
R
D p :
R2 jwj2
It follows that the following smooth map is an inverse for c:
!
1 Rw R2
c .w/ D .; / D p ;p : (3.12)
R2 jwj2 R2 jwj2
we did for stereographic projection above. With .; / defined by (3.12), we have
P
i R dw i Rw i j w j dw j
d D p C 3=2 ;
R2 jwj2 R2 jwj2
P
R2 j w j dw j
d D 3=2 :
R2 jwj2
1 P
It is then straightforward to compute that c 1 gM R D i .d i /2 .d /2 D gM R
2
.
Hyperbolic Spaces 65
W Hn .R/ ! Bn .R/
from the hyperboloid to the ball, called hyperbolic stereographic projection, which
is an isometry between the metrics 1of (a) and (c). Let S 2 Rn;1 denote the point
S D .0; : : : ; 0; R/. For any P D ; : : : ; ; 2 H .R/ Rn;1 , set .P / D u 2
n n
Bn .R/, where U D .u; 0/ 2 Rn;1 is the point where the line through S and P inter-
sects the hyperplane f.; / W D 0g (Fig. 3.5). The point U is characterized by
.U S / D .P S / for some nonzero scalar , or
ui D i ;
(3.13)
R D . C R/:
These equations can be solved in the same manner as in the spherical case to yield
R
.; / D u D ;
RC
which takes its values in Bn .R/ because j .; /j2 D R2 . 2 R2 /. 2 C R2 / < R2 .
A computation similar to the ones before shows that the inverse map is
1 2R2 u R2 C juj2
.u/ D .; / D ;R 2 :
R2 juj2 R juj2
66 3 Model Riemannian Manifolds
so
is a diffeomorphism, called the generalized Cayley transform. The verification
that
gM R
3
D gM R
4
is basically a long calculation, and is left to Problem 3-4. t
u
We often use the generic notation Hn .R/ to refer to any one of the Riemannian
manifolds of Theorem 3.7, and gM R to refer to the corresponding metric; the special
case R D 1 is denoted by .Hn ; g/M and is called simply hyperbolic space, or in the
2-dimensional case, the hyperbolic plane.
Because all of the models for a given value of R are isometric to each other, when
analyzing them geometrically we can use whichever model is most convenient for
the application we have in mind. The next corollary is an example in which the
Poincaré ball and half-space models serve best.
Proof. In either the Poincaré ball model or the half-space model, the identity map
gives a global conformal equivalence with an open subset of Euclidean space. t
u
The examples presented so far might give the impression that most Riemannian
manifolds are locally conformally flat. This is far from the truth, but we do not yet
have the tools to prove it. See Problem 8-25 for some explicit examples of Riemann-
ian manifolds that are not locally conformally flat.
The symmetries of Hn .R/ are most easily seen in the hyperboloid model. Let
O.n; 1/ denote the group of linear maps from Rn;1 to itself that preserve the
Minkowski metric, called the .n C 1/-dimensional Lorentz group. Note that each
element of O.n; 1/ preserves the hyperboloid f 2 jj2 D R2 g, which has two com-
ponents determined by > 0 and < 0. We let OC .n; 1/ denote the subgroup of
O.n; 1/ consisting of maps that take the > 0 component of the hyperboloid to
itself. (This is called the orthochronous Lorentz group, because physically it repre-
sents coordinate changes that preserve the forward time direction.) Then OC .n; 1/
preserves Hn .R/, and because it preserves qx it acts isometrically on Hn .R/.
(Problem 5-11 will show that this is the full isometry group.) Recall that O Hn .R/
denotes the set of all orthonormal bases for all tangent spaces of Hn .R/.
Proposition 3.9. The group OC .n; 1/ acts transitively on O Hn .R/ , and therefore
Hn .R/ is frame-homogeneous.
qx D .ˇ 1 /2 C C .ˇ n /2 .ˇ nC1 /2 :
Thus the matrix whose columns are .b1 ; : : : ; bnC1 / is an element of OC .n; 1/ send-
ing N D .0; : : : ; 0; R/ to p and @i to bi (Fig. 3.6). t
u
both left- and right-invariant. The next lemma shows that left-invariant metrics are
easy to come by.
Lemma 3.10. Let G be a Lie group and let g be its Lie algebra of left-invariant
vector fields.
(a) A Riemannian metric g on G is left-invariant if and only if for all X; Y 2 g, the
function g.X; Y / is constant on G.
(b) The restriction map g 7! ge 2 †2 Te G together with the natural identifica-
tion Te G Š g gives a bijection between left-invariant Riemannian metrics on
G and inner products on g.
the group. Recall that this is the representation Ad W G ! GL.g/ given by Ad.'/ D
.C' / W g ! g, where C' W G ! G is the automorphism defined by conjugation:
C' . / D ' ' 1 . See Appendix C for more details.
Proposition 3.12. Let G be a Lie group and g its Lie algebra. Suppose g is a
left-invariant Riemannian metric on G, and let h; i denote the corresponding inner
product on g as in Lemma 3.10. Then g is bi-invariant if and only if h; i is invariant
under the action of Ad.G/ GL.g/, in the sense that hAd.'/X; Ad.'/Y i D hX; Y i
for all X; Y 2 g and ' 2 G.
Proof. We begin the proof with some preliminary computations. Suppose g is left-
invariant and h; i is the associated inner product on g. Let ' 2 G be arbitrary,
and note that C' is the composition of left multiplication by ' followed by right
multiplication by ' 1 . Thus for every X 2 g, left-invariance implies .R' 1 / X D
.R' 1 / .L' / X D .C' / X D Ad.'/X . Therefore, for all 2 G and X; Y 2 g, we
have
R' 1 g .X ; Y / D g ' 1 R' 1 X ' 1 ; R' 1 Y ' 1
D g ' 1 .Ad.'/X / ' 1 ; .Ad.'/Y / ' 1
D hAd.'/X; Ad.'/Y i:
Now assume that h; i is invariant under Ad.G/. Then the expression
on the last
line above is equal to hX; Y i D g .X ; Y /, which shows that R' 1 g D g. Since
this is true for all ' 2 G, it follows that g is bi-invariant.
Conversely, assuming that g is bi-invariant, we have R' 1 g D g for each
' 2 G, so the above computation yields
hX; Y i D g .X ; Y / D R' 1 g .X ; Y / D hAd.'/X; Ad.'/Y i;
Conversely, suppose H has compact closure in GL.V /, and let K denote the
closure. A simple limiting argument shows that K is itself a subgroup, and thus it is
a Lie group by the closed subgroup theorem (Thm. C.8). Let h; i0 be an arbitrary
inner product on V , and let be a right-invariant volume form on K (for example,
the volume form of some right-invariant metric on K). For fixed x; y 2 V , define
a smooth function fx;y W K ! R by fx;y .k/ D hkx; kyi0 . Then define a new inner
product h; i on V by Z
hx; yi D fx;y :
K
It follows directly from the definition that h; i is symmetric and bilinear over R.
For each nonzero x 2 V , we have fx;x > 0 everywhere on K, so hx; xi > 0, showing
that h; i is indeed an inner product.
To see that it is invariant under K, let k0 2 K be arbitrary. Then for all x; y 2 V
and k 2 K, we have
˝ ˛
fk0 x;k0 y .k/ D kk0 x; kk0 y 0
D fx;y ı Rk0 .k/;
Theorem 3.14 (Existence of Bi-invariant Metrics). Let G be a Lie group and g its
Lie algebra. Then G admits a bi-invariant metric if and only if Ad.G/ has compact
closure in GL.g/.
Proof. Proposition 3.12 shows that there is a bi-invariant metric on G if and only if
there is an Ad.G/-invariant inner product on g, and Lemma 3.13 in turn shows that
the latter is true if and only if Ad.G/ has compact closure in GL.g/. t
u
The most important application of the preceding theorem is to compact groups.
Corollary 3.15. Every compact Lie group admits a bi-invariant Riemannian metric.
Another important application is to prove that certain Lie groups do not admit
bi-invariant metrics. One way to do this is to note that if Ad.G/ has compact closure
in GL.g/, then every orbit of Ad.G/ must be a bounded subset of g with respect to
any choice of norm, because it is contained in the image of the compact set Ad.G/
under a continuous map of the form ' 7! '.X0 / from GL.g/ to g. Thus if one can
find an element X0 2 g and a subset S G such that the elements of the form
Ad.'/X are unbounded in g for ' 2 S , then there is no bi-invariant metric.
Here are some examples.
(a) Every left-invariant metric on an abelian Lie group is bi-invariant, because the
adjoint representation is trivial. Thus the Euclidean metric on Rn and the flat
metric on T n of Example 2.21 are both bi-invariant.
(b) If a metric g on a Lie group G is left-invariant, then the induced metric on
every Lie subgroup H G is easily seen to be left-invariant. Similarly, if g is
bi-invariant, then the induced metric on H is bi-invariant.
(c) The Lie group SL.2; R/ (the group of 2 2 real matrices of determinant 1)
admits many left-invariant metrics (as does every positive-dimensional Lie
group), but no bi-invariant ones. To see this, recall that the Lie algebra of
SL.2; R/ is isomorphic to the algebra sl.2; R/ of trace-free 2 2 matrices,
1
and the adjoint representation
0 1 is given by Ad.A/Xc D0
AXA (see Example
C.10). If we let X0 D 0 0 2 sl.2; R/ and Ac D 0 1=c 2 SL.2; R/ for c > 0,
then Ad.Ac /X0 D 00 0c 2 , which is unbounded as c ! 1. Thus the orbit of
X0 is not contained in any compact subset, which implies that there is no bi-
invariant metric on SL.2; R/. A similar argument shows that SL.n; R/ admits
no bi-invariant metric for any n 2. In view of (b) above, this shows also
that GL.n; R/ admits no bi-invariant metric for n 2. (Of course, GL.1; R/
does admit bi-invariant metrics because it is abelian.)
(d) With S3 regarded as a submanifold of C 2 , the map
w z
.w; z/ 7! (3.16)
x
z wx
gives a diffeomorphism from S3 to SU.2/. Under the inverse of this map, the
round metric on S3 pulls back to a bi-invariant metric on SU.2/, as Problem
3-10 shows.
(e) Let o.n/ denote the Lie algebra of O.n/, identified with the algebra of skew-
symmetric n n matrices, and define a bilinear form on o.n/ by
hA; Bi D tr AT B :
where x and y are treated as column matrices. These are the simplest examples
of nilpotent Lie groups, meaning that the series of subgroups G ŒG; G
ŒG; ŒG; G eventually reaches the trivial subgroup (where for any sub-
groups G1 ; G2 G, the notation ŒG1 ; G2 means the subgroup of G generated
by all elements of the form x1 x2 x11 x21 for xi 2 Gi ). There are many left-
invariant metrics on Hn , but no bi-invariant ones, as Problem 3-13 shows.
(h) Our last example is a group that plays an important role in the classification
of 3-manifolds. Let Sol denote the following 3-dimensional Lie subgroup of
GL.3; R/: ˚ ez 0 x
Sol D 0 e z y W x; y; z 2 R :
0 0 1
This group is the simplest nonnilpotent example of a solvable Lie group,
meaning that the series of subgroups G ŒG; G ŒŒG; G; ŒG; G even-
tually reaches the trivial subgroup. Like the Heisenberg groups, Sol admits
left-invariant metrics but not bi-invariant ones (Problem 3-14). //
In fact, John Milnor showed in 1976 [Mil76] that the only Lie groups that admit
bi-invariant metrics are those that are isomorphic to direct products of compact
groups and abelian groups.
Proof. Assume first that g is a G-invariant metric on M . Then the inner product gp0
on Tp0 M is invariant under the isotropy representation,
so it follows from Lemma
3.13 that Ip0 .Gp0 / has compact closure
in GL T p 0
M .
Conversely, assume that Ip0 Gp0 has compact closure in GL Tp0 M . Lemma
3.13 shows that there is an inner product gp0 on Tp0 .M / that is invariant under
the isotropy representation. For arbitrary p 2 M , we define an inner product gp on
Tp M by choosing an element ' 2 G such that '.p/ D p0 and setting
gp D d'p gp0 :
If '1 ; '2 are any two such elements of G, then '1 D h'2 with h D '1 '21 2 Gp0 , so
d'1 jp gp0 D d.h'2 /p gp0 D d'2 jp dhp0 gp0 D d'2 jp gp0 ;
The next corollary, which follows immediately from Theorem 3.17, addresses
the most commonly encountered case. (Other necessary and sufficient conditions
for the existence of invariant metrics are given in [Poo81, 6.58–6.59].)
Corollary 3.18. If a Lie group G acts smoothly and transitively on a smooth man-
ifold M with compact isotropy groups, then there exists a G-invariant Riemannian
metric on M . t
u
Sketch of proof. The proof relies on the topological classification of compact sur-
faces (see, for example, [LeeTM, Thms. 6.15 and 10.22]), which says that every
Other Homogeneous Riemannian Manifolds 75
Fig. 3.7: A connected sum of tori Fig. 3.8: Constructing a Riemannian covering
Fig. 3.9: The Klein bottle Fig. 3.10: Connected sum of three projective planes
It turns out that
acts freely and properly on R2 , and every
-orbit has a rep-
resentative in the unit square such that two points in the square are in the same
orbit if and only if they satisfy the equivalence relation generated by the relations in
(3.19). Thus the orbit space R2 =
is homeomorphic to the Klein bottle, and since
the group action is isometric, it follows that the Klein bottle inherits a flat, locally
frame-homogeneous metric and the quotient map is a Riemannian covering. Prob-
lem 3-18 asks you to work out the details.
A connected sum of n 3 copies of RP 2 : Such a surface is homeomorphic to a
quotient of a regular 2n-sided polygonal region by side identifications according to
a1 a1 a2 a2 : : : an an [LeeTM, Example 6.13]. As in the case of a connected sum of
tori, there is a compact region P B2 bounded by a 2n-sided regular geodesic poly-
gon whose interior angles are all =n (see Fig. 3.10), and there is a discrete group
of isometries that realizes the appropriate side identifications and yields a quotient
homeomorphic to the connected sum. The new ingredient here is that because such
a connected sum is not orientable, we must work with the full group of isometries
of B2 , not just the (orientation-preserving) ones determined by elements of G; but
Other Homogeneous Riemannian Manifolds 77
otherwise the argument is essentially the same as the one for connected sums of tori.
The details can be found in [Ive92, Section VII.1].
There is one remaining step. The arguments above show that each compact topo-
logical 2-manifold possesses a smooth structure and a locally frame-homogeneous
Riemannian metric, which admits a Riemannian covering by one of the three
frame-homogeneous model spaces. However, we started with a smooth compact
2-manifold, and we are looking for a Riemannian metric that is smooth with respect
to the given smooth structure. To complete the proof, we appeal to a result by James
Munkres [Mun56], which shows that any two smooth structures on a 2-manifold
are related by a diffeomorphism; thus after pulling back the metric by this diffeo-
morphism, we obtain a locally frame-homogeneous metric on M with its originally
given smooth structure. t
u
Locally homogeneous metrics also play a key role in the classification of compact
3-manifolds. In 1982, William Thurston made a conjecture about the classification
of such manifolds, now known as the Thurston geometrization conjecture. The
conjecture says that every compact, orientable 3-manifold can be expressed as a
connected sum of compact manifolds, each of which either admits a Riemannian
covering by a homogeneous Riemannian manifold or can be cut along embedded
tori so that each piece admits a finite-volume locally homogeneous Riemannian
metric. An important ingredient in the analysis leading up to the conjecture was his
classification of all simply connected homogeneous Riemannian 3-manifolds that
admit finite-volume Riemannian quotients. Thurston showed that there are exactly
eight such manifolds (see [Thu97] or [Sco83] for a proof):
R3 with the Euclidean metric
S3 with a round metric
H3 with a hyperbolic metric
S2 R with a product of a round metric and the Euclidean metric
H2 R with a product of a hyperbolic metric and the Euclidean metric
The Heisenberg group H1 of Example 3.16(g) with a left-invariant metric
The group Sol of Example 3.16(h) with a left-invariant metric
The universal covering group of SL.2; R/ with a left-invariant metric
The Thurston geometrization conjecture was proved in 2003 by Grigori Perel-
man. The proof is described in several books [BBBMP, KL08, MF10, MT14].
Symmetric Spaces
We end this section with a brief introduction to another class of Riemannian mani-
folds with abundant symmetry, called symmetric spaces. They turn out to be inter-
mediate between frame-homogeneous and homogeneous Riemannian manifolds.
Here is the definition. If .M; g/ is a Riemannian manifold and p 2 M , a
point reflection at p is an isometry ' W M ! M that fixes p and satisfies d'p D
Id W Tp M ! Tp M . A Riemannian manifold .M; g/ is called a (Riemannian) sym-
metric space if it is connected and for each p 2 M there exists a point reflection at
78 3 Model Riemannian Manifolds
Proof. Suppose .M; g/ satisfies the hypothesis, and let ' W M ! M be a point re-
flection at p 2 M . Given any other point q 2 M , by homogeneity there is an isometry
W M ! M satisfying .p/ D q. Then 'z D ı ' ı 1 is an isometry that fixes
q. Because d p is linear, it commutes with multiplication by 1, so
d 'zq D d p ı IdTpM ı d 1 q D IdTqM ı d p ı d 1 q
D IdTqM :
The definitions of the Euclidean, spherical, and hyperbolic metrics can easily be
adapted to give analogous classes of frame-homogeneous pseudo-Riemannian mani-
folds.
The first example is one we have already seen: the pseudo-Euclidean
space of
signature .r; s/ is the pseudo-Riemannian manifold Rr;s ; qx.r;s/ , where qx.r;s/ is the
pseudo-Riemannian metric defined by (2.24).
There are also pseudo-Riemannian analogues of the spherical and hyperbolic
metrics. For nonnegative integers r and s and a positive real number R, we define the
.r;s/ .r;s/
pseudosphere Sr;s .R/; qV R and the pseudohyperbolic space Hr;s .R/; qM R as
follows. As manifolds, Sr;s .R/ RrC1;s and Hr;s .R/ Rr;sC1 are defined by
n 2 2 2 2 o
Sr;s .R/ D .; / W 1 C C rC1 1 s D R2 ;
n 2 2 2 2 o
Hr;s .R/ D .; / W 1 C C r 1 sC1 D R2 :
The metrics are the ones induced from the respective pseudo-Euclidean metrics:
.r;s/ .r;s/
qV R D qx.rC1;s/ on Sr;s .R/, and qM R D qx.r;sC1/ on Hr;s .R/.
Theorem 3.25. For all r, s, and R as above, Sr;s .R/ and Hr;s .R/ are pseudo-
Riemannian manifolds of signature .r; s/.
Problems
3-1. Show that (3.2) defines a smooth isometric action of E.n/ on .Rn ; gx/, and
the induced action on O .Rn / is transitive. (Used on p. 57.)
3-2. Prove that the metric on RP n described in Example 2.34 is frame-homo-
geneous. (Used on p. 145)
3-3. Prove Proposition 3.1 (about homogeneous and isotropic Riemannian mani-
folds).
3-4. Complete the proof of Theorem 3.7 by showing that
gM R
3
D gM R
4
.
n n 2
3-5. (a) Prove that S .R/; gV R is isometric to S ; R gV for each R > 0.
(b) Prove that Hn .R/; gM R is isometric to Hn ; R2 gM for each R > 0.
(c) We could also have defined a family of metrics on Rn by gxR D R2 gx.
Why did we not bother?
(Used on p. 185.)
3-6. Show that two Riemannian metrics g1 and g2 are conformal if and only if
they define the same angles but not necessarily the same lengths, and that a
diffeomorphism is a conformal equivalence if and only if it preserves angles.
[Hint: Let .Ei / be a local orthonormal frame for g1 , and consider the g2 -angle
between Ei and .cos /Ei C .sin /Ej .] (Used on p. 59.)
3-7. Let U 2 denote the upper half-plane model of the hyperbolic plane (of radius
1), with the metric gM D .dx 2 C dy 2 /=y 2 . Let SL.2; R/ denote the group of
2 2 real matrices of determinant 1. Regard U 2 as a subset of the complex
plane with coordinate z D x C iy, and let
az C b ab
Az D ; AD 2 SL.2; R/:
cz C d cd
for the Lie algebra su.2/. For each positive real number a, define a left-
invariant metric ga on the group SU.2/ by declaring X; Y; aZ to be an
orthonormal frame.
(a) Show that ga is bi-invariant if and only if a D 1.
(b) Show that the map defined by (3.16) is an isometry between S3 ; gV
and .SU.2/; g1 /. [Remark: SU.2/ with any of these metrics is called a
Berger sphere, named after Marcel Berger.]
(Used on pp. 56, 71, 259.)
3-11. Prove that the formula hA; Bi D tr AT B defines a bi-invariant Riemannian
metric on O.n/. (See Example 3.16(e).)
3-12. Regard the upper half-space U n as a Lie group as described in Example
3.16(f).
(a) Show that for each R > 0, the hyperbolic metric gM R 4
on U n is left-
invariant.
(b) Show that U n does not admit any bi-invariant metrics.
(Used on pp. 68, 72.)
3-13. Write down an explicit formula for an arbitrary left-invariant metric on the
Heisenberg group Hn of Example 3.16(g) in terms of global coordinates
.x 1 ; : : : ; x n ; y 1 ; : : : ; y n ; z/, and show that the group has no bi-invariant met-
rics. (Used on pp. 68, 72.)
3-14. Repeat Problem 3-13 for the group Sol of Example 3.16(h). (Used on p. 72.)
3-15. Let Rn;1 be the .n C 1/-dimensional Minkowski space with coordinates
P 2
.; / D 1 ; : : : ; n ; and with the Minkowski metric qx.n;1/ D i d i
d 2 . Let S Rn;1 be the set
˚
S D .; / W . 1 /2 C C . n /2 D D 1 :
82 3 Model Riemannian Manifolds
Ie Ad
? ?
GL.Te G/ Š GL.g/;
where Ie is the isotropy representation of and g is the Lie algebra of G,
and use this to give an alternative proof of Theorem 3.14.
3-18. Let
E.2/ be the subgroup defined by (3.20). Prove that
acts freely and
properly on R2 and the orbit space is homeomorphic to the Klein bottle, and
conclude that the Klein bottle has a flat metric and a Riemannian covering by
the Euclidean plane.
3-19. Show that the Fubini–Study metric on CP n (Example 2.30) is homogeneous,
isotropic, and symmetric. (Used on p. 78.)
3-20. Show that the metric on the Grassmannian Gk .Rn / defined in Problem 2-7 is
homogeneous, isotropic, and symmetric. (Used on p. 78.)
3-21. Let M ; gz be a simply connected Riemannian manifold, and suppose
1
and
2 are countable subgroups of Iso M ; gz acting smoothly, freely, and
properly on M (when endowed with the discrete topology). For i D 1; 2, let
Mi D M =
i , and let gi be the Riemannian metric on Mi that makes the
quotient map i W M ! Mi a Riemannian covering (see Prop. 2.32). Prove
that the Riemannian manifolds .M1 ; g1 / and .M 2 ; g2 / are isometric if and
only if
1 and
2 are conjugate subgroups of Iso M ; gz .
3-22. Prove Theorem 3.25 (showing that pseudospheres and pseudohyperbolic
spaces are pseudo-Riemannian manifolds). [Hint: Mimic the argument in the
proof of Theorem 3.7 that Hn .R/ is Riemannian.]
Problems 83
R2
f .w/ D p :
R2 jwj2
Prove that the image of F is the anti-de Sitter space Hr;1 .R/, and F de-
.r;1/ r;1
fines a pseudo-Riemannian covering of Hr;1 .R/; qR by H z .R/; q .
[Remark: We are tacitly extending the notions of warped product metric and
Riemannian coverings to the pseudo-Riemannian case in the obvious ways.
It follows from the result of Problem 3-24 that Hr;s .R/ is simply
connected
when s 2 but Hr;1 .R/ is not. This shows that H z r;1 .R/; q , called uni-
versal anti-de Sitter space of radius R, is the universal pseudo-Riemannian
.r;1/
covering manifold of Hr;1 .R/; qR .]
Chapter 4
Connections
Our ultimate goal is to define a notion of curvature that makes sense on arbitrary Rie-
mannian manifolds, and to relate it to other geometric and topological properties.
Before we can do so, however, we need to study geodesics, the generalizations to
Riemannian manifolds of straight lines in Euclidean space. There are two key prop-
erties satisfied by straight lines in Rn , either of which serves to characterize them
uniquely: first, every segment of a straight line is the unique shortest path between its
endpoints; and second, straight lines are the only curves that have parametrizations
with zero acceleration.
The first of these characterizations—as shortest paths—is probably the most “ge-
ometric,” so it is tempting to try to use it as a definition of geodesics in Riemannian
manifolds. However, this property turns out to be technically difficult to work with
as a definition, so instead we will use “zero acceleration” as the defining property
and generalize that.
To make sense of acceleration on a manifold, we have to introduce a new object
called a connection—essentially a coordinate-independent set of rules for taking
directional derivatives of vector fields.
We begin this chapter by examining more closely the problem of finding an in-
variant interpretation for the acceleration of a curve, as a way to motivate the defini-
tions that follow. We then give a rather general definition of a connection, in terms
of directional derivatives of sections of vector bundles. After deriving some basic
properties of connections, we show how to use them to differentiate vector fields
along curves, to define geodesics, and to define “parallel transport” of vectors along
curves.
(Here and throughout the book, we use dots to denote ordinary derivatives with
respect to t when there are superscripts that would make primes hard to read.) A
curve in Rn is a straight line if and only if it has a parametrization for which
00 .t / 0.
We can also make sense of directional derivatives of vector fields on Rn , just by
computing ordinary directional derivatives
of
the component functions in standard
coordinates: given a vector field Y 2 X Rn and a vector v 2 Tp Rn , we define the
Euclidean directional derivative of Y in the direction v by the formula
ˇ ˇ
ˇ ˇ
rxv Y D v Y 1 @ ˇ C C v Y n @ ˇ ;
1
@x p ˇ @x ˇp
n
where for each i , v Y i is the result of applying the vector v to the function Y i :
@Y i .p/ i
n @Y .p/
v Y i D v1 C C v :
@x 1 @x n
x X Y by evaluating
If X is another vector field on Rn , we obtain a new vector field r
x Xp Y at each point:
r
xX Y D X Y 1 @ C C X Y n @ :
r (4.3)
@x 1 @x n
More generally, we can play the same game with curves and vector fields on a
submanifold of Rn . Suppose M Rn is an embedded submanifold, and consider
a smooth curve W I ! M . We want to think of a geodesic in M as a curve in M
that is “as straight as possible.” Of course, if M itself is curved, then 0 .t / (thought
of as a vector in Rn ) will probably have to vary, or else the curve will leave M .
But we can try to insist that the velocity not change any more than necessary for
the curve to stay in M . One way to do this is to compute the Euclidean acceleration
00 .t / as above, and then apply the tangential projection > W T.t/ Rn ! T.t/ M
(see Prop. 2.16). This yields a vector 00 .t /> D > . 00 .t // tangent to M , which we
call the tangential acceleration of . It is reasonable to say that is as straight as it
is possible for a curve in M to be if its tangential acceleration is zero.
Similarly, suppose Y is a smooth vector field on (an open subset of) M , and we
wish to ask how much Y is varying in M in the direction of a vector v 2 Tp M . Just
as in the case of velocity vectors, if we look at it from the point of view of Rn , the
vector field Y might be forced to vary just so that it can remain tangent to M . But
The Problem of Differentiating Vector Fields 87
one plausible way to answer the question is to extend Y to a smooth vector field Yz
on an open subset of Rn , compute the Euclidean directional derivative of Yz in the
direction v, and then project orthogonally onto Tp M . Let us define the tangential
directional derivative of Y in the direction v to be
rv> Y D > r x v Yz : (4.4)
Problem 4-1 shows that the tangential directional derivative is well defined and pre-
served by rigid motions of Rn . However, at this point there is little reason to believe
that the tangential directional derivative is an intrinsic invariant of M (one that de-
pends only on the Riemannian geometry of M with its induced metric).
On an abstract Riemannian manifold, for which there is no “ambient Euclidean
space” in which to differentiate, this technique is not available. Thus we have to
find some way to make sense of the acceleration of a smooth curve in an abstract
manifold. Let W I ! M be such a curve. As you know from your study of smooth
manifold theory, at each time t 2 I , the velocity of is a well-defined vector 0 .t / 2
T.t/ M (see Appendix A), whose representation in any coordinates is given by (4.1),
just as in Euclidean space.
However, unlike velocity, acceleration has no such coordinate-independent in-
terpretation. For example, consider the parametrized circle in the plane given in
Cartesian coordinates by .t / D .x.t /; y.t // D .cos t; sin t / (Fig. 4.1). As a smooth
curve in R2 , it has an acceleration vector at time t given by
ˇ ˇ
00 00 @ ˇˇ 00 @ ˇˇ
.t / D x .t / C y .t /
@x ˇ.t/ @x ˇ.t/
ˇ ˇ
@ ˇˇ @ ˇˇ
D cos t sin t :
@x ˇ .t/ @x ˇ.t/
88 4 Connections
But in polar coordinates, the same curve is described by .r.t /; .t // D .1; t / (Fig.
4.2). In these coordinates, if we try to compute the acceleration vector by the anal-
ogous formula, we get
ˇ ˇ
@ˇ @ ˇˇ
00 .t / D r 00 .t / ˇˇ C 00 .t / D 0:
@r .t/ @ ˇ.t/
Connections
see Examples 4.8 and 4.9.) After defining connections in this general setting, we
will adapt the definition to the case of vector fields along curves.
Let W E ! M be a smooth vector bundle over a smooth manifold M with
or without boundary, and let .E/ denote the space of smooth sections of E. A
connection in E is a map
rX .a1 Y1 C a2 Y2 / D a1 rX Y1 C a2 rX Y2 :
rX .f Y / D f rX Y C .Xf /Y:
The symbol r is read “del” or “nabla,” and rX Y is called the covariant derivative
of Y in the direction X . (This use of the word “covariant” has nothing to do with
covariant functors in category theory. It is related, albeit indirectly, to the use of
the word in the context of covariant and contravariant tensors, in that it reflects the
fact that the components of the covariant derivative have a transformation law that
“varies correctly” to give a well-defined meaning independent of coordinates. From
the modern coordinate-free point of view, “invariant derivative” would probably be
a better term.)
There is a variety of types of connections that are useful in different circum-
stances. The type of connection we have defined here is sometimes called a Koszul
connection to distinguish it from other types. Since we have no need to consider
other types of connections in this book, we refer to Koszul connections simply as
connections.
Although a connection is defined by its action on global sections, it follows from
the definitions that it is actually a local operator, as the next lemma shows.
Lemma 4.1 (Locality). Suppose r is a connection in a smooth vector bundle E !
M . For every X 2 X.M /, Y 2 .E/, and p 2 M , the covariant derivative rX Y jp
depends only on the values of X and Y in an arbitrarily small neighborhood of
p. More precisely, if X D Xz and Y D Yz on a neighborhood of p, then rX Y jp D
rXz Yz jp .
I Exercise 4.4. Complete the proof of the preceding proposition by showing that r U is
a connection.
In the situation of this proposition, we typically just refer to the restricted con-
nection as r instead of r U ; the proposition guarantees that there is no ambiguity in
doing so.
Lemma 4.1 tells us that we can compute the value of rX Y at p knowing only the
values of X and Y in a neighborhood of p. In fact, as the next proposition shows,
we need only know the value of X at p itself.
Proposition 4.5. Under the hypotheses of Lemma 4.1, rX Y jp depends only on the
values of Y in a neighborhood of p and the value of X at p.
Connections 91
Proof. The claim about Y was proved in Lemma 4.1. To prove the claim about X ,
it suffices by linearity to assume that Xp D 0 and show that rX Y jp D 0. Choose
a coordinate neighborhood U of p, and write X D X i @i in coordinates on U , with
X i .p/ D 0. Thanks to Proposition 4.3, it suffices to work with the restricted con-
nection on U , which we also denote by r. For every Y 2 .EjU /, we have
rX Y jp D rX i @i Y jp D X i .p/r@i Y jp D 0: t
u
Thanks to Propositions 4.3 and 4.5, we can make sense of the expression rv Y
when v is some element of Tp M and Y is a smooth local section of E defined only
on some neighborhood of p. To evaluate it, let X be a vector field on a neighbor-
hood of p whose value at p is v, and set rv Y D rX Y jp . Proposition 4.5 shows
that the result does not depend on the extension chosen. Henceforth, we will inter-
pret covariant derivatives of local sections of bundles in this way without further
comment.
Proposition 4.6. Let M be a smooth manifold with or without boundary, and let
r be a connection in TM . Suppose .Ei / is a smooth local frame over an open sub-
92 4 Connections
˚
set U M , and let ijk be the connection coefficients of r with respect to this
frame. For smooth vector fields X; Y 2 X.U /, written in terms of the frame as
X D X i Ei , Y D Y j Ej , one has
rX Y D X Y k C X i Y j ijk Ek : (4.9)
Existence of Connections
So far, we have studied properties of connections but have not produced any, so you
might be wondering whether they are plentiful or rare. In fact, they are quite plenti-
ful, as we will show shortly. Let us begin with the simplest example.
where > is the orthogonal projection onto TM (Prop. 2.16), r x is the Euclidean
z z
connection on R (Example 4.8), and X and Y are smooth extensions of X and
n
Proposition 4.12. The tangent bundle of every smooth manifold with or without
boundary admits a connection.
94 4 Connections
Proof. Let M be a smooth manifold with or without boundary, and cover M with
coordinate charts fU˛ g; the preceding lemma guarantees the existence of a connec-
tion r ˛ on each U˛ . Choose a partition of unity f'˛ g subordinate to fU˛ g. We would
like to patch the various r ˛ ’s together by the formula
X
˛
rX Y D '˛ rX Y: (4.11)
˛
Because the set of supports of the '˛ ’s is locally finite, the sum on the right-hand
side has only finitely many nonzero terms in a neighborhood of each point, so it
defines a smooth vector field on M . It is immediate from this definition that rX Y is
linear over R in Y and linear over C 1 .M / in X . We have to be a bit careful with the
product rule, though, since a linear combination of connections is not necessarily a
connection. (You can check, for example, that if r 0 and r 1 are connections, then
neither 2r 0 nor r 0 C r 1 satisfies the product rule.) By direct computation,
X
˛
rX .f Y / D '˛ rX .f Y /
˛
X
˛
D '˛ .Xf /Y C f rX Y
˛
X X
˛
D .Xf /Y '˛ C f '˛ rX Y
˛ ˛
D .Xf /Y C f rX Y: t
u
Although a connection is not a tensor field, the next proposition shows that the
difference between two connections is.
Then D is bilinear over C 1 .M /, and thus defines a .1; 2/-tensor field called the
difference tensor between r 0 and r 1 .
Proof. It is immediate from the definition that D is linear over C 1 .M / in its first
argument, because both r 0 and r 1 are. To show that it is linear over C 1 .M / in the
second argument, expand D.X; f Y / using the product rule, and note that the two
terms in which f is differentiated cancel each other.
Now that we know there is always one connection in TM , we can use the result
of the preceding proposition to say exactly how many there are.
Theorem 4.14. Let M be a smooth manifold with or without boundary, and let r 0
be any connection in TM . Then the set A.TM / of all connections in TM is equal
to the following affine space:
Covariant Derivatives of Tensor Fields 95
˚
A.TM / D r 0 C D W D 2 T .1;2/ TM ;
where D 2 T .1;2/ TM is interpreted as a map from X.M / X.M / to X.M / as
in Proposition B.1, and r 0 C D W X.M / X.M / ! X.M / is defined by
0 0
r C D X Y D rX Y C D.X; Y /:
Proposition 4.15. Let M be a smooth manifold with or without boundary, and let
r be a connection in TM . Then r uniquely determines a connection in each tensor
bundle T .k;l/ TM , also denoted by r, such that the following four conditions are
satisfied.
(i) In T .1;0/ TM D TM , r agrees with the given connection.
(ii) In T .0;0/ TM D M R, r is given by ordinary differentiation of functions:
rX f D Xf:
(iii) r obeys the following product rule with respect to tensor products:
rX .F ˝ G/ D .rX F / ˝ G C F ˝ .rX G/ :
(iv) r commutes with all contractions: if “tr” denotes a trace on any pair of in-
dices, one covariant and one contravariant, then
rX .tr F / D tr .rX F / :
.rX F / ! 1 ; : : : ; ! k ; Y1 ; : : : ; Yl D X F ! 1 ; : : : ; ! k ; Y1 ; : : : ; Yl
X
k
F ! 1 ; : : : ; rX ! i ; : : : ; ! k ; Y1 ; : : : ; Yl
iD1 (4.12)
X
l
F ! 1 ; : : : ; ! k ; Y1 ; : : : ; rX Yj ; : : : ; Yl :
j D1
Proof. First we show that every family of connections on all tensor bundles satisfying
(i)–(iv) also satisfies (a) and (b). Suppose we are given such a family of connections,
all denoted by r. To prove (a), note that h!; Y i D tr.! ˝ Y /, as can be seen by
evaluating both sides in coordinates, where they both reduce to !i Y i . Therefore,
(i)–(iv) imply
rX h!; Y i D rX tr.! ˝ Y /
D tr rX .! ˝ Y /
D tr rX ! ˝ Y C ! ˝ rX Y
D hrX !; Y i C h!; rX Y i :
where each trace operator acts on an upper index of F and the lower index of the
corresponding 1-form, or a lower index of F and the upper index of the correspond-
ing vector field.
Next we address uniqueness. Assume again that r represents a family of connec-
tions satisfying (i)–(iv), and hence also (a) and (b). Observe that (ii) and (a) imply
that the covariant derivative of every 1-form ! can be computed by
are both evident from (4.12) and (4.13), and the product rule in F follows easily
from the fact that differentiation of functions by X satisfies the product rule.
While (4.12) and (4.13) are useful for proving the existence and uniqueness of the
connections in tensor bundles, they are not very practical for computation, because
computing the value of rX F at a point requires extending all of its arguments to
vector fields and covector fields in an open set, and computing a great number of
derivatives. For computing the components of a covariant derivative in terms of a
local frame, the formulas in the following proposition are far more useful.
Proposition 4.16. Let M be a smooth manifold with or without boundary, and let
r be a connection in TM . Suppose .Ei / is a local frame for M , "j is its dual
˚
coframe, and ijk are the connection coefficients of r with respect to this frame.
Let X be a smooth vector field, and let X i Ei be its local expression in terms of this
frame.
(a) The covariant derivative of a 1-form ! D !i "i is given locally by
rX .!/ D X.!k / X j !i ji k "k :
(b) If F 2 T .k;l/ TM is a smooth mixed tensor field of any rank, expressed
locally as
i :::i
F D Fj11:::jkl Ei1 ˝ ˝ Eik ˝ "j1 ˝ ˝ "jl ;
then the covariant derivative of F is given locally by
Xk X
l
i :::i i :::p:::i i :::ik p
rX F D X Fj11:::jkl C X m Fj11:::jl k mp
is
X m Fj11:::p:::j
l mjs
sD1 sD1
rF W 1 .M / 1 .M / X.M / X.M / ! C 1 .M /
„ ƒ‚ … „ ƒ‚ …
k copies lC1 copies
given by
.rF / ! 1 ; : : : ; ! k ; Y1 ; : : : ; Yl ; X D .rX F / ! 1 ; : : : ; ! k ; Y1 ; : : : ; Yl (4.14)
98 4 Connections
defines a smooth .k; l C 1/-tensor field on M called the total covariant derivative
of F.
Proof. This follows immediately from the tensor characterization lemma (Lemma
B.6): rX F is a tensor field, so it is multilinear over C 1 .M / in its k C l arguments;
and it is linear over C 1 .M / in X by definition of a connection.
When we write the components of a total covariant derivative in terms of a local
frame, it is standard practice to use a semicolon to separate indices resulting from
differentiation from the preceding indices. Thus, for example, if Y is a vector field
written in coordinates as Y D Y i Ei , the components of the .1; 1/-tensor field rY
are written Y i Ij , so that
rY D Y i Ij Ei ˝ "j ;
with
Y i Ij D Ej Y i C Y k ji k :
For a 1-form !, the formulas read
More generally, the next lemma gives a formula for the components of total co-
variant derivatives of arbitrary tensor fields.
Proposition 4.18. Let M be a smooth manifold with or without boundary and let r
be a connection in TM ; and let .Ei / be a smooth local frame for TM and fijk g
the corresponding connection coefficients. The components of the total covariant
derivative of a .k; l/-tensor field F with respect to this frame are given by
X
k X
l
i :::i i1 :::ik i1 :::p:::ik is i :::ik
Fj11:::jkl Im D Em Fj1 :::jl C Fj1 :::jl mp Fj11:::p:::j p :
l mjs
sD1 sD1
I Exercise 4.20. Suppose F is a smooth .k; l/-tensor field and G is a smooth .r; s/-
tensor field. Show that the components of the total covariant derivative of F ˝ G are given
by
i1 :::i p1 :::pr i :::i i :::i p :::p
r.F ˝ G/ j1 :::jkl q1 :::qs Im D Fj11:::jkl Im Gqp11:::q
:::pr
s
C Fj11:::jkl Gq11:::qsrIm :
[Remark: This formula is often written in the following way, more suggestive of the product
rule for ordinary derivatives:
i :::i i :::i i :::i p :::p
Fj11:::jkl Gqp11:::q
:::pr
s
D Fj11:::jkl Im Gqp11:::q
:::pr
s
C Fj11:::jkl Gq11:::qsrIm :
Im
Notice that this does not say that r.F ˝ G/ D .rF / ˝ G C F ˝ .rG/, because in the
first term on the right-hand side of this latter formula, the index resulting from differentia-
tion is not the last lower index.]
Covariant Derivatives of Tensor Fields 99
Having defined the tensor field rF for a .k; l/-tensor field F , we can in turn take its
total covariant derivative and obtain a .k; l C 2/-tensor field r 2 F D r.rF /. Given
vector fields X; Y 2 X.M /, let us introduce the notation rX;Y2
F for the .k; l/-tensor
field obtained by inserting X; Y in the last two slots of r 2 F :
2
rX;Y F .: : : / D r 2 F .: : : ; Y; X /:
Note the reversal of order of X and Y : this is necessitated by our convention that
the last index position in rF is the one resulting from differentiation, while it is
conventional to let rX;Y
2
stand for differentiating first in the Y direction, then in the
X direction. (For this reason, some authors adopt the convention that the new index
position introduced by differentiation is the first instead of the last. As usual, be sure
to check each author’s conventions when you read.)
It is important to be aware that rX;Y
2
F is not the same as rX .rY F /. The main
reason is that the former is linear over C 1 .M / in Y , while the latter is not. The
relationship between the two expressions is given in the following proposition.
Proposition 4.21. Let M be a smooth manifold with or without boundary and let r
be a connection in TM . For every smooth vector field or tensor field F ,
2
rX;Y F D rX .rY F / r.rX Y / F:
2
rX;Y F D tr tr.r 2 F ˝ X / ˝ Y :
(First trace the last index of r 2 F with that of X , and then trace the last remaining
free index—originally the second-to-last in r 2 F —with that of Y .)
Therefore, since rX commutes with contraction and satisfies the product rule
with respect to tensor products (Prop. 4.15), we have
rX .rY F / D rX tr.rF ˝ Y /
D tr rX .rF ˝ Y /
D tr rX .rF / ˝ Y C rF ˝ rX Y
D tr tr.r 2 F ˝ X / ˝ Y C tr rF ˝ rX Y
2
D rX;Y F C r.rX Y / F: t
u
100 4 Connections
r 2 u.Y; X / D rX;Y
2
u D rX .rY u/ r.rX Y / u D Y.Xu/ .rY X /u:
Now we can address the question that originally motivated the definition of connec-
tions: How can we make sense of the derivative of a vector field along a curve?
Let M be a smooth manifold with or without boundary. Given a smooth curve
W I ! M , a vector field along is a continuous map V W I ! TM such that
V .t / 2 T.t/ M for every t 2 I ; it is a smooth vector field along if it is smooth as
a map from I to TM . We let X. / denote the set of all smooth vector fields along
. It is a real vector space under pointwise vector addition and multiplication by
constants, and it is a module over C 1 .I / with multiplication defined pointwise:
.f X /.t / D f .t /X.t /:
The most obvious example of a vector field along a smooth curve is the curve’s
velocity: 0 .t / 2 T.t/ M for each t , and its coordinate expression (4.1) shows that
it is smooth. Here is another example: if is a curve in R2 , let N.t / D R 0 .t /,
where R is counterclockwise rotationby =2, so N.t / is normal to 0 .t /. In standard
coordinates, N.t / D P .t /; P 1 .t / , so N is a smooth vector field along .
2
Fig. 4.4: Extendible vector field Fig. 4.5: Nonextendible vector field
Fig. 4.6: The image of the figure eight curve of Example 4.23
Its image is a set that looks like a figure eight in the plane (Fig. 4.6). Problem 4-7
asks you to show that is an injective smooth immersion, but its velocity vector
field is not extendible. //
Dt W X. / ! X. /;
102 4 Connections
Dt .f V / D f 0 V C f Dt V for f 2 C 1 .I /:
Dt V .t / D r 0 .t/ Vz :
There is an analogous operator on the space of smooth tensor fields of any type
along .
Proof. For simplicity, we prove the theorem for the case of vector fields along ;
the proof for arbitrary tensor fields is essentially identical except for notation.
First we show uniqueness. Suppose Dt is such an operator, and let t0 2 I be
arbitrary. An argument similar to that of Lemma 4.1 shows that the value of Dt V at
t0 depends only on the values of V in any interval .t0 "; t0 C "/ containing t0 . (If
t0 is an endpoint of I , extend a coordinate representation of to a slightly bigger
and then restrict back to I .)
open interval, prove the lemma there,
Choose smooth coordinates x i for M in a neighborhood of .t0 /, and write
ˇ
V .t / D V j .t /@j ˇ.t/
I Exercise 4.25. Complete the proof of Theorem 4.24 by showing that the operator Dt
defined in coordinates by (4.15) satisfies properties (i)–(iii).
Geodesics 103
Apart from its use in proving existence of the covariant derivative along a curve,
(4.15) also gives a practical formula for computing such covariant derivatives in
coordinates.
Now we can improve Proposition 4.5 by showing that rv Y actually depends only
on the values of Y along any curve through p whose velocity is v.
Proof. We can define a smooth vector field Z along by Z.t / D Y.t/ D Yz.t/ .
Since both Y and Yz are extensions of Z, it follows from condition (iii) in Theorem
4.24 that both rv Y and rv Yz are equal to Dt Z.t0 /.
Geodesics
Armed with the notion of covariant differentiation along curves, we can now define
acceleration and geodesics.
Let M be a smooth manifold with or without boundary and let r be a connection
in TM . For every smooth curve W I ! M , we define the acceleration of to be the
vector field Dt 0 along . A smooth curve is called a geodesic (with respect i to
r) if its acceleration is zero: Dt 0 0. In terms
of smooth coordinates
x , if
we write the component functions of as .t / D x 1 .t /; : : : ; x n .t / , then it follows
from (4.15) that is a geodesic if and only if its component functions satisfy the
following geodesic equation:
component functions satisfy (4.16). The standard trick for proving existence and
uniqueness for such a second-order system is to introduce auxiliary variables v i D
xP i to convert it to the following equivalent first-order system in twice the number of
variables:
xP k .t / D v k .t /;
(4.17)
vP k .t / D v i .t /v j .t /ijk .x.t //:
Treating x 1 ; : : : ; x n ; v 1 ; : : : ; v n / as coordinates on U Rn , we can recognize (4.17)
as the equations for the flow of the vector field G 2 X.U Rn / given by
ˇ ˇ
k @ ˇ
ˇ i j k @ ˇˇ
G.x;v/ D v v v ij .x/ k ˇ : (4.18)
@x k ˇ.x;v/ @v .x;v/
By the fundamental theorem on flows (Thm. A.42), for each .p; w/ 2 U Rn and
t0 2 R, there exist an open interval I0 containing t0 and a unique smooth solution
W I0 ! U Rn to this system satisfying the initial i condition .t0 / D .p; w/. If
we write the component functions 1 of as .t/ D x .t /; v i
.t / , then we can easily
check that the curve .t / D x .t /; : : : ; x .t / in U satisfies the existence claim of
n
the theorem.
To prove the uniqueness claim, suppose ; z W I ! M are both geodesics defined
on some open interval with .t0 / D z.t0 / and 0 .t0 / D z0 .t0 /. In any local coordi-
nates around .t0 /, we can define smooth curves ; z W .t0 "; t0 C "/ ! U Rn
as above. These curves both satisfy the same initial value problem for the sys-
tem (4.17), so by the uniqueness of ODE solutions, they agree on .t0 "; t0 C "/
for some " > 0. Suppose for the sake of contradiction that .b/ ¤ z.b/ for some
b 2 I . First suppose b > t0 , and let ˇ be the infimum of numbers b 2 I such that
b > t0 and .b/ ¤ z.b/ (Fig. 4.7). Then ˇ 2 I , and by continuity, .ˇ/ D z.ˇ/ and
0 .ˇ/ D z0 .ˇ/. Applying local uniqueness in a neighborhood of ˇ, we conclude that
and z agree on a neighborhood of ˇ, which contradicts our choice of ˇ. Arguing
similarly to the left of t0 , we conclude that z on all of I .
A geodesic W I ! M is said to be maximal if it cannot be extended to a
geodesic on a larger interval, that is, if there does not exist a geodesic z W Iz ! M
defined on an interval Iz properly containing I and satisfying zjI D . A geodesic
segment is a geodesic whose domain is a compact interval.
Parallel Transport 105
Proof. Given p 2 M and v 2 Tp M , let I be the union of all open intervals con-
taining 0 on which there is a geodesic with the given initial conditions. By Theo-
rem 4.27, all such geodesics agree where they overlap, so they define a geodesic
W I ! M , which is obviously the unique maximal geodesic with the given initial
conditions.
I Exercise 4.29. Show that the maximal geodesics on Rn with respect to the Euclidean
connection (4.3) are exactly the constant curves and the straight lines with constant-speed
parametrizations.
The unique maximal geodesic with .0/ D p and 0 .0/ D v is often called
simply the geodesic with initial point p and initial velocity v, and is denoted by v .
(For simplicity, we do not specify the initial point p in the notation; it can implicitly
be recovered from v by p D .v/, where W TM ! M is the natural projection.)
Parallel Transport
Another construction involving covariant differentiation along curves that will be
useful later is called parallel transport. As we did with geodesics, we restrict atten-
tion here to manifolds without boundary.
Let M be a smooth manifold and let r be a connection in TM . A smooth vector
or tensor field V along a smooth curve is said to be parallel along (with respect
to r) if Dt V 0 (Fig. 4.8). Thus a geodesic can be characterized as a curve whose
velocity vector field is parallel along the curve.
106 4 Connections
I Exercise 4.30. Let W I ! Rn be a smooth curve, and let V be a smooth vector field
along . Show that V is parallel along with respect to the Euclidean connection if and
only if its component functions (with respect to the standard basis) are constants.
The fundamental fact about parallel vector and tensor fields along curves is that
every tangent vector or tensor at any point on a curve can be uniquely extended to a
parallel field along the entire curve. Before we prove this claim, let us examine what
the equation of parallelism looks like in coordinates. Given a smooth curve with a
local coordinate representation .t / D 1 .t /; : : : ; n .t / , formula (4.15) shows that
a vector field V is parallel along if and only if
with analogous expressions based on Proposition 4.18 for tensor fields of other
types. In each case, this is a system of first-order linear ordinary differential equa-
tions for the unknown
coefficients of the vector or tensor field—in the vector case,
the functions V 1 .t /; : : : ; V n .t / . The usual ODE theorem guarantees the existence
and uniqueness of a solution for a short time, given any initial values at t D t0 ; but
since the equation is linear, we can actually show much more: there exists a unique
solution on the entire parameter interval.
Theorem 4.31 (Existence, Uniqueness, and Smoothness for Linear ODEs). Let
I R be an open interval, and for 1 j; k n, let Akj W I ! R be smooth func-
tions. For all t0 2 I and every initial vector c 1 ; : : : ; c n 2 Rn , the linear initial value
problem
VP k .t / D Akj .t /V j .t /;
(4.20)
V k .t0 / D c k ;
has a unique smooth solution on all of I , and the solution depends smoothly on
.t; c/ 2 I Rn .
Proof. First assume t0 D 0. Let x 0 ; x 1 ; : : : ; x n denote standard
coordinates
on the
manifold I Rn RnC1 , and consider the vector field Y 2 X I Rn defined by
@ @ @
Y D 0
C A1j x 0 x j 1 C C Anj x 0 x j n :
@x @x @x
1
If V .t / D V .t /; : : : ; V n .t / is a solution
to (4.20) with t0 D 0 defined on some
interval I0 I , then the curve .t / D t; V 1 .t /; : : : ; V n .t / is an integral curve of
Y defined on I0 satisfying the initial condition
.0/ D 0; c 1 ; : : : ; c n : (4.21)
1
Conversely, for each c ; : : : ; c n 2 Rn , there is an integral curve of Y defined
on some open interval I0 I containing 0 and satisfying (4.21).
If we write the
component functions of as .t / D 0 .t /; 1 .t /; : : : ; n .t/ , then P 0 .t / 1 and
0 .0/ D 0, so .t / D t for all t . It then follows that V .t / D 1 .t /; : : : ; n .t / solves
Parallel Transport 107
d X
jV .t /j2 D 2 VP k .t /V k .t /
dt
k
X
D2 Akj .t /V j .t /V k .t /
j;k
Now suppose .I / is not covered by a single chart. Let ˇ denote the supremum
of all b > t0 for which a unique parallel transport exists on Œt0 ; b
. (The argument
for t < t0 is similar.) We know that ˇ > t0 , since for b close enough to t0 , .Œt0 ; b
/
is contained in a single chart and the above argument applies. Then a unique par-
allel transport V exists on Œt0 ; ˇ/ (Fig. 4.9). If ˇ is equal to sup I , we are done. If
not, choose smooth coordinates on an open set containing .ˇ ı; ˇ C ı/ for some
positive ı. Then there exists a unique parallel vector field Vz on .ˇ ı; ˇ C ı/ sat-
isfying the initial condition Vz .ˇ ı=2/ D V .ˇ ı=2/. By uniqueness, V D Vz on
their common domain, and therefore Vz is a parallel extension of V past ˇ, which is
a contradiction.
The vector or tensor field whose existence and uniqueness are proved in Theorem
4.32 is called the parallel transport of v along . For each t0 ; t1 2 I , we define a map
ˇ P Y.h/ Yp
rX Y ˇp D lim h0 ; (4.25)
h!0 h
110 4 Connections
Pullback Connections
Like vector fields, connections in the tangent bundle cannot be either pushed for-
ward or pulled back by arbitrary smooth maps. However, there is a natural way to
pull back such connections by means of a diffeomorphism. In this section we define
this operation and enumerate some of its most important properties.
Suppose M and M are smooth manifolds and ' W M ! M is a diffeomorphism.
For a smooth vector field X 2 X.M /, recall that the pushforward of X is the
that satisfies d'p .Xp / D .' X /'.p/ for all p 2 M .
unique vector field ' X 2 X M
(See Lemma A.36.)
Lemma 4.37 (Pullback Connections). Suppose M and M are smooth manifolds
z is a connection in T M
with or without boundary. If r and ' W M ! M is a diffeo-
morphism, then the map ' rz W X.M / X.M / ! X.M / defined by
' r z Y D ' 1 r z ' X .' Y / (4.26)
X
Finally, to prove the product rule in Y , let f and fz be as above, and note that (A.7)
implies .' X / fz D .Xf / ı ' 1 . Thus
z .f Y / D ' 1 r
' r z ' X fz' Y
X
D ' 1 fzr z ' X .' Y / C .' X / fz ' Y
D f 'rz Y C .Xf /Y: t
u
X
The next proposition shows that various important concepts defined in terms of
connections—covariant derivatives along curves, parallel transport, and geodesics—
all behave as expected with respect to pullback connections.
Problems
4-1. Let M Rn be an embedded submanifold and Y 2 X.M /. For every point
p 2 M and vector v 2 Tp M , define rv> Y by (4.4).
(a) Show that rv> Y does not depend on the choice of extension Yz of Y .
[Hint: Use Prop. A.28.]
112 4 Connections
(b) Show that rv> Y is invariant under rigid motions of Rn , in the following
sense: if F 2 E.n/ and Mz D F .M /, then dFp rv> Y D r > .F Y /.
dFp .v/
.X; Y / D rX Y rY X ŒX; Y
:
(a) Show that is a .1; 2/-tensor field, called the torsion tensor of r .
(b) We say that r is symmetric if its torsion vanishes identically. Show that
r is symmetric if and only if its connection coefficients with respect
to every coordinate frame are symmetric: ijk D jki . [Warning: They
might not be symmetric with respect to other frames.]
(c) Show that r is symmetric if and only if the covariant Hessian r 2 u of
every smooth function u 2 C 1 .M / is a symmetric 2-tensor field. (See
Example 4.22.)
(d) Show that the Euclidean connection r x on Rn is symmetric.
d "j D "i ^ !i j C j ;
If we are to use geodesics and covariant derivatives as tools for studying Riemann-
ian geometry, it is evident that we need a way to single out a particular connection
on a Riemannian manifold that reflects the properties of the metric. In this chap-
ter, guided by the example of the tangential connection on a submanifold of Rn ,
we describe two properties that determine a unique connection on every Riemann-
ian manifold. The first property, compatibility with the metric, is easy to motivate
and understand. The second, symmetry, is a bit more mysterious; but it is motivated
by the fact that it is invariantly defined, and is always satisfied by the tangential
connection. It turns out that these two conditions are enough to determine a unique
connection associated with any Riemannian or pseudo-Riemannian metric, called the
Levi-Civita connection after the early twentieth-century Italian differential geometer
Tullio Levi-Civita.
After defining the Levi-Civita connection, we investigate the exponential map,
which conveniently encodes the collective behavior of geodesics and allows us to
study how they change as the initial point and initial velocity vary. Having estab-
lished the properties of this map, we introduce normal neighborhoods and normal
coordinates, which are essential computational and theoretical tools for studying
local geometric properties near a point. Then we introduce the analogous notion
for studying properties near a submanifold: tubular neighborhoods and Fermi
coordinates. Finally, we return to our three main model Riemannian manifolds and
determine their geodesics.
Except where noted otherwise, the results and proofs of this chapter do not use
positivity of the metric, so they apply equally well to Riemannian and pseudo-
Riemannian manifolds.
try. Since we get most of our intuition about Riemannian manifolds from studying
submanifolds of Rn with the induced metric, let us start by examining that case.
Let M Rn be an embedded submanifold. As a guiding principle, consider the
idea mentioned at the beginning of Chapter 4: a geodesic in M should be “as straight
as possible.” A reasonable way to make this rigorous is to require that the geodesic
have no acceleration in directions tangent to the manifold, or in other words that its
acceleration vector have zero orthogonal projection onto TM .
The tangential connection defined in Example 4.9 is perfectly suited to this task,
because it computes covariant derivatives on M by taking ordinary derivatives in
Rn and projecting them orthogonally to TM .
It is easy to compute covariant derivatives along curves in M with respect to
the tangential connection. Suppose W I ! M is a smooth curve. Then can be
regarded as either a smooth curve in M or a smooth curve in Rn , and a smooth
vector field V along that takes its values in TM can be regarded as either a
vector field along in M or a vector field along in Rn . Let D x t V denote the
n
covariant derivative of V along (as a curve in R ) with respect to the Euclidean
connection r, x and let D > V denote its covariant derivative along (as a curve in
t
M ) with respect to the tangential connection r > . The next proposition shows that
the two covariant derivatives along have a simple relationship to each other.
for some smooth functions V 1 ; : : : ; V k W .t0 "; t0 C "/ ! R. Formula (4.15) yields
X
k
ˇ ˇ
> Dx t V .t / D > VP i .t /Ei ˇ.t/ C V i .t /r
x 0 .t/ Ei ˇ
.t/
iD1
k
X ˇ ˇ
D VP i .t /Ei ˇ.t/ C V i .t / > rx 0 .t/ Ei ˇ
.t/
iD1
k
X ˇ ˇ
D VP i .t /Ei ˇ.t/ C V i .t /r>0 .t/ Ei ˇ.t/
iD1
D Dt> V .t /: t
u
Connections on Abstract Riemannian Manifolds 117
Proof. As noted in Example 4.8, the connection coefficients of the Euclidean con-
nection on Rn are all zero. Thus it follows from (4.15) that the Euclidean covariant
x t 0 .t / D 00 .t /. The corol-
derivative of 0 along is just its ordinary acceleration: D
lary then follows from Proposition 5.1. t
u
These
considerations
can be extended to pseudo-Riemannian manifolds as well.
Let Rr;s ; qx.r;s/ be the pseudo-Euclidean space of signature .r; s/. If M Rr;s
is an embedded Riemannian or pseudo-Riemannian submanifold, then for each
p 2 M , the tangent space Tp Rr;s decomposes as a direct sum Tp M ˚ Np M , where
Np M D .Tp M /? is the orthogonal complement of Tp M with respect to qx.r;s/ . We
let > W Tp Rr;s ! Tp M be the qx.r;s/ -orthogonal projection, and define the tangen-
tial connection r > on M by
rX>
Y D > r x z Yz ;
X
Metric Connections
The Euclidean connection on Rn has one very nice property with respect to the
Euclidean metric: it satisfies the product rule
˝ ˛ ˝ ˛
x X hY; Zi D r
r xX Z ;
x X Y; Z C Y; r
as you can verify easily by computing in terms of the standard basis. (In this for-
mula, the left-hand side represents the covariant derivative of the real-valued func-
tion hY; Zi regarded as a .0; 0/-tensor field, which is really just X hY; Zi by virtue
of property (ii) of Prop. 4.15.) The Euclidean connection has the same property with
respect to the pseudo-Euclidean metric on Rr;s . It is almost immediate that the tan-
gential connection on a Riemannian or pseudo-Riemannian submanifold satisfies
the same product rule, if we now interpret all the vector fields as being tangent to M
and interpret the inner products as being taken with respect to the induced metric on
M (see Prop. 5.8 below).
This property makes sense on an abstract Riemannian or pseudo-Riemannian
manifold. Let g be a Riemannian or pseudo-Riemannian metric on a smooth mani-
fold M (with or without boundary). A connection r on TM is said to be compatible
with g, or to be a metric connection, if it satisfies the following product rule for all
X; Y; Z 2 X.M /:
rX hY; Zi D hrX Y; Zi C hY; rX Zi : (5.1)
The next proposition gives several alternative characterizations of compatibility
with a metric, any one of which could be used as the definition.
d
hV; W i D hDt V ; W i C hV; Dt W i : (5.3)
dt
(e) If V; W are parallel vector fields along a smooth curve in M , then hV; W i is
constant along .
Connections on Abstract Riemannian Manifolds 119
(f) Given any smooth curve in M , every parallel transport map along is a
linear isometry.
(g) Given any smooth curve in M , every orthonormal basis at a point of can
be extended to a parallel orthonormal frame along (Fig. 5.1).
Proof. First we prove (a) , (b). By (4.14) and (4.12), the total covariant derivative
of the symmetric 2-tensor g is given by
This is zero for all X; Y; Z if and only if (5.1) is satisfied for all X; Y; Z.
To prove (b) , (c), note that Proposition 4.18 shows that the components of rg
in terms of a smooth local frame .Ei / are
l l
gij Ik D Ek .gij / ki glj kj gil :
d d i j
hV; W i D V W h@i ; @j i
dt dt
D VP i W j C V i WP j h@i ; @j i C V i W j hr 0 .t/ @i ; @j i C h@i ; r 0 .t/ @j i
D hDt V; W i C hV; Dt W i;
which proves (d). Conversely, if (d) holds, then in particular it holds for extendible
vector fields along , and then (a) follows from part (iii) of Theorem 4.24.
Now we will prove (d) ) (e) ) (f) ) (g) ) (d). Assume first that (d) holds. If
V and W are parallel along , then (5.3) shows that hV; W i has zero derivative with
respect to t , so it is constant along .
120 5 The Levi-Civita Connection
Symmetric Connections
This expression has the virtue that it is coordinate-independent and makes sense
for every connection on the tangent bundle. We say that a connection r on the
tangent bundle of a smooth manifold M is symmetric if
The symmetry condition can also be expressed in terms of the torsion tensor
of the connection, which was introduced in Problem 4-6; this is the smooth .1; 2/-
tensor field W X.M / X.M / ! X.M / defined by
.X; Y / D rX Y rY X ŒX; Y :
Using the symmetry condition on the last term in each line, this can be rewritten as
Adding the first two of these equations and subtracting the third, we obtain
Now suppose r 1 and r 2 are two connections on TM that are symmetric and
compatible with g. Since the right-hand side of (5.5) does not depend on the con-
Connections on Abstract Riemannian Manifolds 123
1 2
nection, it follows that hrX Y rX Y; Zi D 0 for all X; Y; Z. This can happen only
1 2
if rX Y D rX Y for all X and Y , so r 1 D r 2 .
To prove existence, we use (5.5), or rather a coordinate version of it. It suffices
to prove that such a connection exists in each coordinate chart, for then uniqueness
ensures that the
connections in different charts agree where they overlap.
Let U; x i be any smooth local coordinate chart. Applying (5.5) to the coordi-
nate vector fields, whose Lie brackets are zero, we obtain
˝ ˛ ˝ ˛ ˝ ˛
r@i @j ; @l D 12 @i @j ; @l C @j h@l ; @i i @l @i ; @j : (5.6)
Recall the definitions of the metric coefficients and the connection coefficients:
˝ ˛
gij D @i ; @j ; r@i @j D ijm @m :
Finally, multiplying both sides by the inverse matrix g kl and noting that gml g kl D
k
ım , we get
ijk D 12 g kl @i gj l C @j gil @l gij : (5.8)
This formula certainly defines a connection in each chart, and it is evident from
the formula that ijk D jki , so the connection is symmetric by Problem 4-6(b). Thus
only compatibility with the metric needs to be checked. Using (5.7) twice, we get
l l
ki glj C kj gil D 12 @k gij C @i gkj @j gki C 12 @k gj i C @j gki @i gkj
D @k gij :
(c) I N A LOCAL FRAME : Let .Ei / be a smooth local frame on an open subset
U M , and let cijk W U ! R be the n3 smooth functions defined by
Proof. We derived (5.9) and (5.10) in the proof of Theorem 5.10. To prove (5.12),
apply formula (5.9) with X D Ei , Y D Ej , and Z D El , to obtain
Multiplying both sides by g kl and simplifying yields (5.12). Finally, under the
hypotheses of (d), we have gij D ıij , so (5.12) reduces to (5.13) after rearranging
and using the fact that cijk is antisymmetric in i; j . t
u
On every Riemannian or pseudo-Riemannian manifold, we will always use the
Levi-Civita connection from now on without further comment. Geodesics with respect
to this connection are called Riemannian (or pseudo-Riemannian) geodesics, or
simply “geodesics” as long as there is no risk of confusion. The connection coeffi-
cients ijk of the Levi-Civita connection in coordinates, given by (5.10), are called
the Christoffel symbols of g.
The next proposition shows that these connections are familiar ones in the case
of embedded submanifolds of Euclidean or pseudo-Euclidean spaces.
Proposition 5.12.
(a) The Levi-Civita connection on a (pseudo-)Euclidean space is equal to the
Euclidean connection.
(b) Suppose M is an embedded (pseudo-)Riemannian submanifold of a (pseudo-)
Euclidean space. Then the Levi-Civita connection on M is equal to the tangen-
tial connection r > .
Proof. We observed earlier in this chapter that the Euclidean connection is sym-
metric and compatible with both the Euclidean metric gx and the pseudo-Euclidean
metrics qx.r;s/ , which implies (a). Part (b) then follows from Propositions 5.8 and
5.9. t
u
Connections on Abstract Riemannian Manifolds 125
Proposition
5.13 (Naturality of the Levi-Civita Connection). Suppose .M; g/
and M
; gz are Riemannian or pseudo-Riemannian manifolds with or without
boundary, and let r denote the Levi-Civita connection of g and r z that of gz. If
z
' W M ! M is an isometry, then ' r D r.
which shows that the pullback connection is compatible with g. Symmetry is proved
as follows:
z Y 'r
' r z X D ' 1 r z ' X .' Y / r
z ' Y .' X /
X Y
D ' 1 ' X; ' Y
D ŒX; Y : t
u
Corollary 5.14 (Naturality of Geodesics). Suppose .M; g/ and M
; gz are Rie-
mannian or pseudo-Riemannian manifolds with or without boundary, and ' W M !
Proof. This is an immediate consequence of Proposition 4.38, together with the fact
that being a geodesic is a local property. t
u
Like every connection on the tangent bundle, the Levi-Civita connection induces
connections on all tensor bundles.
Proof. Since every tensor field can be written as a sum of tensor products of vector
and/or covector fields, it suffices to consider the case in which F D ˛1 ˝ ˝ ˛kCl
and G D ˇ1 ˝ ˝ ˇkCl , where ˛i and ˇi are covariant or contravariant 1-tensor
fields, as appropriate. In this case, the formula follows from (2.15) by a routine
computation. t
u
Proposition 5.17. The musical isomorphisms commute with the total covariant
derivative operator: if F is any smooth tensor field with a contravariant i th index
position, and [ represents the operation of lowering the i th index, then
r F [ D .rF /[ : (5.14)
Similarly, if G has a covariant i th position and ] denotes raising the i th index, then
r G ] D .rG/] : (5.15)
Proof. The discussion on page 27 shows that F [ D tr.F ˝ g/, where the trace is
taken on the i th and last indices of F ˝ g. Because g is parallel, for every vector
field X we have rX .F ˝ g/ D .rX F / ˝ g. Because rX commutes with traces,
therefore,
rX F [ D rX tr.F ˝ g/ D tr .rX F / ˝ g D .rX F /[ :
This shows that when X is inserted into the last index position on both sides of
(5.14), the results are equal. Since X is arbitrary, this proves (5.14).
Because the sharp and flat operators are inverses of each other when applied
to the same index position, (5.15) follows by substituting F D G ] into (5.14) and
applying ] to both sides. t
u
dependence of geodesics on the initial data is encoded in a map from the tangent
bundle into the manifold, called the exponential map, whose properties are funda-
mental to the further study of Riemannian geometry.
(It is worth noting that the existence of the exponential map and the basic prop-
erties expressed in Proposition 5.19 below hold for every connection in TM , not just
for the Levi-Civita connection. For simplicity, we restrict attention here to the latter
case, because that is all we need. We also restrict to manifolds without boundary, in
order to avoid complications with geodesics running into a boundary.)
The next lemma shows that geodesics with proportional initial velocities are
related in a simple way.
Proof. If c D 0, then both sides of (5.16) are equal to p for all t 2 R, so we may
assume that c ¤ 0. It suffices to show that cv .t / exists and (5.16) holds whenever
the right-hand side is defined. (The same argument with the substitutions v D c 0 v 0 ,
t D c 0 t 0 , and c D 1=c 0 then implies that the conclusion holds when only the left-hand
side is known to be defined.)
Suppose the maximal domain of v is the open interval I R. For simplicity,
write D v , and define a new curve z W c 1 I ! M by z.t/ D .ct /, where c 1 I D
fc 1 t W t 2 I g. We will show that z is a geodesic with initial point p and initial
velocity cv; it then follows by uniqueness and maximality that it must be equal to
cv .
It is immediate from the definition that z.0/ D .0/ D p. Choose any smooth
local
1 coordinates on M and write the coordinate representation of as .t / D
.t /; : : : ; n .t / ; then the chain rule gives
d
zP i .t / D i .ct /
dt
D c P i .ct /:
exp.v/ D v .1/:
v .t / D exp.t v/ (5.17)
(Here and whenever convenient, we use the notations .p; v/ and v interchangeably
for an element v 2 Tp M , depending on whether we wish to emphasize the point at
which v is tangent.) Since this formula is independent of coordinates, it shows that
the various definitions of G given by (4.18) in different coordinate systems agree.
To prove that G satisfies (5.18), we write the components of the geodesic v .t /
as x i .t / and those of its velocity as v i .t / D xP i .t /. Using the chain rule and the
geodesic equation in the form (4.17), we can write the right-hand side of (5.18) as
ˇ
@f @f ˇ
k
k
.x.t /; v.t //xP .t / C k
.x.t /; v.t //vP .t / ˇˇ
k
@x @v tD0
@f @f
D k .p; v/v k k .p; v/v i v j ijk .p/
@x @v
D Gf .p; v/:
The fundamental theorem on flows (Thm. A.42) shows that there exist an open
set D R TM containing f0g TM and a smooth map W D ! TM , such that
each curve .p;v/ .t / D .t; .p; v// is the unique maximal integral curve of G starting
at .p; v/, defined on an open interval containing 0.
Now suppose .p; v/ 2 E. This means that the geodesic v is defined at least on
the interval Œ0; 1, and therefore so is the integral curve of G starting at .p; v/ 2 TM .
Since .1; .p; v// 2 D, there is a neighborhood of .1; .p; v// in R TM on which the
flow of G is defined (Fig. 5.2). In particular, this means that there is a neighborhood of
.p; v/ on which the flow exists for t 2 Œ0; 1, and therefore on which the exponential
map is defined. This shows that E is open.
130 5 The Levi-Civita Connection
Since geodesics are projections of integral curves of G, it follows that the expo-
nential map can be expressed as
wherever it is defined, and therefore expp .v/ is a smooth function of .p; v/.
To compute d expp 0 .v/ for an arbitrary vector v 2 Tp M , we just need to
choose a curve in Tp M starting at 0 whose initial velocity is v, and compute
the initial velocity of expp ı . A convenient curve is .t / D t v, which yields
ˇ ˇ ˇ
d ˇˇ d ˇˇ d ˇˇ
d expp 0 .v/ D .expp ı /.t / D expp .t v/ D v .t / D v:
dt ˇtD0 dt ˇtD0 dt ˇtD0
Thus d expp 0 is the identity map. t
u
Corollary 5.14 on the naturality of geodesics translates into the following impor-
tant property of the exponential map.
Proposition
5.20 (Naturality of the Exponential Map). Suppose .M; g/ and
M
; gz are Riemannian or pseudo-Riemannian manifolds and ' W M ! M
is a
local isometry. Then for every p 2 M , the following diagram commutes:
d'p- z
Ep E'.p/
expp exp'.p/
? ?
' -
M M;
Normal Neighborhoods and Normal Coordinates 131
B 1
- Rn
Tp M
-
6 '
.expp jV /1
U:
132 5 The Levi-Civita Connection
Proof. Let ' be a normal coordinate chart on U centered at p, with coordinate func-
tions x i . By definition, this means that ' D B 1 ıexp1
p , where B W R n
! T p M is
the basis isomorphism
determined by some orthonormal
basis .b i / for T p M . Note
that d'p1 D d expp 0 ıdB0 D B because d expp 0 is the identity and B is linear.
Thus @i jp D d'p1 .@i j0 / D B.@i j0 / D bi , which shows that the coordinate basis is
orthonormal at p. Conversely, every orthonormal basis .bi / for Tp M yields a basis
isomorphism B and thus a normal coordinate chart ' D B 1 ıexp1 p , which satisfies
@i jp D bi by the computation above.
If 'z D Bz 1 ı exp1
p is another such chart, then
Proof. Part (a) follows directly from the definition of normal coordinates, and parts
(b) and (c) follow from Propositions 5.23 and 5.19(b), respectively.
To prove (d), let v D v i @i jp 2 Tp M be arbitrary. The geodesic equation (4.16)
for v .t / D t v 1 ; : : : ; t v n simplifies to
ijk .t v/v i v j D 0:
Evaluating this expression at t D 0 shows that ijk .0/v i v j D 0 for every index k and
k
every vector v. In particular, with v D @a for some fixed a, this shows that aa D0
for each a and k (no summation). Substituting v D @a C @b and v D @b @a for any
k
fixed pair of indices a and b and subtracting, we conclude also that ab D 0 at p for
all a; b; k. Finally, (e) follows from (d) together with (5.2) in the case Ek D @k . u t
Because they are given by the simple formula (5.20), the geodesics starting at p
and lying in a normal neighborhood of p are called radial geodesics. (But be warned
that geodesics that do not pass through p do not in general have a simple form in
normal coordinates.)
for some ı > 0. (Here we are using the fact that P is embedded in M , so it has the
subspace topology.)
To complete the proof, we need to show that there is a set V EP of the form
(5.21) on which E is a diffeomorphism onto its image. For each point x 2 P , define
˚
.x/ D sup ı 1 W E is a diffeomorphism from Vı .x/ to its image : (5.23)
The argument in the preceding paragraph implies that .x/ is positive for each
x. Note that E is injective on the entire set V.x/ .x/, because any two points
.x1 ; v1 /; .x2 ; v2 / in this set are in Vı .x/ for some ı < .x/. Because it is an injective
local diffeomorphism, E is actually a diffeomorphism from V.x/ .x/ onto its image.
Next we show that the function W P ! R is continuous. For any x; x 0 2 P ,
if dg .x; x 0 / < .x/, then the triangle inequality shows that Vı .x 0 / is contained in
V.x/ .x/ for ı D .x/ dg .x; x 0 /, which implies that .x 0 / .x/ dg .x; x 0 /,
or
.x/ .x 0 / dg .x; x 0 /: (5.24)
Tubular Neighborhoods and Fermi Coordinates 135
If dg .x; x 0 / .x/, then (5.24) holds for trivial reasons. Reversing the roles of x
and x 0 yields an analogous inequality, which shows that j.x/ .x 0 /j dg .x; x 0 /,
so is continuous.
˚
Let V D .x; v/ 2 NP W jvjg < 12 .x/ , which is an open subset of NP contain-
ing P0 . We show that E is injective on V . Suppose .x; v/ and .x 0 ; v 0 / are points in
V such that E.x; v/ D E.x 0 ; v 0 / (Fig. 5.4). Assume without loss of generality that
.x 0 / .x/. Because expx .v/ D expx 0 .v 0 /, there is an admissible curve from x
to x 0 of length jvjg C jv 0 jg , and thus
Therefore, both .x; v/ and .x 0 ; v 0 / are in V.x/ .x/. Since E is injective on this set,
this implies .x; v/ D .x 0 ; v 0 /.
The set U D E.V / is open in M because EjV is a local diffeomorphism and
thus an open map, and E W V ! U is a diffeomorphism. Therefore, U is a tubular
neighborhood of P .
Finally, if P is compact, then the continuous function 12 achieves a minimum
value " > 0 on P , so U contains a uniform tubular neighborhood of radius ". t
u
Fermi Coordinates
B 1
- W 0 Rnp Rn
NP jW0
-
1 6 '
EjV0
U0 :
136 5 The Levi-Civita Connection
Coordinates of this form are called Fermi coordinates, after the Italian physicist
Enrico Fermi (1901–1954), who first introduced them in the special case in which
P is the image of a geodesic in M . The generalization to arbitrary submanifolds
was first introduced by Alfred Gray [Gra82]. (See also [Gra04] for a detailed study
of the geometry of tubular neighborhoods.)
Here is the analogue of Proposition 5.24 for Fermi coordinates.
Euclidean Space
On Rn with the Euclidean metric, Proposition 5.12 shows that the Levi-Civita
connection is the Euclidean connection. Therefore, as one would expect, constant-
coefficient vector fields are parallel, and the Euclidean geodesics are straight lines
with constant-speed parametrizations (Exercises 4.29 and 4.30). Every Euclidean
space is geodesically complete.
Spheres
Because the round metric on the sphere Sn .R/ is induced by the Euclidean metric on
RnC1 , it is easy to determine the geodesics on a sphere using Corollary 5.2. Define
a great circle on Sn .R/ to be any subset of the form Sn .R/ \ ˘ , where ˘ RnC1
is a 2-dimensional linear subspace.
Proof. Let p 2 Sn .R/ be arbitrary. Because f .x/ D jxj2 is a defining function for
Sn .R/, a vector v 2 Tp RnC1 is tangent to Sn .R/ if and only if dfp .v/ D 2hv; pi D
0, where we think of p as a vector by means of the usual identification of RnC1 with
Tp RnC1 . Thus Tp Sn .R/ is exactly the set of vectors orthogonal to p.
Suppose v is an arbitrary nonzero vector in Tp Sn .R/. Let a D jvj=R and vy D v=a
v j D R), and consider the smooth curve W R ! RnC1 given by
(so jy
0 .t / D a.sin at /p C a.cos at /y
v;
00 .t / D a2 .cos at /p a2 .sin at /y
v:
Hyperbolic Spaces
Before considering the other three models, note that since maximal geodesics in
Hn .R/ are constant-speed embeddings of R, it follows from naturality that maximal
geodesics in each of the other models are also constant-speed embeddings of R.
Thus each model is geodesically complete, and to determine the geodesics in the
other models we need only determine their images.
Consider the Beltrami–Klein model. Recall the isometry c W Hn .R/ ! Kn .R/
given by c.; / D R= (see (3.11)). The image of a maximal geodesic in Hn .R/
is a great hyperbola, which is the set of points .; / 2 Hn .R/ that solve a system
of n 1 independent linear equations. Simple algebra shows that .; / satisfies a
linear equation ˛i i C ˇ D 0 if and only if w D c.; / D R= satisfies the affine
equation ˛i w i D ˇR. Thus c maps each great hyperbola onto the intersection of
Kn .R/ with an affine subspace of Rn , and since it is the image of a smooth curve,
it must be the intersection of Kn .R/ with a straight line.
Next consider the Poincaré ball model. First consider the 2-dimensional case,
and recall the inverse hyperbolic stereographic projection 1 W B2 .R/ ! H2 .R/
constructed in Chapter 3:
2R2 u R2 C juj2
1 .u/ D .; / D ; R :
R2 juj2 R2 juj2
In this case, a great hyperbola is the set of points on H2 .R/ that satisfy a single
linear equation ˛i i C ˇ D 0. In the special case ˇ D 0, this hyperbola is mapped
by to a straight line segment through the origin, as can easily be seen from the
geometric definition of . If ˇ ¤ 0, we can assume (after multiplying through by a
constant if necessary) that ˇ D 1, and write the linear equation as D ˛i i D ˛
(where the dot represents the Euclidean dot product between elements of R2 ). Under
1 , this pulls back to the equation
R2 C juj2 2R2 ˛ u
R 2 2
D 2
R juj R juj2
on the disk, which simplifies to
juj2 2R˛ u C R2 D 0:
If j˛j2 1, this locus is either empty or a point on @B2 .R/, so it contains no points
in B2 .R/. Since we are assuming that it is the image of a maximal geodesic, we must
therefore havepj˛j2 > 1. In that case, (5.26) is the equation of a circle with center R˛
and radius R j˛j2 1. At a point u0 where the circle intersects @B2 .R/, the three
points 0, u0 , and R˛ form a triangle with sides ju0 j D R, jR˛j, and ju0 R˛j (Fig.
5.9), which satisfy the Pythagorean identity by (5.26); therefore the circle meets
@B2 .R/ in a right angle.
Geodesics of the Model Spaces 141
Fig. 5.9: Geodesics are arcs of circles orthogonal to the boundary of H2 .R/
jz iRj2 z iR zN C iR
R2 2
iR2 ˛N C iR2 ˛ C R2 j˛j2 D R2 .j˛j2 1/:
jz C iRj z C iR zN iR
This is the equation of a circle with center on the x-axis, unless b D 1, in which
case the condition j˛j2 > 1 forces a ¤ 0, and then it is a straight line x D constant.
The other class of geodesics on the ball, line segments through the origin, can be
handled similarly.
In the higher-dimensional case, suppose first that W R ! U n .R/ ˚ is a maximal
geodesic such that .0/ lies on the y-axis and 0 .0/ is in the span of @=@x 1 ; @=@y .
From the explicit formula (3.15) for
, it˚follows that
ı .0/ lies on the v-axis in
the ball, and .
ı /0 .0/ is in the span of @=@u1 ; @=@v . The image of the geodesic
ı is either part of a line through the origin or an arc of a circle perpendicu-
lar to @Bn .R/, both of which are contained in the u1 ; v -plane. By the argument
142 5 The Levi-Civita Connection
in 1the preceding paragraph, it then follows that the image of is contained in the
x ; y -plane and is either a vertical half-line or a semicircle centered on the y D 0
hyperplane. For the general case, note that translations and orthogonal transforma-
tions in the x-variables preserve vertical half-lines and circles centered on the y D 0
3
hyperplane in U n .R/, and they also preserve the metric gM R . Given an arbitrary max-
n
imal geodesic W R ! U .R/, after applying an x-translation we may assume that
.0/ lies on the y-axis, and after an orthogonal
˚ transformation
in the x variables,
we may assume that 0 .0/ is in the span of @=@x 1 ; @=@y ; then the argument above
shows that the image of is either a vertical half-line or a semicircle centered on
the y D 0 hyperplane. t
u
Problems
gj k !i k C gik !j k D dgij :
Show that
this
implies that with respect to every local orthonormal frame, the
matrix !i j is skew-symmetric.
5-3. Define a connection on R3 by setting (in standard coordinates)
3 1 2
12 D 23 D 31 D 1;
3 1 2
21 D 32 D 13 D 1;
with all other connection coefficients equal to zero. Show that this connec-
tion is compatible with the Euclidean metric and has the same geodesics as
the Euclidean connection, but is not symmetric. (See Problem 4-9.)
5-4. Let C be an embedded smooth curve in the half-plane H D f.r; z/ W r > 0g,
and let SC R3 be the surface of revolution determined by C as in Example
2.20. Let .t / D .a.t /; b.t // be a unit-speed local parametrization of C , and
let X be the parametrization of SC given by (2.11).
(a) Compute the Christoffel symbols of the induced metric on SC in .t; /
coordinates.
146 5 The Levi-Civita Connection
(a) Show that for every pair of vector fields X; Y 2 X.M /, we have
˝ ˛
z Yz D hX; Y i ı I
X;
A
z Yz H D ŒX; Y I
X;
z W is vertical if W 2 X M
X;
is vertical.
A
z Xz Yz D rX Y C 1 Xz ; Yz V :
r 2
(5.27)
z be a horizontal lift and W a vertical vector field on M
[Hint: Let Z
,
˝ ˛ ˝ ˛
and compute r z Xz Yz ; Z
z and rz Xz Yz ; W using (5.9).]
(Used on p. 224.)
5-7. Suppose .M1 ; g1 / and .M2 ; g2 / are Riemannian manifolds.
(a) Prove that if M1 M2 is endowed with the product metric, then a curve
W I ! M1 M2 of the form .t / D .1 .t /; 2 .t // is a geodesic if and
only if i is a geodesic in .Mi ; gi / for i D 1; 2.
(b) Now suppose f W M1 ! RC is a strictly positive smooth function, and
M1 f M2 is the resulting warped product (see Example 2.24). Let
1 W I ! M1 be a smooth curve and q0 a point in M2 , and define
Problems 147
x i x j ijk .x/ 0:
5-10. Prove Proposition 5.22 (a local isometry is determined by its value and dif-
ferential at one point).
5-11. Recall the groups E.n/, O.n C 1/, and OC .n; 1/ defined
in Chapter
n 3, which
n
act isometrically
n on the model Riemannian manifolds R ; x
g , S .R/; gV R ,
and H .R/; gM R , respectively.
(a) Show that
148 5 The Levi-Civita Connection
Iso Rn ; gx D E.n/;
Iso Sn .R/; gV R D O.n C 1/;
Iso Hn .R/; gM R D OC .n; 1/:
(b) Show that in each case, for each point p in Rn , Sn .R/, or Hn .R/, the
isotropy group at p is a subgroup isomorphic to O.n/.
(c) Strengthen the result above
by showing
that if .M;
g/ is one of the
Riemannian manifolds Rn ; gx , Sn .R/; gV R , or Hn .R/; gM R , U is
a connected open subset of M , and ' W U ! M is a local isometry,
then ' is the restriction to U of an element of Iso.M; g/.
(Used on pp. 57, 58, 67, 348, 349.)
5-12. Suppose M is a connected n-dimensional Riemannian manifold, and G is
a Lie group acting isometrically and effectively on M . Show that dim G
n.n C 1/=2. (Used on p. 261.)
5-13. Let .M; g/ be a Riemannian manifold.
(a) Show that the following formula holds for every smooth 1-form 2
1 .M /:
d.X; Y / D .rX /.Y / .rY /.X /:
(b) Generalize this to an arbitrary k-form 2 k .M / as follows:
d D .1/k .k C 1/ Alt.r/;
where Alt denotes the alternation operator defined in (B.9). [Hint: For
each p 2 M , do the computation in normal coordinates centered at p,
and note that both sides of the equation are well defined, independently
of the choice of coordinates.]
(Used on p. 209.)
5-14. Suppose .M; g/ is a Riemannian manifold, and let div and be the diver-
gence and Laplace operators defined on pages 32–33.
(a) Show that for every vector field X 2 X.M /, div X can be written in
terms of the total covariant derivative as div X D tr.rX /, and that if
X D X i Ei in terms of some local frame, then div X D X i Ii . [Hint:
Show that it suffices to prove the formulas at the origin in normal
coordinates.]
(b) Show that the Laplace operator acting on a smooth function u can be
expressed as
u D trg r 2 u ; (5.28)
and in terms of any local frame,
[Hint: Consider the 2-tensor field r 2 u n1 .u/g, and use one of Green’s
identities (Prob. 2-23).] (Used on p. 223.)
5-16. By analogy with the formula div X D tr.rX / developed in Problem 5-14,
we can define a divergence operator on tensor fields of any rank on a Rie-
mannian manifold. If F is any smooth k-tensor field (covariant, contravari-
ant, or mixed), we define the divergence of F by
where the trace is taken on the last two indices of the .k C 1/-tensor field
rF . (If F is purely contravariant, then trg can be replaced with tr, because
the next-to-last index of rF is already an upper index.) Extend the integra-
tion by parts formula of Problem 2-22 as follows: if F is a smooth covariant
k-tensor field and G is a smooth covariant .k C 1/-tensor field on a compact
smooth Riemannian manifold .M; g/ with boundary, then
Z Z Z
˝ ˛
hrF; Gi d Vg D F ˝ N [ ; G d Vgy hF; div Gi d Vg ;
M @M M
5-20. Show that single elliptic geometry (Example 5.30) satisfies Hilbert’s inci-
dence postulates and the elliptic parallel postulate if points are defined as
elements of RP 2 and lines are defined as images of maximal geodesics.
5-21. Let .M; g/ be a Riemannian or pseudo-Riemannian manifold and p 2 M .
Show that for every orthonormal basis .b1 ; : : : ; bn / for Tp M; there is a
smooth orthonormal frame .Ei / on a neighborhood of p such that Ei jp D bi
and .rEi /p D 0 for each i .
5-22. A smooth vector field X on a Riemannian manifold is called a Killing vector
field if the Lie derivative of the metric with respect to X vanishes. By Propo-
sition B.10, this is equivalent to the requirement that the metric be invariant
under the flow of X . Prove that X is a Killing vector field if and only if
the covariant 2-tensor field .rX /[ is antisymmetric. [Hint: Use Prop. B.9.]
(Used on pp. 190, 315.)
5-23. Let .M; g/ be a connected Riemannian manifold and p 2 M . An admissible
loop based at p is an admissible curve W Œa; b ! M such that .a/ D
.b/ D p. For each such loop , let P denote the parallel
transport
operator
Pab W Tp M ! Tp M along , and let Hol.p/ GL Tp M denote the set of
all automorphisms of Tp M obtained in this way:
˚
Hol.p/ D P W is an admissible loop based at p :
(a) Show that Hol.p/ is a subgroup of O Tp M (the set of all linear
isometries of Tp M ), called the holonomy group at p.
(b) Let Hol0 .p/ Hol.p/ denote the subset obtained by restricting to
loops that are path-homotopic to the constant loop. Show that Hol0 .p/
is a normal subgroup of Hol.p/, called the restricted holonomy group
at p.
(c) Given p; q2 M , show that there is an isomorphism of GL Tp M with
GL Tq M that takes Hol.p/ to Hol.q/.
(d) Show that M is orientable if and only if Hol.p/ SO Tp M (the set
of linear isometries with determinant C1) for some p 2 M .
(e) Show that g is flat if and only if Hol0 .p/ is the trivial group for some
p 2 M.
Chapter 6
Geodesics and Distance
In this chapter, we explore the relationships among geodesics, lengths, and distances
on a Riemannian manifold. A primary goal is to show that all length-minimizing
curves are geodesics, and that all geodesics are locally length-minimizing. Later, we
prove the Hopf–Rinow theorem, which states that a connected Riemannian manifold
is geodesically complete if and only if it is complete as a metric space. At the end
of the chapter, we study distance functions (which express the distance to a point
or other subset) and show that they can be used to construct coordinates, called
semigeodesic coordinates, that put a metric in a particularly simple form.
Most of the results of this chapter do not apply to general pseudo-Riemannian
metrics, at least not without substantial modification. For this reason, we restrict
our focus here to the Riemannian case. (For a treatment of lengths of curves in the
pseudo-Riemannian setting, see [O’N83].) Also, the theory of minimizing curves
becomes considerably more complicated in the presence of a nonempty boundary;
thus, unless otherwise stated, throughout this chapter we assume that .M; g/ is a Rie-
mannian manifold without boundary. And because we will be using the Riemannian
distance function, we assume for most results that M is connected.
Families of Curves
To make this rigorous, we introduce some more definitions. Let .M; g/ be a Rie-
mannian manifold.
Given intervals I; J R, a continuous map W J I ! M is called a one-
parameter family of curves. Such a family defines two collections of curves in M :
the main curves s .t / D .s; t / defined for t 2 I by holding s constant, and the
transverse curves .t/ .s/ D .s; t / defined for s 2 J by holding t constant.
If such a family is smooth (or at least continuously differentiable), we denote
the velocity vectors of the main and transverse curves by
Proof. Suppose and V satisfy the hypotheses, and set .s; t / D exp.t/ .sV .t //
(Fig. 6.2). By compactness of Œa; b, there is some positive " such that is de-
fined on ."; "/ Œa; b. By composition, is smooth on ."; "/ Œai1 ; ai for
each subinterval Œai1 ; ai on which V is smooth, and it is continuous on its whole
domain. By the properties of the exponential map, the variation field of is V .
Moreover, if V .a/ D 0 and V .b/ D 0, the definition gives .s; a/ .a/ and
.s; b/ .b/, so is proper. t
u
If V is a piecewise smooth vector field along ; we can compute the covariant
derivative of V either along the main curves (at points where V is smooth) or along
the transverse curves; the resulting vector fields along are denoted by Dt V and
Ds V respectively.
A key ingredient in the proof that minimizing curves are geodesics is the sym-
metry of the Levi-Civita connection. It enters into our proofs in the form of the
154 6 Geodesics and Distance
following lemma. (Although we state and use this lemma only for the Levi-Civita
connection, the proof shows that it is actually true for every symmetric connection in
TM .)
We can now compute an expression for the derivative of the length functional along
a variation of a curve. Traditionally, the derivative of a functional on a space of maps
is called its first variation.
where we have used the symmetry lemma in the last line. Setting s D 0 and noting
that S.0; t / D V .t / and T .0; t / D 0 .t / (which has length 1), we get
156 6 Geodesics and Distance
ˇ Z
d ˇˇ ˇ ˝
ai ˛
Lg s ˇŒa D Dt V ; 0 dt
ds ˇ sD0
i 1 ;ai
a
Z ia1
i d ˝ ˛
D hV; 0 i V; Dt 0 dt
a dt
˝ i 1 ˛ ˝ C
˛
D V .ai /; 0 .ai / V .ai1 /; 0 .ai1 /
Z ai
˝ ˛
V; Dt 0 dt:
ai 1
(The second equality follows from (5.3), and the third from the fundamental theorem
of calculus.) Finally, summing over i , we obtain (6.1). t
u
Because every admissible curve has a unit-speed parametrization and length is
independent of parametrization, the requirement in the above proposition that be
of unit speed is not a real restriction, but rather just a computational convenience.
Since the integrand is nonnegative and ' > 0 on .ai1 ; ai /, this shows that Dt 0 D 0
on each such subinterval.
Next we need to show that i 0 D 0 for each i between 0 and k, which is to say
that has no corners. For each such i , we can use a bump function in a coordinate
chart to construct a piecewise smooth vector field V along such that V .ai / D i 0
and V .aj / D 0 for j ¤ i . Then (6.2) reduces to ji 0 j2 D 0, so i 0 D 0 for each
i.
Finally, since the two one-sided velocity vectors of match up at each ai , it
follows from uniqueness of geodesics that jŒai ;ai C1 is the continuation of the
geodesic jŒai 1 ;ai , and therefore is smooth. t
u
The preceding proof has an enlightening geometric interpretation. Under the as-
sumption that Dt 0 ¤ 0, the first variation with V D 'Dt 0 is negative, which shows
that deforming in the direction of its acceleration vector field decreases its length
(Fig. 6.4). Similarly, the length of a broken geodesic is decreased by deforming it
Geodesics and Minimizing Curves 157
in the direction of a vector field V such that V .ai / D i 0 (Fig. 6.5). Geometrically,
this corresponds to “rounding the corner.”
The first variation formula actually tells us a bit more than is claimed in Theorem
6.4. In proving that is a geodesic, we did not use the full strength of the assumption
that the length of s takes a minimum when s D 0; we only used the fact that its
derivative is zero. We say that an admissible curve is a critical point of Lg if for
every proper variation s of , the derivative of Lg .s / with respect to s is zero at
s D 0. Therefore we can strengthen Theorem 6.4 in the following way.
Corollary 6.5. A unit-speed admissible curve is a critical point for Lg if and only
if it is a geodesic.
Proof. If is a critical point, the proof of Theorem 6.4 goes through without mod-
ification to show that is a geodesic. Conversely, if is a geodesic, then the first
term on the right-hand side of (6.2) vanishes by the geodesic equation, and the sec-
ond term vanishes because 0 has no jumps. t
u
The geodesic equation Dt 0 D 0 thus characterizes the critical points of the
length functional. In general, the equation that characterizes critical points of a func-
tional on a space of maps is called the variational equation or the Euler–Lagrange
equation of the functional. Many interesting equations in differential geometry arise
as variational equations. We touch briefly on three others in this book: the Einstein
equation (7.34), the Yamabe equation (7.59), and the minimal hypersurface equation
H 0 (Thm. 8.18).
Next we turn to the converse of Theorem 6.4. It is easy to see that the literal con-
verse is not true, because not every geodesic segment is minimizing. For example,
every geodesic segment on Sn that goes more than halfway around the sphere is
not minimizing, because the other portion of the same great circle is a shorter curve
segment between the same two points. For that reason, we concentrate initially on
local minimization properties of geodesics.
As usual, let .M; g/ be a Riemannian manifold. A regular (or piecewise regular)
curve W I ! M is said to be locally minimizing if every t0 2 I has a neighborhood
I0 I such that whenever a; b 2 I0 with a < b, the restriction of to Œa; b is
minimizing.
158 6 Geodesics and Distance
Our goal in this section is to show that geodesics are locally minimizing. The
proof will be based on a careful analysis of the geodesic equation in Riemannian
normal coordinates.
If " is a positive number such that expp is a diffeomorphism from the ball
B" .0/ Tp M to its image (where the radius of the ball is
measured with respect to
the norm defined by gp ), then the image set expp B" .0/ is a normal neighborhood
of p, called a geodesic ball in M , or sometimes an open geodesic ball for clarity.
Also, if the closed ball Bx" .0/ is contained in an open
set V Tp M on which
expp is a diffeomorphism onto its image, then expp Bx" .0/ is called a closed
geodesic ball, and expp @B" .0/ is called a geodesic sphere. Given such a V , by
compactness there exists "0 > " such that B"0 .0/ V , so every closed geodesic ball
is contained in an open geodesic ball of larger radius. In Riemannian normal co-
ordinates centered at p, the open and closed geodesic balls and geodesic spheres
centered at p are just the coordinate balls and spheres.
iSuppose
U is a normal neighborhood of p 2 M . Given any normal coordinates
x on U centered at p, define the radial distance function r W U ! R by
p
r.x/ D .x 1 /2 C C .x n /2 ; (6.4)
and the radial vector field on U X fpg (see Fig. 6.6), denoted by @r , by
xi @
@r D : (6.5)
r.x/ @x i
In Euclidean space, r.x/ is the distance to the origin, and @r is the unit vector field
pointing radially outward from the origin. (The notation is suggested by the fact that
@r is a coordinate derivative in polar or spherical coordinates.)
Geodesics and Minimizing Curves 159
Theorem 6.9 (The Gauss Lemma). Let .M; g/ be a Riemannian manifold, let U
be a geodesic ball centered at p 2 M , and let @r denote the radial vector field on
U Xfpg. Then @r is a unit vector field orthogonal to the geodesic spheres in U Xfpg.
Proof. We will work entirely in normal coordinates x i on U centered at p, using
the properties of normal coordinates described in Proposition 5.24.
Let q 2 U X fpg be arbitrary, with coordinate representation q 1 ; : : : ; q n , and let
p
b D r.q/ D .q 1 /2 C C .q n /2 , where r is the radial distance function defined by
(6.4). It follows that @r jq has the coordinate representation
ˇ
ˇ q i @ ˇˇ
ˇ
@r q D :
b @x i ˇq
For each s 2 ."; "/, s is a geodesic by Proposition 5.24(c). Its initial velocity is
s0 .0/ D .1=b/ i .s/@i jp, which is a unit vector by (6.6) and the fact that gp is the
Euclidean metric in coordinates; thus each s is a unit-speed geodesic.
As before, let S D @s and T D @t . It follows from the definitions that
ˇ
d ˇˇ
S.0; 0/ D s .0/ D 0I
ds ˇsD0
ˇ
d ˇˇ
T .0; 0/ D v .t / D vI
dt ˇtD0
ˇ
d ˇˇ
S.0; b/ D .s/ D wI
ds ˇsD0
ˇ
d ˇˇ
T .0; b/ D v .t / D v0 .b/ D @r jq :
dt ˇtDb
˝ ˛
Therefore hS; T i is zero when .s; t / D .0; 0/ and equal to w; @r jq when .s; t / D
.0; b/, so to prove the theorem it suffices to show that hS; T i is independent of t .
We compute
@
hS; T i D hDt S; T i C hS; Dt T i (compatibility with the metric)
@t
D hDs T ; T i C hS; Dt T i (symmetry lemma)
(6.7)
D hDs T ; T i C 0 (each s is a geodesic)
1 @
D jT j2 D 0 (jT j D js0 j 1 for all .s; t /):
2 @s
Geodesics and Minimizing Curves 161
Proof. By the result of Problem 2-10, it suffices to show that @r is orthogonal to the
level sets of r and @r .r/ j@r j2g . The first claim follows directly from the Gauss
lemma, and the second from the fact that @r .r/ 1 by direct computation in normal
coordinates, which in turn is equal to j@r j2g by the Gauss lemma. t
u
Here is the payoff: our first step toward proving that geodesics are locally mini-
mizing. Note that this is not yet the full strength of the theorem we are aiming for,
because it shows only that for each point on a geodesic, sufficiently small segments
of the geodesic starting at that point are minimizing. We will remove this restriction
after a little more work below.
Z b0
d
r. .b0 // r. .a0 // D r. .t // dt
a0 dt
Z b0
D dr. 0 .t // dt
a0
Z b0 ˝ ˛
D grad rj.t/ ; 0 .t / dt
a0 (6.8)
Z b0 ˇ ˇˇ ˇ
ˇgrad rj.t/ ˇ ˇ 0 .t /ˇ dt
a0
Z b0
D j 0 .t /j dt
a0
D Lg jŒa0 ;b0 Lg . /:
Thus Lg . / r. .b0 // r. .a0 // D c, so is minimizing.
Now suppose Lg . / D c. Then b D c, and both inequalities in (6.8) are equal-
ities. Because we assume that is a unit-speed curve, the second of these equali-
ties implies that a0 D 0 and b0 D b D c, since otherwise the segments of before
t D a0 and after t D b0 would contribute ˇ positive lengths.
ˇˇ ˇ The
˝ first equality then ˛
implies that the nonnegative expression ˇgrad rj.t/ ˇ ˇ 0 .t /ˇ grad rj.t/ ; 0 .t / is
identically zero on Œ0; b, which is possible only if 0 .t / is a positive multiple
of grad rj.t/ for each t . Since we assume that has unit speed, we must have
0 .t / D grad rj.t/ D @r j.t/ . Thus and are both integral curves of @r passing
through q at time t D c, so D . t
u
The next two corollaries show how radial distance functions, balls, and spheres
in normal coordinates are related to their global metric counterparts.
Proof. Since every closed geodesic ball is contained in an open geodesic ball of
larger radius, we need only consider the open case. If x is in the open geodesic ball
expp Bc .0/ , the radial geodesic from p to x is minimizing by Proposition 6.11.
Since its velocity is equal to @r , which is a unit vector in both the g-norm and the
Euclidean norm in normal coordinates, the g-length of is equal to its Euclidean
length, which is r.x/. t
u
is also in the closed metric ball of radius c. Conversely, suppose q … V . Let S be
the geodesic sphere expp @Bc .0/ . The complement of S is the disjoint union of
the open sets expp Bc .0/ and M X expp Bxc .0/ , and hence disconnected. Thus if
W Œa; b ! M is any admissible curve from p to q, there must be a time t0 2 .a; b/
when .t0 / 2 S , and then Corollary 6.12 shows that the length of jŒa;t0 must be at
least c. Since jŒt0 ;b must have positive length, it follows that dg .p; q/ > c, so q is
not in the closed metric ball of radius c around p.
Next, let W D expp Bc .0/ be an open geodesic ball of radius c. Since W is
the union of all closed geodesic balls around p of radius less than c, and the open
metric ball of radius c is similarly the union of all closed metric metric balls of
smaller radii, the result of the preceding paragraph shows that W is equal to the
open metric ball of radius c.
Finally, if S D expp @Bc .0/ is a geodesic sphere of radius c, the arguments
above show that S is equal to the closed metric ball of radius c minus the open
metric ball of radius c, which is exactly the metric sphere of radius c. t
u
The last corollary suggests a simplified notation for geodesic balls and spheres
in M . From now on, we will use the notations Bc .p/ D expp Bc .0/ , Bxc .p/ D
expp Bxc .0/ , and Sc .p/ D expp @Bc .0/ for open and closed geodesic balls and
geodesic spheres, which we now know are also open and closed metric balls and
spheres. (To avoid confusion, we refrain from using this notation for other metric
balls and spheres unless explicitly stated.)
we may assume that the metric g satisfies an estimate of the form (2.21) for all
x 2 W . This means that for each x 2 W , the coordinate ball B" .0/ Tx M contains
a gx -ball of radius at least "=c. Put ı D "=c; we have shown that for each x 2 W ,
there is a g-geodesic ball of radius ı in M centered at x.
Now, shrinking W once more, we may assume that its diameter (with respect
to the metric dg ) is less than ı. It follows that for each x 2 W , the entire set W
is contained in the metric ball of radius ı around x, and Corollary 6.13 shows that
this metric ball is also a geodesic ball of radius ı. Thus W is the required uniformly
normal neighborhood of p. t
u
We are now ready to prove the main result of this section.
Uniformly Normal Neighborhoods 165
If there is no upper bound to the radii of such balls (as is the case, for example, on
Rn ), then we set inj.p/ D 1. Then we define the injectivity radius of M , denoted
by inj.M /, to be the infimum of inj.p/ as p ranges over points of M . It can be zero,
positive, or infinite. (The terminology is explained by Problem 10-24.)
Lemma 6.16. If .M; g/ is a compact Riemannian manifold, then inj.M / is positive.
Proof. For each x 2 M , there is a positive number ı.x/ such that x is contained in a
uniformly ı.x/-normal neighborhood Wx , and inj.x 0 / ı.x/ for each x 0 2 W . Since
M is compact, it is covered by finitely many such neighborhoods Wx1 ; : : : ; Wxk .
Therefore, inj.M / is at least equal to the minimum of ı.x1 /; : : : ; ı.xk /. It cannot be
infinite, because a compact metric space is bounded, and a geodesic ball of radius c
contains points whose distances from the center are arbitrarily close to c. t
u
In addition to uniformly normal neighborhoods, there is another, more special-
ized, kind of normal neighborhood that is frequently useful. Let .M; g/ be a Rie-
mannian manifold. A subset U M is said to be geodesically convex if for each
p; q 2 U , there is a unique minimizing geodesic segment from p to q in M , and the
image of this geodesic segment lies entirely in U .
The next theorem says that every sufficiently small geodesic ball is geodesically
convex.
Theorem 6.17. Let .M; g/ be a Riemannian manifold. For each p 2 M , there exists
"0 > 0 such that every geodesic ball centered at p of radius less than or equal to "0
is geodesically convex.
Proof. Problem 6-5. t
u
Completeness
Suppose .M; g/ is a connected Riemannian manifold. Now that we can view M as
a metric space, it is time to address one of the most important questions one can ask
about a metric space: Is it complete? In general, the answer is no: for example, if M
is an open ball in Rn with its Euclidean metric, then every sequence in M that converges
in Rn to a point in @M is Cauchy, but not convergent in M .
In Chapter 5, we introduced another notion of completeness for Riemannian and
pseudo-Riemannian manifolds: recall that such a manifold is said to be geodesically
complete if every maximal geodesic is defined for all t 2 R. For clarity, we will
use the phrase metrically complete for a connected Riemannian manifold that is
complete as a metric space with the Riemannian distance function, in the sense that
every Cauchy sequence converges.
The Hopf–Rinow theorem, which we will state and prove below, shows that these
two notions of completeness are equivalent for connected Riemannian manifolds.
Before we prove it, let us establish a preliminary result, which will have other im-
portant consequences besides the Hopf–Rinow theorem itself.
Completeness 167
(This would be the case, for example, if were an initial segment of a minimizing
geodesic from p to q; but we are not assuming that.) To prove (a), it suffices to show
that there is a geodesic segment W Œa; b ! M that begins at p, aims at q, and has
length equal to dg .p; q/, for then the fact that is minimizing means that
dg .p; .b// D Lg . / D dg .p; q/, and (6.9) becomes
To this end, let W Œa0 ; b0 ! M be any admissible curve from p to q. Let t0 be the
first time hits S" .p/, and let 1 and 2 denote the restrictions of 1 to Œa0 ; t0 and
Œt0 ; b0 , respectively (Fig. 6.11). Since every point in S" .p/ is at a distance " from
168 6 Geodesics and Distance
p, we have Lg .1 / dg .p; .t0 // D dg .p; x/; and by our choice of x we have
Lg .2 / dg . .t0 /; q/ dg .x; q/. Putting these two inequalities together yields
Taking the infimum over all such , we find that dg .p; q/ dg .p; x/ C dg .x; q/.
The opposite inequality is just the triangle inequality, so (6.10) holds.
Now let T D dg .p; q/ and
˚ ˇ
A D b 2 Œ0; T W ˇŒ0;b aims at q :
We have just shown that " 2 A. Let A D sup A ". By continuity of the distance
function, it is easy to see that A is closed in Œ0; T , and therefore A 2 A. If A D T ,
then jŒ0;T is a geodesic of length T D dg .p; q/ that aims at q, and by the remark
above we are done. So we assume A < T and derive a contradiction.
Let y D .A/, and choose ı > 0 such that there is a closed geodesic ball Bxı .y/
around y, small enough that it does not contain q (Fig. 6.12). The fact that A 2 A
means that
dg .y; q/ D dg .p; q/ dg .p; y/ D T A:
Let z 2 Sı .y/ be a point where dg .z; q/ attains its minimum, and let W Œ0; ı ! M
be the unit-speed radial geodesic from y to z. By exactly the same argument as
before, aims at q, so
Therefore, the admissible curve consisting of jŒ0;A (of length A) followed by (of
length ı) is a minimizing curve from p to z. This means that it has no corners, so z must
lie on , and in fact, z D .A C ı/. But then (6.11) says that
Completeness 169
Note that both expressions on the right-hand side are geodesics, and they have the
same position and velocity when t D tj . Therefore, by uniqueness of geodesics, the
two definitions agree where they overlap. Since tj C ı > b, z is an extension of
past b, which is a contradiction. t
u
A connected Riemannian manifold is simply said to be complete if it is either
geodesically complete or metrically complete; the Hopf–Rinow theorem then im-
plies that it is both. For disconnected manifolds, we interpret “complete” to mean
geodesically complete, which is equivalent to the requirement that each component
be a complete metric space. As mentioned in the previous chapter, complete mani-
folds are the natural setting for global questions in Riemannian geometry.
We conclude this section by stating three important corollaries, whose proofs are
easy applications of Lemma 6.18 and the Hopf–Rinow theorem.
The Hopf–Rinow theorem and Corollary 6.20 are key ingredients in the follow-
ing theorem about Riemannian covering maps. This theorem will play a key role in
the proofs of some of the local-to-global theorems in Chapter 12.
Theorem 6.23. Suppose M ; gz and .M; g/ are connected Riemannian manifolds
with M complete, and W M ! M is a local isometry. Then M is complete and
is a Riemannian covering map.
S
z˛ \ U
Fig. 6.16: Proof that U zˇ D ¿ Fig. 6.17: Proof that 1 .U / z
˛ U˛
Closed Geodesics
Round spheres have the remarkable property that all of their geodesics are closed
when restricted to appropriate intervals. Of course, this is not typically the case,
even for compact Riemannian manifolds; but it is natural to wonder whether closed
geodesics exist in more general manifolds. Much work has been done in Riemannian
geometry to determine how many closed geodesics exist in various situations. Here
we can only touch on the simplest case; these results will be useful in some of the
proofs of local-to-global theorems in Chapter 12.
A continuous path W Œ0; 1 ! M is called a loop if .0/ D .1/. Two loops
0 ; 1 W Œ0; 1 ! M are said to be freely homotopic if they are homotopic through
closed paths (but not necessarily preserving the base point), that is, if there exists a
homotopy H W Œ0; 1 Œ0; 1 ! M satisfying
This is an equivalence relation on the set of all loops in M , and an equivalence class
is called a free homotopy class. The trivial free homotopy class is the equivalence
class of any constant path.
I Exercise 6.27. Given a connected manifold M and a point x 2 M , show that a loop
based at x represents the trivial free homotopy class if and only if it represents the identity
element of 1 .M; x/.
The next proposition shows that closed geodesics are easy to find on compact
Riemannian manifolds that are not simply connected.
Proposition 6.28 (Existence of Closed Geodesics). Suppose .M; g/ is a compact,
connected Riemannian manifold. Every nontrivial free homotopy class in M is rep-
resented by a closed geodesic that has minimum length among all admissible loops
in the given free homotopy class.
Proof. Problem 6-17. t
u
The previous proposition guarantees the existence of at least one closed geodesic
on every non-simply-connected compact Riemannian manifold. In fact, it was
proved in 1951 by the Russian mathematicians Lazar Lyusternik and Abram Fet
that closed geodesics exist on every compact Riemannian manifold, but the proof
in the simply connected case is considerably harder. Proofs can be found in [Jos17]
and [Kli95].
Distance Functions
Suppose .M; g/ is a connected Riemannian manifold and S M is any subset. For
each point x 2 M , we define the distance from x to S to be
˚
dg .x; S / D inf dg .x; p/ W p 2 S :
Distance Functions 175
The simplest example of a distance function occurs when the set S is just a
singleton, S D fpg. Inside a geodesic ball around p, Corollary 6.12 shows that
dg .x; S / D r.x/, the radial distance function, and Corollary 6.10 shows that it has
unit gradient where it it smooth (everywhere inside the geodesic ball except at p
itself). The next theorem is a far-reaching generalization of that result.
On the other hand, for 0 t ", the fact that ˛.t / 2 Bx" .x/ implies
ˇ ˇ
ˇd ˇ ˇ˝ ˛ˇ ˇ ˇ
ˇ f .˛.t //ˇ D ˇ grad f j˛.t/ ; ˛ 0 .t / ˇ ˇgrad f j˛.t/ ˇ j˛ 0 .t /j 1 ı:
ˇ dt ˇ
Thus f .˛.t // f .x/ .1 ı/t for all such t . In particular, for t D ", this means
that
dg .˛."/; S / dg .x; S / .1 ı/": (6.14)
Combining (6.13) and (6.14) yields c > "ı, contradicting our choice of c. t
u
Motivated by the previous theorem, if .M; g/ is a Riemannian manifold and U
M is an open subset, we define a local distance function on U to be a continuously
differentiable function f W U ! R such that j grad f jg 1 in U . Theorem 6.34
and Corollary 6.35 below will justify this terminology. But first, we develop some
important general properties of local distance functions.
Proof. Let F 2 X.U / denote the unit vector field grad f . The definition of the
gradient shows that for every vector field W , we have
Wf D df .W / D hF; W i; (6.15)
and therefore
Ff D hF; F i D j grad f j2 D 1: (6.16)
For every smooth vector field W on U , we have
Dt 0 .t / D rF F j.t/ D 0;
so is a geodesic. t
u
Proof. This is proved exactly as in (6.8), noting that the only properties of r we used
in that computation were that it is continuous on the image of and continuously
differentiable on .a0 ; b0 / , and its gradient has unit length there. t
u
The next theorem and its corollary explain why the name “local distance func-
tion” is justified. Its proof is an adaptation of the proof of Proposition 6.11.
Proof. For each p 2 S , there are positive numbers "p ; ıp such that B"p .p/ is a
uniformly ıp -normal geodesic ball and B2"p .p/ U . This means that B"p .p/ is
contained in the open geodesic ball of radius ıp around each of its points. In partic-
ular, B"p .p/ Bıp .p/, which means that ıp "p , and thus every geodesic starting
at a point of B"p .p/ is defined at least for t 2 ."p ; "p /. Let U0 be the union of all
of the geodesic balls B"p .p/ for p 2 S , which is a neighborhood of S contained in
U (Fig. 6.18).
Let x 2 U0 be arbitrary, and let c D f .x/. We will show that dg .x; S / D c. If
x 2 S , then dg .x; S / D 0 D c, so we may as well assume x … S .
There is some p 2 S such that x 2 B"p .p/, which means that dg .x; S / < "p
and geodesics starting at x are defined at least on ."p ; "p /. Also, if ˛ W Œ0; b !
B" .p/ is the radial geodesic segment from p to x, it follows from Lemma 6.33 that
Lg .˛/ jf .x/ f .p/j D c, and we conclude that c Lg .˛/ < "p as well.
Let W ."p ; "p / ! U be the unit-speed geodesic starting at x with initial veloc-
ity equal to grad f jx . By Theorem 6.32 and uniqueness of geodesics, coincides
with an integral curve of grad f as long as .t / 2 U X S , which is to say as long
as f ..t // ¤ 0. For all such t we have
d ˝ ˛ ˇ ˇ2
f ..t // D grad f j.t/ ; 0 .t / D ˇgrad f j.t/ ˇ D 1;
dt
178 6 Geodesics and Distance
The most important local distance functions are those associated with embedded
submanifolds. As we will see in this section, such distance functions are always
smooth near the manifold.
Suppose .M; g/ is a Riemannian n-manifold (without boundary) and P M
is an embedded k-dimensional submanifold. Let NP denote the normal bundle of
P in M , and let U M be a normal neighborhood of P in M , which is the dif-
feomorphic image of a certain open subset V NP under the normal exponential
map. (Such a neighborhood always exists by Thm. 5.25.) We begin by constructing
generalizations of the radial distance function and radial vector field (see (6.4) and
(6.5)). Recall the definition of Fermi coordinates from Chapter 5 (see Prop. 5.26).
Proposition 6.37. Let P be an embedded submanifold of a Riemannian manifold
.M; g/ and let U be any normal neighborhood of P in M . There exist a unique
continuous function r W U ! Œ0; 1/ and smooth vector field @r on U X P that
have
1 the following coordinate
representations in terms of any Fermi coordinates
x ; : : : ; x k ; v 1 ; : : : ; v nk for P on a subset U0 U :
1 k 1 nk
q
r x ;:::;x ;v ;:::;v D .v 1 /2 C C .v nk /2 ; (6.17)
1 nk
v @ v @
@r D 1
CC : (6.18)
r.x; v/ @v r.x; v/ @v nk
Proof. The uniqueness, continuity, and smoothness claims follow immediately from
the coordinate expressions (6.17) and (6.18), so we need only prove that r and @r
can be globally defined so as to have the indicated coordinate expressions in any
Fermi coordinates.
Let V NS be the subset that is mapped diffeomorphically onto U by the
normal exponential map E. Define a function
W V ! Œ0; 1/ by
.p; v/ D jvjg ,
and define r W U ! Œ0; 1/ by r D
ı E 1 . Any Fermi coordinates for P are
defined by choosing local coordinates x 1 ; : : : ; x k for P and a local orthonormal
1 k 1 nk
frame .E˛ / for NP , and assigning
the coordinates x .p/; : : : ; x .p/; v ; : : : ; v
to the point E p; v ˛ E˛ jp . (Here we are using the summation convention with
Greek indices
running from 1 to n k.) Because the frame is orthonormal, for each
.p; v/ D p; v ˛ E˛ jp 2 V we have r.E.p; v//2 D
.p; v/2 D .v 1 /2 C C.v nk /2 ,
which shows that r has the coordinate representation (6.17).
To define @r , let q be an arbitrary point in U X P . Then q D expp .v/ for a unique
.p; v/ 2 V , and the curve W Œ0; 1 ! U given by .t / D expp .t v/ is a geodesic
from p to q. Define
1 0
@r jq D .1/; (6.19)
r.q/
which is independent of the choice of coordinates. Proposition 5.26 shows that in
coordinates, if we writev D v E˛ jp , then has the coordinate formula
˛
any Fermi
.t / D x 1 ; : : : ; x k ; t v 1 ; : : : ; t v nk , and therefore 0 .t / D v ˛ @=@v ˛ j.t/ . It follows
that @r has the coordinate formula (6.18). t
u
By analogy with the special case in which P is a point, we call r the radial
distance function for P and @r the radial vector field for P.
Theorem 6.38 (Gauss Lemma for Submanifolds). Let P be an embedded sub-
manifold of a Riemannian manifold .M; g/, let U be a normal neighborhood of P
in M , and let r and @r be defined as in Proposition 6.37. On U X P , @r is a unit
vector field orthogonal to the level sets of r.
Proof. The proof is a dressed-up version of the proof of the ordinary Gauss lemma.
Let q 2 U X P be arbitrary, and let x 1 ; : : : ; x k ; v 1 ; : : : ; v nk be the coordinate
representation of q in some choice of Fermi coordinates associated with a local
orthonormal frame .E˛ / for NP . As in the proof of Proposition 6.37, q D .1/,
where is the geodesic expp .t v/ for some p 2 P and v D v ˛ E˛ jp 2 Np M . Since
the frame .E˛ / is orthonormal, we have
q
j 0 .0/jg D jvjg D .v 1 /2 C C .v nk /2 D r.q/:
Since geodesics have constant speed, it follows that j 0 .1/jg D r.q/ as well, and
then (6.19) shows that @r jq is a unit vector.
Next we show that @r is orthogonal to the level sets of r. Suppose q 2 U X P , and
write q D expp0 .v0 / for some p0 2 P and v0 2 Np P with v0 ¤ 0. Let b D r.q/ D
jv0 jg , so q lies in the level set r 1 .b/. The coordinate representation (6.17) shows
that this is a regular level set, and hence an embedded submanifold of U .
180 6 Geodesics and Distance
Let w 2 Tq M be an arbitrary vector tangent to this level set, and let W ."; "/ !
U be a smooth curve lying in the same level set, with .0/ D q and 0 .0/ D w. We
can write .s/ D expx.s/ .v.s//, where x.s/ 2 P and v.s/ 2 Nx.s/ P with jv.s/jg D
b. The initial condition .0/ D q translates to x.0/ D p0 and v.0/ D v0 . Define a
smooth one-parameter family of curves W ."; "/ Œ0; b ! M by
t 1 t n
.s; t / D expx.s/ v .s/; : : : ; v .s/ :
b b
Then the same computation as in (6.7) shows that .@=@t /hS; T ig 0, and therefore
hw; @r jq ig D hS.0; b/; T .0; b/ig D hS.0; 0/; T .0; 0/ig D .1=b/hx 0 .0/; v0 ig , which
is zero because x 0 .0/ is tangent to P and v0 is normal to it. This proves that @r is
orthogonal to the level sets of r. t
u
Corollary 6.39. Assume the hypotheses of Theorem 6.38.
(a) @r is equal to the gradient of r on U X P .
(b) r is a local distance function.
(c) Each unit-speed geodesic W Œa; b/ ! U with 0 .a/ normal to P coincides with
an integral curve of @r on .a; b/.
(d) P has a tubular neighborhood in which the distance in M to P is equal to r.
Proof. By direct computation in Fermi coordinates using formulas (6.17) and
(6.18), @r .r/ 1, which is equal to j@r j2g by the previous theorem. Thus @r D grad r
on U X P by Problem 2-10. Because grad r is a unit vector field, r is a local dis-
tance function. By Proposition 5.26, the geodesics in U that start normal to P are
represented in any Fermi coordinates by t 7! .x 1 ; : : : ; x k ; t v 1 ; : : : ; t v nk /, and such
a geodesic has unit speed if and only if .v 1 /2 C C .v nk /2 D 1. Another direct
computation shows that each such curve is an integral curve of @r wherever r ¤ 0.
Finally, to prove (d), note that Theorem 6.34 shows that there is some neighborhood
Uz0 of P in M on which r.x/ D dg .x; P /; if we take U0 to be a tubular neighborhood
of P in Uz0 , then U0 satisfies the conclusion. t
u
When P is compact, we can say more.
Semigeodesic Coordinates 181
Proof. First we show that r can be extended continuously to Ux" by setting r.q/ D "
for q 2 @U" . Indeed, suppose q 2 @U" and qi is any sequence of points in U" converg-
ing to q. Then lim supi r.qi / " because r.qi / < " for each i . Let c D lim infi r.qi /;
we will prove the result by showing that c D ". Suppose for the sake of contra-
diction that c < ". By passing to a subsequence, we may assume that r.qi / ! c.
We can write qi D exppi .vi / for pi 2 P and vi 2 Npi P , and because P is com-
pact and limi jvi jg D limi r.qi / D c, we can pass to a further subsequence and
assume that .pi ; vi / ! .p; v/ 2 NP with jvjg D c < ". Then we have q D limi qi D
limi exppi .vi / D expp v, which lies in the open set U" , contradicting our assumption
that q 2 @U" . Henceforth, we regard r as a continuous function on Ux" .
Now to prove the theorem, let W" denote the "-neighborhood of P in the metric
space sense. Suppose first that q 2 M X U" , and suppose ˛ W Œa; b ! M is any
admissible curve from a point of P to q. There is a first time b0 2 Œa; b that ˛.b0 / 2
@U" , and then Lemma 6.33 shows that
Lg .˛/ Lg ˛jŒa;b0 jr.˛.b0 // r.˛.a//j D ":
where a0 is the last time that ˛.a0 / 2 P . On the other hand, if ˛.t / does not remain
in U , then there is a first time b0 such that ˛.b0 / 2 @U" , and the argument in the
preceding paragraph shows that Lg .˛/ " > r.q/. Thus dg .q; P / D r.q/ for all
q 2 U" . Since r.q/ < " for all such q, it follows also that U" W" . t
u
Semigeodesic Coordinates
Local distance functions can be used to build coordinate charts near submanifolds
in which the metric has a particularly simple form. We begin by describing the kind
of coordinates we are looking for.
1Let .M;n g/ be an n-dimensional Riemannian manifold. Smooth local coordinates
x ; : : : ; x on an open subset
U M are called semigeodesic coordinates if each
x n -coordinate curve t 7! x 1 ; : : : ; x n1 ; t is a unit-speed geodesic that meets each
level set of x n orthogonally.
182 6 Geodesics and Distance
Because of the distinguished role played by the last coordinate function, through-
out the rest of this section we will use the summation convention with Latin indices
running from 1 to n and Greek indices running from 1 to n 1.
We will see below that semigeodesic coordinates are easy to construct. But first,
let us develop some alternative characterizations of them.
Proof. We begin by showing that (b) , (c) , (d) , (e) and (c) ,(f). Note that (b)
is equivalent to the coordinate matrix of g having the block form 0 10 , where the
asterisk represents an arbitrary .n 1/ .n 1/ positive definite symmetric matrix,
while (c) is equivalent to the inverse matrix having the same form. It follows from
Cramer’s rule that the matrix of g has this form if and only if its inverse does, and
thus (b) is equivalent to (c).
The equivalence of (c) and (d) follows from the definitions of the gradient and of
the inner product on 1-forms: for all i; j D 1; : : : ; n,
˝ i ˛ ˝ ˛ ˝ ˛
dx ; dx j g D .dx i /# ; .dx j /# g D grad x i ; grad x j g :
The equivalence of (d) and (e) also follows from the definition of the gradient:
hgrad x ˛ ; grad x n i g D dx ˛ grad x n D .grad x n /.x ˛ / for each ˛, which means that
x ˛ is constant along the grad x n integral curves if and only if hgrad x ˛ ; grad x n i g D
0. Finally, by examining the individual components of the coordinate formula
grad x n D g nj @j , we see that (c) is also equivalent to (f).
To complete the proof, we show that (a) , (b). Assume first that (a) holds.
Because the x n -coordinate curves have unit speed, it follows that j@n jg 1. The
tangent space to any x n -level set is spanned at each point by @1 ; : : : ; @n1 , and (a)
guarantees that @n is orthogonal to each of these, showing that (b) holds. Con-
versely, if we assume (b), the first part of the proof shows that (f) holds as well,
so j grad x n jg j@n jg 1, showing that x n is a local distance function. Thus the
x n -coordinate curves are also integral curves of grad x n and hence are unit-speed
geodesics by Theorem 6.32. The fact that h@˛ ; @n i D 0 for ˛ D 2; : : : ; n implies that
these geodesics are orthogonal to the level sets of x n , thus proving (a). t
u
Part (b) of this proposition leads to the following simplified coordinate represen-
tations for the metric and Christoffel symbols in semigeodesic coordinates. Recall
that implied summations with Greek indices run from 1 to n 1.
Semigeodesic Coordinates 183
Corollary 6.42. Let x i be semigeodesic coordinates on an open subset of a Rie-
mannian n-manifold .M; g/.
(a) The metric has the following coordinate expression:
2
g D dx n C g˛ˇ x 1 ; : : : ; x n dx ˛ dx ˇ :
Proof. Part (a) follows immediately from part (b) of Proposition 6.41, and (b) is
proved by inserting gnn D 1 and g˛n D gn˛ D 0 into formula (5.8) for the Christoffel
symbols. t
u
Proposition 6.41(e) gives us an effective way to construct semigeodesic coordi-
nates: if r is any smooth local distance function (for example, the distance from
a point or a smooth submanifold), just set x n D r, choose any local coordinates
x 1 ; : : : ; x n1 for a level set of r, and then extend them to be constant along the
integral curves of grad r. Here are some explicit examples.
Example 6.45 (Polar Normal Coordinates). Polar coordinates for Rn are con-
structed by choosing a smooth local parametrization y W Uy ! U Sn1 for an
open subset U of Sn1 , and defining y W Uy RC ! Rn by y 1 ; : : : ; n1 ; r D
r y 1 ; : : : ; n1 . It is straightforward to show that the differential of y van-
y y 1 is a smooth coordinate map on the open subset
ishes nowhere,
nD
so
y y C
U D U R R X f0g. Familiar examples are ordinary polar coordinates in
the plane and spherical coordinates in R3 . Such coordinates have the property
that the last coordinate function is r.x/ D jxj.
Now let .M; g/ be a Riemannian n-manifold, p a point in M , and ' any normal
coordinate chart defined
on a normal neighborhood V of p. For every choice of polar
coordinates U;
y for Rn X f0g, we obtain a smooth coordinate map
D
y ı ' on
an open subset of V X fpg (see Fig. 6.19). Such coordinates are called polar normal
coordinates. They have the property that the last coordinate function r is the radial
distance function on V , and the other coordinates are constant along the integral
curves of grad r, so they are semigeodesic coordinates. //
Problems
hp; qi
dgV R .p; q/ D R arccos ;
R2
where h; i is the Euclideaninner product
on R
nC1
.
(b) Prove that the metric space S .R/; dgV R has diameter R.
n
where j j represents the Euclidean norm in Rn . [Hint: First use the result of
Problem 3-5 to show that it suffices to consider the case R D 1. Then use a
rotation to reduce to the case n D 2, and use the group action of Problem 3-8
to show that it suffices to consider the case in which p is the origin.]
186 6 Geodesics and Distance
6-4. In Chapter 2, we started with a Riemannian metric and used it to define the
Riemannian distance function. This problem shows how to go back the other
way: the distance function determines the Riemannian metric. Let .M; g/ be
a connected Riemannian manifold.
(a) Show that if W ."; "/ ! M is any smooth curve, then
dg ..0/; .t //
j 0 .0/jg D lim :
t&0 t
(b) Show that if g and gz are two Riemannian metrics on M such that
dg .p; q/ D dgz .p; q/ for all p; q 2 M , then g D gz.
6-5. Prove Theorem 6.17 (sufficiently small geodesic balls are geodesically con-
vex) as follows.
(a) Let .M; g/ be a Riemannian manifold, let p 2 M be fixed, and let W
be a uniformly normal neighborhood of p. For " > 0 small enough that
B3" .p/ W , define a subset W" TM R by
Define f W W" ! R by
dg .expp t v; expp t w/
lim D jv wjg :
t!0 t
[Hint: Use the Taylor series for g in Riemannian normal coordinates on
a convex geodesic ball centered at p.]
(b) Show that ' takes geodesics to geodesics.
(c) For each p 2 M , show that there exist an open ball B" .0/ Tp M
and a continuous map W B" .0/ ! T'.p/ M satisfying .0/ D 0 and
exp'.p/ .v/ D '.expp v/ for all v 2 B" .0/.
(d) With as above, show that j .v/ .w/jgz D jv wjg for all v; w 2
B" .0/, and conclude from Problem 2-2 that is the restriction of a
linear isometry.
(e) With p and as above, show that ' is smooth on a neighborhood of p
and d'p D .
(f) Conclude that ' is a Riemannian isometry.
6-8. Suppose .M; g/ and M ; gz are connected Riemannian manifolds (not nec-
essarily complete), and for each i 2 ZC , 'i W M ! M is a Riemannian isom-
etry such that 'i converges pointwise to a map ' W M ! M . Show that '
is a Riemannian isometry. [Hint: Once you have shown that ' is a local
isometry, to show that ' is surjective, suppose y is a limit point of '.M /.
Choose x 2 M such that '.x/ lies in a uniformly normal neighborhood of
sequence of points .xi ; vi / 2 TM
y, and show that there exists a convergent
such that 'i .xi / D '.x/ and 'i expxi vi D y.]
6-9. Suppose .M; g/ is a (not necessarily connected) Riemannian manifold. In
Problem 2-30, you were asked to show that there is a distance function
d W M M ! R that induces the given topology and restricts to the Rie-
mannian distance on each connected component of M . Show that if each
component of M is geodesically complete, then d can be chosen so that
.M; d / is a complete metric space. Show also that if M has infinitely many
components, then d can be chosen so that .M; d / is not complete.
6-10. A curve W Œ0; b/ ! M (with 0 < b 1) is said to diverge to infinity if
for every compact set K M , there is a time T 2 Œ0; b/ such that .t / … K
for t > T . (For those who are familiar with one-point compactifications, this
means that .t / converges to the “point at infinity” in the one-point com-
pactification of M as t % b.) Prove that a connected Riemannian manifold is
complete if and only if every regular curve that diverges to infinity has
infinite length. (The length of a curve whose domain is not compact is just
the supremum of the lengths of its restrictions to compact subintervals.)
6-11. Suppose .M; g/ is a connected Riemannian manifold, P M is a connected
embedded submanifold, and gy is the induced Riemannian metric on P .
(a) Show that dgy .p; q/ dg .p; q/ for p; q 2 P .
188 6 Geodesics and Distance
(b) Prove that if .M; g/ is complete and P is closed in M , then P; gy is
complete.
(c) Give an example of a complete Riemannian manifold .M; g/ and a con-
nected embedded submanifold P M that is complete but not closed
in M .
6-12. Let .M; g/ be a connected Riemannian manifold.
(a) Suppose there exists ı > 0 such that for each p 2 M , every maximal
unit-speed geodesic starting at p is defined at least on an interval of the
form .ı; ı/. Prove that M is complete.
(b) Prove that if M has positive or infinite injectivity radius, then it is com-
plete.
(c) Prove that if M is homogeneous, then it is complete.
(d) Give an example of a complete, connected Riemannian manifold that
has zero injectivity radius.
6-13. Let G be a connected compact Lie group. Show that the Lie group expo-
nential map of G is surjective. [Hint: Use Problem 5-8.]
6-14. Let .M; g/ be a connected Riemannian manifold.
(a) Show that M is complete if and only if the compact subsets of M are
exactly the closed and bounded ones.
(b) Show that M is compact if and only if it is complete and bounded.
6-15. Let S be the unit 2-sphere minus its north and south poles, and let M D
.0; / R. Define q W M ! S by q.'; / D .sin ' cos ; sin ' sin ; cos '/,
and let g be the metric on M given by pulling back the round metric: g D
q g.
V (Think of M as an infinitely long onion skin wrapping infinitely many
times around an onion.) Prove that the Riemannian manifold .M; g/ has diam-
eter , but contains infinitely long properly embedded geodesics.
6-16. Suppose .M; g/ is a complete, connected Riemannian manifold with positive
or infinite injectivity radius.
(a) Let
2 .0; 1 denote the injectivity radius of M , and define T M to be
the subset of TM consisting of vectors of length less than
, and D to
be the subset f.p; q/ W dg .p; q/ <
g M M . Define E W T M ! D
by E.x; v/ D .x; expx v/. Prove that E is a diffeomorphism.
(b) Use part (a) to prove that if B is a topological space and F; G W B ! M
are continuous maps such that dg .F .x/; G.x// < inj.M / for all x 2 B,
then F and G are homotopic.
6-17. Prove Proposition 6.28 (existence of a closed geodesic in a free homotopy
class). [Hint: Use Prop. 6.25 to show that the given free homotopy class is
represented by a geodesic loop, i.e., a geodesic whose starting and ending
points are the same. Show that the lengths of such loops have a positive great-
est lower bound; then choose a sequence of geodesic loops whose lengths
approach that lower bound, and show that a subsequence converges uniformly
Problems 189
to a geodesic loop whose length is equal to the lower bound. Use Problem
6-16 to show that the limiting curve is in the given free homotopy class, and
apply the first variation formula to show that the limiting curve is in fact a
closed geodesic.]
6-18. A connected Riemannian manifold .M; g/ is said to be k-point homoge-
neous if for any two ordered k-tuples .p1 ; : : : ; pk / and .q1 ; : : : ; qk / of points
in M such that dg .pi ; pj / D dg .qi ; qj / for all i; j , there is an isometry
' W M ! M such that '.pi / D qi for i D 1; : : : ; k. Show that .M; g/ is 2-point
homogeneous if and only if it is isotropic. [Hint: Assuming that M is isotropic,
first show that it is homogeneous by considering the midpoint of a geodesic
segment joining sufficiently nearby points p; q 2 M , and then use the result
of Problem 6-12(c) to show that it is complete.] (Used on pp. 56, 261.)
6-19. Prove that every Riemannian symmetric space is homogeneous. [Hint: Pro-
ceed as in Problem 6-18.] (Used on p. 78.)
6-20. A connected Riemannian manifold is said to be extendible if it is isometric
to a proper open subset of a larger connected Riemannian manifold.
(a) Show that every complete, connected Riemannian manifold is nonex-
tendible.
(b) Show that the converse is false by giving an example of a nonextendible
connected Riemannian manifold that is not complete.
6-21. Let .M; g/ be a complete, connected, noncompact Riemannian manifold.
Define a ray in M to be a geodesic whose domain is Œ0; 1/, and such that
the restriction of to Œ0; b is minimizing for every b > 0. Prove that for each
p 2 M there is a ray in M starting at p.
6-22. Let .M; g/ be a connected Riemannian manifold with boundary. Prove that
.M; dg / is a complete metric space if and only if the following condition
holds: for every geodesic W Œ0; b/ ! M that cannot be extended to a geodesic
on any interval Œ0; b 0 / with b 0 > b, .t / converges to a point of @M as t % b.
6-23. In some treatments of Riemannian geometry, instead of minimizing the
length functional, one considers the following energy functional for an ad-
missible curve W Œa; b ! M :
Z b
E. / D 1
2
j 0 .t /j2 dt:
a
Tq0 M and a subsequence .'mj / such that 'mj .pi / ! expq0 ."Abi / for
i D 1; : : : ; n. [Hint: Use Prop. 5.20.]
(c) Prove that there is an isometry ' 2 G such that 'mj ! ' pointwise,
meaning that 'mj .x/ ! '.x/ for every x 2 M .
(d) Prove that 'mj ! ' in the topology of G. [Hint: Define a map F W G !
M nC1 by F . / D . .p0 /; .p1 /; : : : ; .pn //, where p1 ; : : : ; pn are
the points introduced in part (b). Show that F is continuous, injective,
and closed, and therefore is a homeomorphism onto its image.]
(Used on p. 349.)
6-29. Suppose .M; g/ is a Riemannian n-manifold that admits a nonzero parallel
vector field X . Show that for each p 2 M, there exist a neighborhood U of
p, a Riemannian .n 1/-manifold N, and an open interval I R with its
Euclidean metric such that U is isometric to the Riemannian product N I .
[Hint: It might be helpful to prove that the 1-form X [ is closed and to compute
the Lie derivative LX g.]
6-30. Suppose .M; g/ is a Riemannian n-manifold, p 2 M , and B" .p/ is a
geodesic ball centered at p. Prove that for every ı such that 0 < ı < ",
Z
ı
Vol Bxı .p/ D Area.@Br .p// dr;
0
Uı D fx 2 U W dg .x; P / ıg;
Pı D fx 2 U W dg .x; P / D ıg:
In this chapter, we begin our study of the local invariants of Riemannian metrics.
Starting with the question whether all Riemannian metrics are locally isometric,
we are led to a definition of the Riemannian curvature tensor as a measure of the
failure of second covariant derivatives to commute. Then we prove the main result
of this chapter: a Riemannian manifold has zero curvature if and only if it is flat, or
locally isometric to Euclidean space. Next, we derive the basic symmetries of the
curvature tensor, and introduce the Ricci, scalar, and Weyl curvature tensors. At the
end of the chapter, we explore how the curvature can be used to detect conformal
flatness. As you will see, the results of this chapter apply essentially unchanged to
pseudo-Riemannian metrics.
Local Invariants
Fig. 7.1: Result of parallel transport along the x 1 -axis and the x 2 -coordinate lines
On the other hand, Problem 5-5 showed that the round 2-sphere and the Euclidean
plane are not locally isometric.
The most important technique for proving that two geometric structures are not
locally equivalent is to find local invariants, which are quantities that must be pre-
served by local equivalences. In order to address the general problem of local equiv-
alence of Riemannian or pseudo-Riemannian metrics, we will define a local invari-
ant for all such metrics called curvature. Initially, its definition will have nothing to
do with the curvature of curves described in Chapter 1, but later we will see that the
two concepts are intimately related.
To motivate the definition of curvature, let us look back at the argument outlined
in Problem 5-5 for showing that the sphere and the plane are not locally isometric.
The key idea is that every tangent vector in the plane can be extended to a parallel
vector field, so every Riemannian manifold that is locally isometric to R2 must have
the same property locally.
Given a Riemannian 2-manifold M , here is one way to attempt to construct a
parallel extension of a vector z 2 Tp M : working in any smooth local coordinates
.x 1 ; x 2 / centered at p, first parallel transport z along the x 1 -axis, and then parallel
transport the resulting vectors along the coordinate lines parallel to the x 2 -axis (Fig.
7.1). The result is a vector field Z that, by construction, is parallel along every x 2 -
coordinate line and along the x 1 -axis. The question is whether this vector field is
parallel along x 1 -coordinate lines other than the x 1 -axis, or in other words, whether
r@1 Z 0. Observe that r@1 Z vanishes when x 2 D 0. If we could show that
then it would follow that r@1 Z 0, because the zero vector field is the unique
parallel transport of zero along the x 2 -curves. If we knew that
Local Invariants 195
and r x@ r x Z is equal to the same thing, because ordinary second partial deriva-
1 @2
tives commute. However, (7.2) might not hold for an arbitrary Riemannian metric;
indeed, it is precisely the noncommutativity of such second covariant derivatives
that forces this construction to fail on the sphere. Lurking behind this noncommuta-
tivity is the fact that the sphere is “curved.”
To express this noncommutativity in a coordinate-independent way, let us look
more closely at the quantity r xX r xY r
xY Z r x X Z when X , Y , and Z are smooth
vector fields. On R2 with the Euclidean connection, we just showed that this always
vanishes if X D @1 and Y D @2 ; however, for arbitrary vector fields this may no
longer be true. In fact, in Rn with the Euclidean connection we have
xX r
r xY Z D r x X Y Z k @k D X Y Z k @k ;
and similarly rx r x Z D YX Z k @k . The difference between these two expressions
k Y X k
is X Y Z YX Z @k D r x ŒX;Y Z. Therefore, the following relation holds for
all vector fields X; Y; Z defined on an open subset of Rn :
xX r
r xY r
xY Z r x ŒX;Y Z:
xX Z D r
rX rY Z rY rX Z D rŒX;Y Z: (7.3)
Example 7.1. The metric on the n-torus induced by the embedding in R2n given in
Example 2.21 is flat, because each point has a coordinate neighborhood in which
the metric is Euclidean. //
Proof. We just showed that the Euclidean connection on Rn satisfies (7.3). By nat-
urality, the Levi-Civita connection on every manifold that is locally isometric to a
Euclidean or pseudo-Euclidean space must also satisfy the same identity. t
u
R.X; Y /Z D rX rY Z rY rX Z rŒX;Y Z:
Proposition 7.3. The map R defined above is multilinear over C 1 .M /, and thus
defines a .1; 3/-tensor field on M .
Proof. The map R is obviously multilinear over R. For f 2 C 1 .M /,
R.X; f Y /Z D rX rf Y Z rf Y rX Z rŒX;f Y Z
D rX .f rY Z/ f rY rX Z rf ŒX;Y C.Xf /Y Z
D .Xf /rY Z C f rX rY Z f rY rX Z
f rŒX;Y Z .Xf /rY Z
D fR.X; Y /Z:
R D Rij k l dx i ˝ dx j ˝ dx k ˝ @l ;
Proof. This is a local question, so for each .s; t / 2 J I , we can choose smooth
coordinates x i defined on a neighborhood of .s; t / and write
ˇ
.s; t / D 1 .s; t /; : : : ; n .s; t / ; V .s; t / D V j .s; t /@j ˇ .s;t/ :
@2 V i @V i @V i
Ds Dt V D @i C Ds @i C Dt @i C V i Ds Dt @i :
@s@t @t @s
Interchanging s and t and subtracting, we see that all the terms except the last cancel:
Ds Dt V Dt Ds V D V i .Ds Dt @i Dt Ds @i / : (7.6)
Now we need to compute the commutator in parentheses. For brevity, let us write
@ k @ j
S D @s D @k I T D @t D @j :
@s @t
Because @i is extendible,
@ j
Dt @i D rT @i D r@j @i ;
@t
198 7 Curvature
@ j @ k
Ds Dt @i Dt Ds @i D r@k r@j @i r@j r@k @i
@t @s
@ j @ k
D R.@k ; @j /@i D R.S; T /@i :
@t @s
Finally, inserting this into (7.6) yields the result. t
u
For many purposes, the information contained in the curvature endomorphism is
much more conveniently encoded in the form of a covariant 4-tensor. We define the
(Riemann) curvature tensor to be the .0; 4/-tensor field Rm D R[ (also denoted by
Riem by some authors) obtained from the .1; 3/-curvature tensor R by lowering its
last index. Its action on vector fields is given by
Rm D Rij kl dx i ˝ dx j ˝ dx k ˝ dx l ;
It is appropriate to note here that there is much variation in the literature with
respect to index positions in the definitions of the curvature endomorphism and
curvature tensor. While almost all authors define the .1; 3/-curvature tensor as we
have, there are a few (notably [dC92, GHL04]) whose definition is the negative of
ours. There is much less agreement on the definition of the .0; 4/-curvature tensor:
whichever definition is chosen for the curvature endomorphism, you will see the
curvature tensor defined as in (7.7) but with various permutations of .X; Y; Z; W /
on the right-hand side. After applying the symmetries of the curvature tensor that we
will prove later in this chapter, however, all of the definitions agree up to sign. There
are various arguments to support one choice or another; we have made a choice that
makes equation (7.7) easy to remember. You just have to be careful when you begin
reading any book or article to determine the author’s sign convention.
Flat Manifolds 199
The next proposition gives one reason for our interest in the curvature tensor.
Proposition
7.6. The curvature tensor is a local isometry invariant: if .M; g/ and
M ; gz are Riemannian or pseudo-Riemannian manifolds and ' W M ! M is a
f D Rm.
local isometry, then ' Rm
I Exercise 7.7. Prove Proposition 7.6.
Flat Manifolds
To give a qualitative geometric interpretation to the curvature tensor, we will show
that it is precisely the obstruction to being locally isometric to Euclidean (or pseudo-
Euclidean) space. (In Chapter 8, after we have developed more machinery, we will
be able to give a far more detailed quantitative interpretation.) The crux of the proof
is the following lemma.
This shows that r@i V is parallel along the x kC1 -curves starting on Mk . Since r@i V
vanishes on Mk and the zero vector field is the unique parallel transport of zero, we
conclude that r@i V is zero on each x kC1 -curve. Since every point of MkC1 is on
one of these curves, it follows that r@i V D 0 on all of MkC1 . This completes the
inductive step to show that V is parallel. t
u
I Exercise 7.9. Prove that the vector field V constructed in the preceding proof is
smooth.
Fig. 7.3: The curvature endomorphism and parallel transport around a closed loop
Thus the vector fields .E1 ; : : : ; En / form a commuting orthonormal frame on U . The
canonical form theorem
forcommuting vector fields (Prop. A.48) shows that there
are coordinates y 1 ; : : : ; y n on a (possibly smaller) neighborhood of p such that
Ei D @=@y i for i D 1; : : : ; n. In any such coordinates, gij D g.@i ; @j / D g.Ei ; Ej / D
˙ıij , so the map y D .y 1 ; : : : ; y n / is an isometry from a neighborhood of p to an
open subset of the appropriate Euclidean or pseudo-Euclidean space. t
u
Using similar ideas, we can give a more precise interpretation of the meaning of
the curvature tensor: it is a measure of the extent to which parallel transport around
a small rectangle fails to be the identity map.
Proof. Define a vector field Z along by first parallel transporting z along the
curve t 7! .0; t /, and then for each t , parallel transporting Z.0; t / along the curve
s 7! .s; t /. The resulting vector field along is smooth by another application
of Theorem 4.31; and by construction, it satisfies Dt Z.0; t / D 0 for all t 2 I , and
Ds Z.s; t / D 0 for all .s; t / 2 I I . Proposition 7.5 shows that
Thus we need only show that Ds Dt Z.0; 0/ is equal to the limit on the right-hand
side of (7.10).
From Theorem 4.34, we have
s;0
Ps;" .Z.s; "// Z.s; 0/
Dt Z.s; 0/ D lim ; (7.11)
"!0 "
0;0
Pı;0 .Dt Z.ı; 0// Dt Z.0; 0/
Ds Dt Z.0; 0/ D lim : (7.12)
ı!0 ı
Evaluating (7.11) first at s D ı and then at s D 0, and inserting the resulting expres-
sions into (7.12), we obtain
Ds Dt Z.0; 0/
0;0 ı;0 0;0 0;0
Pı;0 ı Pı;" .Z.ı; "// Pı;0 .Z.ı; 0// P0;" .Z.0; "// C Z.0; 0/
D lim : (7.13)
ı;"!0 ı"
Here we have used the fact that parallel transport is linear, so the "-limit can be
0;0
pulled past Pı;0 .
Now, the fact that Z is parallel along t 7! .0; t / and along all of the curves
s 7! .s; t / implies
0;0 0;0
Pı;0 .Z.ı; 0// D P0;" .Z.0; "// D Z.0; 0/ D z;
ı;" ı;" 0;"
Z.ı; "/ D P0;" .Z.0; "// D P0;" ı P0;0 .z/:
Before we begin the proof, a few remarks are in order. First, as the proof will
show, (a) is a trivial consequence of the definition of the curvature endomorphism;
(b) follows from the compatibility of the Levi-Civita connection with the metric; (d)
Symmetries of the Curvature Tensor 203
follows from the symmetry of the connection; and (c) follows from (a), (b), and (d).
The identity in (d) is called the algebraic Bianchi identity (or more traditionally
but less informatively, the first Bianchi identity). It is easy to show using (a)–(d)
that a three-term sum obtained by cyclically permuting any three arguments of Rm
is also zero. Finally, it is useful to record the form of these symmetries in terms of
components with respect to any basis:
(a0 ) Rij kl D Rj ikl .
(b0 ) Rij kl D Rij lk .
(c0 ) Rij kl D Rklij .
(d0 ) Rij kl C Rj kil C Rkij l D 0.
Proof of Proposition 7.12. Identity (a) is immediate from the definition of the
curvature tensor, because R.W; X /Y D R.X; W /Y . To prove (b), it suffices to
show that Rm.W; X; Y; Y / D 0 for all Y , for then (b) follows from the expansion of
Rm.W; X; Y C Z; Y C Z/ D 0. Using compatibility with the metric, we have
When we subtract the second and third equations from the first, the left-hand side
is zero. The terms 2hrX Y; rW Y i and 2hrW Y; rX Y i cancel on the right-hand side,
giving
Next we prove (d). From the definition of Rm, this will follow immediately from
Using the definition of R and the symmetry of the connection, the left-hand side
expands to
.rW rX Y rX rW Y rŒW;X Y /
C .rX rY W rY rX W rŒX;Y W /
C .rY rW X rW rY X rŒY;W X /
D rW .rX Y rY X / C rX .rY W rW Y / C rY .rW X rX W /
rŒW;X Y rŒX;Y W rŒY;W X
D rW ŒX; Y C rX ŒY; W C rY ŒW; X
rŒW;X Y rŒX;Y W rŒY;W X
D ŒW; ŒX; Y C ŒX; ŒY; W C ŒY; ŒW; X :
204 7 Curvature
Now add up all four equations. Applying (b) four times makes all the terms in
the first two columns cancel. Then applying (a) and (b) in the last column yields
2Rm.Y; W; X; Z/ 2Rm.X; Z; Y; W / D 0, which is equivalent to (c). t
u
There is one more identity that is satisfied by the covariant derivatives of the
curvature tensor on every Riemannian manifold. Classically, it was called the sec-
ond Bianchi identity, but modern authors tend to use the more informative name
differential Bianchi identity.
In components, this is
This can be proved by a long and tedious computation, but there is a standard short-
cut for such calculations in Riemannian geometry that makes our task immeasurably
easier. To prove that (7.16) holds at a particular point p, it suffices by multilinear-
ity to prove the formula when X; Y; Z; V; W are basis vectors with respect to some
frame. The shortcut consists in choosing a special frame for each point p to simplify
the computations there.
Let p be an arbitrary point, let x i be normal coordinates centered at p, and let
X; Y; Z; V; W be arbitrary coordinate basis vector fields. These vectors satisfy two
properties that simplify our computations enormously: (1) their commutators vanish
identically, since Œ@i ; @j 0; and (2) their covariant derivatives vanish at p, since
ijk .p/ D 0 (Prop. 5.24(d)).
Using these facts and the compatibility of the connection with the metric, the first
term in (7.16) evaluated at p becomes
The Ricci Identities 205
.rW Rm/.Z; V; X; Y / D rW Rm.Z; V; X; Y /
D rW hR.Z; V /X; Y i
D rW hrZ rV X rV rZ X rŒZ;V X; Y i
D hrW rZ rV X rW rV rZ X; Y i:
Write this equation three times, with the vector fields W; Z; V cyclically permuted.
Summing all three gives
If ˇ is a smooth 1-form,
2 2
rX;Y ˇ rY;X ˇ D R.X; Y / ˇ: (7.18)
2 2
rX;Y B rY;X B .! 1 ; : : : ; ! k ; V1 ; : : : ; Vl /
D B R.X; Y / ! 1 ; ! 2 ; : : : ; ! k ; V1 ; : : : ; Vl C
C B ! 1 ; : : : ; ! k1 ; R.X; Y / ! k ; V1 ; : : : ; Vl (7.19)
B ! 1 ; : : : ; ! k ; R.X; Y /V1 ; V2 ; : : : ; Vl
B ! 1 ; : : : ; ! k ; V1 ; : : : ; Vl1 ; R.X; Y /Vl ;
for all covector fields ! i and vector fields Vj . In terms of any local frame, the
component versions of these formulas read
Proof. For any tensor field B and vector fields X; Y , Proposition 4.21 implies
2 2
rX;Y B rY;X B D rX rY B r.rX Y / B rY rX B C r.rY X/ B
(7.23)
D rX rY B rY rX B rŒX;Y B;
where the last equality follows from the symmetry of the connection. In particular,
this holds when B D Z is a vector field, so (7.17) follows directly from the definition
of the curvature endomorphism.
Next we prove (7.18). Using (4.13) repeatedly, we compute
.rX rY ˇ/.Z/ D X .rY ˇ/.Z/ .rY ˇ/.rX Z/
D X Y.ˇ.Z// ˇ.rY Z/ .rY ˇ/.rX Z/
D X Y.ˇ.Z// .rX ˇ/.rY Z/ ˇ.rX rY Z/ .rY ˇ/.rX Z/:
(7.24)
Reversing the roles of X and Y , we get
Now subtract (7.25) and (7.26) from (7.24): all but three of the terms cancel, yielding
rX rY ˇ rY rX ˇ rŒX;Y ˇ .Z/ D ˇ rX rY Z rY rX Z rŒX;Y Z
D ˇ R.X; Y /Z ;
A simple induction using this relation together with (7.17) and (7.18) shows that for
all smooth vector fields W1 ; : : : ; Wk and 1-forms 1 ; : : : ; l ,
2 2
rX;Y rY;X W1 ˝ ˝ Wk ˝ 1 ˝ ˝ l
D R.X; Y /W1 ˝ W2 ˝ ˝ Wk ˝ 1 ˝ ˝ l C
C W1 ˝ ˝ Wk1 ˝ R.X; Y /Wk ˝ 1 ˝ ˝ l
C W1 ˝ ˝ Wk ˝ R.X; Y / 1 ˝ 2 ˝ ˝ l C
C W1 ˝ ˝ Wk ˝ 1 ˝ ˝ l1 ˝ R.X; Y / l :
Since every tensor field can be written as a sum of tensor products of vector fields
and 1-forms, this implies (7.19).
Finally, the component formula (7.22) follows by applying (7.19) to
2 2
rEq ;Ep B rE p ;Eq
B "i1 ; : : : ; "ik ; Ej1 ; : : : ; Ejl ;
where .Ej / and ."i / represent a local frame and its dual coframe, respectively, and
using
The scalar curvature is the function S defined as the trace of the Ricci tensor:
S D trg Rc D Ri i D g ij Rij :
It is probably not clear at this point why the Ricci tensor or scalar curvature might be
interesting, and we do not yet have the tools to give them geometric interpretations.
But be assured that there is such an interpretation; see Proposition 8.32.
Lemma 7.15. The Ricci curvature is a symmetric 2-tensor field. It can be expressed
in any of the following ways:
I Exercise 7.16. Prove Lemma 7.15, using the symmetries of the curvature tensor.
It is sometimes useful to decompose the Ricci tensor into a multiple of the metric
and a complementary piece with zero trace. Define the traceless Ricci tensor of g
as the following symmetric 2-tensor:
1
V D Rc Sg:
Rc
n
Proposition 7.17. Let .M; g/ be a Riemannian or pseudo-Riemannian n-manifold.
V 0, and the Ricci tensor decomposes orthogonally as
Then trg Rc
1
V C Sg:
Rc D Rc (7.27)
n
Therefore, in the Riemannian case,
ˇ ˇ2 1 2
Vˇ C S :
jRcj2g D ˇRc g
(7.28)
n
Remark. The statement about norms, and others like it that we will prove below,
works only in the Riemannian case because of the additional absolute value signs
required to compute norms in the˝ pseudo-Riemannian
˛ case. The pseudo-Riemannian
analogue would be hRc; Rcig D Rc; V Rc
V C 1 S 2 , but this is not as useful.
g n
V that trg Rc
It then follows directly from the definition of Rc V 0 and (7.27) holds.
The fact that the decomposition is orthogonal follows easily from the fact that for
every symmetric 2-tensor h, we have
Ricci and Scalar Curvatures 209
where the trace in each case is on the first and last indices. In components, this is
Proof. Start with the component form (7.15) of the differential Bianchi identity,
raise the index m, and then contract on the indices i; m to obtain (7.32). (Note that
covariant differentiation commutes with contraction by Proposition 4.15 and with
the musical isomorphisms by Proposition 5.17, so it does not matter whether the
indices that are raised and contracted come before or after the semicolon.) Then do
the same with the indices j; k and simplify to obtain (7.33). The coordinate-free
formulas (7.30) and (7.31) follow by expanding everything out in components. u t
It is important to note that if the sign convention chosen for the curvature tensor
is the opposite of ours, then the Ricci tensor must be defined as the trace of Rm on
the first and third (or second and fourth) indices. (The trace on the first two or last
210 7 Curvature
two indices is always zero by antisymmetry.) The definition is chosen so that the
Ricci and scalar curvatures have the same meaning for everyone, regardless of the
conventions chosen for the full curvature tensor. So, for example, if a manifold is
said to have positive scalar curvature, there is no ambiguity as to what is meant.
A Riemannian or pseudo-Riemannian metric is said to be an Einstein metric if
its Ricci tensor is a constant multiple of the metric—that is,
This equation is known as the Einstein equation. As the next proposition shows,
for connected manifolds of dimension greater than 2, it is not necessary to assume
that is constant; just assuming that the Ricci tensor is a function times the metric
is sufficient.
Proof. Suppose first that g is an Einstein metric with Rc D g. Taking traces of both
V D Rc g D 0. Conversely, if Rc
sides, we find that D n1 S , and therefore Rc V D 0,
Schur’s lemma implies that g is Einstein. t
u
By an argument analogous to those of Chapter 6, Hilbert showed (see [Bes87,
Thm. 4.21]) that Einstein
R metrics are critical points for the total scalar curvature
functional .g/ D M S d Vg on the space of all metrics on M with fixed volume.
Thus Einstein metrics can be viewed as “optimal” metrics in a certain sense, and as
such they form an appealing higher-dimensional analogue of locally homogeneous
metrics on 2-manifolds, with which one might hope to prove some sort of general-
ization of the uniformization theorem (Thm. 3.22). Although the statement of such
a theorem cannot be as elegant as that of its 2-dimensional ancestor because there
Ricci and Scalar Curvatures 211
are known examples of smooth, compact manifolds that admit no Einstein metrics
[Bes87, Chap. 6], there is still a reasonable hope that “most” higher-dimensional
manifolds (in some sense) admit Einstein metrics. This is an active field of current
research; see [Bes87] for a sweeping survey of Einstein metrics.
The term “Einstein metric” originated, as you might guess, in physics: the central
assertion of Einstein’s general theory of relativity is that physical spacetime is mod-
eled by a 4-manifold that carries a Lorentz metric whose Ricci curvature satisfies
the following Einstein field equation:
1
Rc Sg D T; (7.35)
2
where T is a certain symmetric 2-tensor field (the stress–energy tensor) that
describes the density, momentum, and stress of the matter and energy present at
each point in spacetime. It is shown in physics books (e.g., [CB09, pp. 51–53]) that
(7.35) is the variational equation for a certain functional, called the Einstein–Hilbert
action, on the space of all Lorentz metrics on a given 4-manifold. Einstein’s theory
can then be interpreted as the assertion that a physically realistic spacetime must be
a critical point for this functional.
In the special case T 0, (7.35) reduces to the vacuum Einstein field
equation Rc D 12 Sg. Taking traces of both sides and recalling that trg g D dim M D
4, we obtain S D 2S , which implies S D 0. Therefore, the vacuum Einstein equation
is equivalent to Rc D 0, which means that g is a (pseudo-Riemannian) Einstein met-
ric in the mathematical sense of the word. (At one point in the development of the
theory, Einstein considered adding a term g to the left-hand side of (7.35), where
is a constant that he called the cosmological constant. With this modification the
vacuum Einstein field equation would be exactly the same as the mathematicians’
Einstein equation (7.34). Einstein soon decided that the cosmological constant
was a mistake on physical grounds; however, researchers in general relativity have
recently begun to believe that a theory with a nonzero cosmological constant might in
fact have physical relevance.)
Other than these special cases and the obvious formal similarity between (7.35)
and (7.34), there is no direct connection between the physicists’ version of the
Einstein equation and the mathematicians’ version. The mathematical interest in
Riemannian Einstein metrics stems more from their potential applications to uni-
formization in higher dimensions than from their relation to physics.
Another approach to generalizing the uniformization theorem to higher dimen-
sions is to search for metrics of constant scalar curvature. These are also critical
points of the total scalar curvature functional, but only with respect to variations of
the metric with fixed volume within a given conformal equivalence class. Thus it
makes sense to ask whether, given a metric g on a manifold M , there exists a metric
gz conformal to g that has constant scalar curvature. This is called the Yamabe prob-
lem, because it was first posed in 1960 by Hidehiko Yamabe, who claimed to have
proved that the answer is always yes when M is compact. Yamabe’s proof was
later found to be in error, and it was two dozen years before the proof was finally
completed by Richard Schoen; see [LP87] for an expository account of Schoen’s
212 7 Curvature
solution. When M is noncompact, the issues are much subtler, and much current
research is focused on determining exactly which conformal classes contain metrics
of constant scalar curvature.
It is easy to check that ˚.B/ satisfies (a)–(c), so ˚.B/ 2 B.V /, and that ˚ is a
linear map. In fact, it is an isomorphism, which we prove by constructing an inverse
for it. Choose a basis .b1 ; : : : ; bn / for V , so the fbi ^ bj W i < j g is a basis
collection
for ƒ2 .V /. Define a map W B.V / ! †2 ƒ2 .V / by setting
.T /.bi ^ bj ; bk ^ bl / D T .bi ; bj ; bk ; bl /
(It should be noted that the Kulkarni–Nomizu product must be defined as the nega-
tive of this expression when the Riemann curvature tensor is defined as the negative
of ours.)
Lemma 7.22 (Properties of the Kulkarni–Nomizu Product). Let V be an n-
dimensional vector space endowed with a scalar product g, let h and k be sym-
metric 2-tensors on V , let T be an algebraic curvature tensor on V , and let trg
denote the trace on the first and last indices.
(a) h k is an algebraic curvature tensor.
(b) h k D k h.
(c) trg .h g/ D .n 2/h C .trg h/g.
(d) trg .g g/ D 2.n 1/g.
(e) hT; h gig D 4htrg T; hig .
(f) In case g is positive definite, jg hj2g D 4.n 2/jhj2g C 4.trg h/2 .
Proof. It is evident from the definition that h k has three of the four symmetries of
an algebraic curvature tensor: it is antisymmetric in its first two arguments and also
in its last two, and its value is unchanged when the first two and last two arguments
are interchanged. Thus to prove (a), only the algebraic Bianchi identity needs to
be checked. This is a straightforward computation: when h k.w; x; y; z/ is written
three times with the arguments w; x; y cyclically permuted and the three expressions
are added together, all the terms cancel due to the symmetry of h and k.
Part (b) is immediate from the definition. To prove (c), choose a basis and use the
definition to compute
.trg .h g//j k D g im him gj l C hj l gim hil gj m hj m gil
D hi i gj l C nhj l hj l hj l ;
which is equivalent to (c). Then (d) follows from (c) and the fact that trg g D n.
The proofs of (e) and (f) are left to Problem 7-9. t
u
Here is the primary application of the Kulkarni–Nomizu product.
Proposition 7.23. Let .V; g/ be an n-dimensional scalar product space with n 3,
and define a linear map G W †2 .V / ! R.V / by
1 trg h
G.h/ D h g g: (7.38)
n2 2.n 1/
Then G is a right inverse for trg , and its image is the orthogonal complement of the
kernel of trg in R.V /.
Proof. The fact that G is a right inverse is a straightforward computation based
on the definition and Lemma 7.22(c,d). This implies that G is injective and trg
is surjective, so the dimension of Im G is equal to the codimension of Ker.trg /,
which in turn is equal to the dimension of Ker.trg /? . If T 2 R.V / is an algebraic
curvature tensor such that trg T D 0, then Lemma 7.22(e) shows that hT; G.h/i D 0,
so it follows by dimensionality that Im G D Ker.trg /? . t
u
The Weyl Tensor 215
and define the Weyl tensor of g to be the following algebraic curvature tensor field:
W D Rm P g
1 S
D Rm Rc g C g g:
n2 2.n 1/.n 2/
Rm D P g D Rc g 14 Sg g: (7.39)
Rm D 14 Sg g; Rc D 12 Sg:
Proof. In dimension 2, it follows from Corollary 7.25(b) that there is some scalar
function f such that Rm D fg g. Taking traces, we find from Lemma 7.22(d) that
Rc D trg .Rm/ D 2fg, and then S D trg .Rc/ D 2f trg .g/ D 4f . The results follow
by substituting f D 14 S back into these equations. t
u
Although the traceless Ricci tensor is always zero on a 2-manifold, this does not
imply that S is constant, since the proof of Schur’s lemma fails in dimension 2.
Einstein metrics in dimension 2 are simply those with constant scalar curvature.
Returning now to dimensions greater than 2, we can use (7.27) to further decom-
pose the Schouten tensor into a part determined by the traceless Ricci tensor and a
purely scalar part. The next proposition is the analogue of Proposition 7.17 for the
full curvature tensor.
Proposition 7.28 (The Ricci Decomposition of the Curvature Tensor). Let .M; g/
be a Riemannian or pseudo-Riemannian manifold of dimension n 3. Then the
.0; 4/-curvature tensor of g has the following orthogonal decomposition:
1 1
Rm D W C V gC
Rc S g g: (7.40)
n2 2n.n 1/
Therefore, in the Riemannian case,
1 ˇ ˇ 1
jRmj2g D jW j2g C V g ˇ2 C
ˇRc jS g gj2g
.n 2/2 g 4n2 .n 1/2
(7.41)
4 ˇˇ ˇˇ2 2
D jW j2g C V C
Rc S 2:
n2 g n.n 1/
Proof. The decomposition (7.40) follows immediately by substituting (7.27) into
the definition of the Weyl tensor and simplifying. The decomposition is orthogonal
thanks to Lemma 7.22(e), and (7.41) follows from Lemma 7.22(f). t
u
But it is a remarkable fact that the Weyl tensor has a very simple transformation law
under conformal changes of metric. In this section, we derive that law.
First we need to determine how the Levi-Civita connection changes when a met-
ric is changed conformally. Given conformal metrics g and gz, we can always write
gz D e 2f g for some smooth real-valued function f .
where the gradient on the right-hand side is that of g. In any local coordinates, the
Christoffel symbols of the two connections are related by
zijk D ijk C fIi ıjk C fIj ıik g kl fIl gij ;
(7.43)
Proof. Formula (7.43) is a straightforward computation using formula (5.10) for the
Christoffel symbols in coordinates, and then (7.42) follows by expanding everything
in coordinates and using (7.43). t
u
This result leads to transformation laws for the various curvature tensors.
where the curvatures and covariant derivatives on the right are those of g, and
f D div.grad f / is the Laplacian of f defined in Chapter 2. If in addition n 3,
then
In the pseudo-Riemannian case, the same formulas hold with each occurrence of
jdf j2g replaced by hdf; df i2g .
218 7 Curvature
Proof. We begin with (7.44). The plan is to choose local coordinates and insert for-
mula (7.43) for the Christoffel symbols into the coordinate formula (7.8) for the co-
efficients of the curvature tensor. As in the proof of the differential Bianchi identity,
we can make the computations much more tractable by computing the components
of these tensors at a point p 2 M in normal coordinates for g centered at p, so that
the equations gij D ıij , @k gij D 0, and ijk D 0 hold at p. This has the following
consequences at p:
fIij D @j @i f;
zijk D fIi ıjk C fIj ıik g kl fIl gij ;
zijk D @m ijk C fIim ıjk C fIj m ıik g kl fIlm gij ;
@m
Rij k l D @i jl k @j ik
l
:
C .fIi fIl gj k C fIj fIk gil fIi fIk gj l fIj fIl gik /
g mp fIm fIp .gil gj k gik gj l / ;
which is the coordinate version of (7.44). (See Problem 5-14 for the formula for the
Laplacian in terms of covariant derivatives.) The rest of the formulas follow easily
from this, using the identities of Lemma 7.22 and the fact that g ij D e 2f gzij . u
t
The next corollary begins to explain the geometric significance of the Weyl ten-
sor. Recall that a Riemannian manifold is said to be locally conformally flat if
every point has a neighborhood that is conformally equivalent to an open subset of
Euclidean space. Similarly, a pseudo-Riemannian manifold is locally conformally
flat if every point has a neighborhood conformally equivalent to an open set in a
pseudo-Euclidean space.
Proof. Suppose .M; g/ is locally conformally flat. Then for each p 2 M there exist
a neighborhood U and an embedding ' W U ! Rn such that ' pulls back a flat
(Riemannian or pseudo-Riemannian) metric on Rn to a metric of the form gz D
e 2f g. This implies that gz has zero Weyl tensor, because its entire curvature tensor
is zero. By virtue of (7.48), the Weyl tensor of g is also zero. t
u
Curvatures of Conformally Related Metrics 219
Proof. Writing W D Rm P g and using the component form of the first con-
tracted Bianchi identity (7.32), we obtain
Note that Pj kI i gil D Pj kIl and Pj lI i gik D Pj lIk . To simplify the other two terms,
we use the definition of the Schouten tensor and the second contracted Bianchi
identity (7.33) to obtain that
i 1 i S Ii 1
PilI D RilI gil D SIl :
n2 2.n 1/ 2.n 1/
The analogous formula holds for PikI i . When we insert these expressions into (7.51)
and simplify, we find that the terms involving derivatives of the Ricci and scalar
curvatures combine to yield
The next proposition expresses another important feature of the Cotton tensor.
Corollary 7.35. If .M; g/ is a locally conformally flat 3-manifold, then the Cotton
tensor of g vanishes identically.
The real significance of the Weyl and Cotton tensors is explained by the following
important theorem.
Proof. The necessity of each condition was proved in Corollaries 7.31 and 7.35. To
prove sufficiency, suppose .M; g/ satisfies the hypothesis appropriate to its dimen-
sion; then it follows from Corollaries 7.26 and 7.33 that the Weyl and Cotton tensors
of g are both identically zero. Every metric gz D e 2f g conformal to g also has zero
f D Pz gz. We will show that in
Weyl tensor, and therefore its curvature tensor is Rm
a neighborhood of each point, the function f can be chosen to make Pz 0, which
implies that Rmf 0 and therefore gz is flat.
From (7.47), it follows that Pz 0 if and only if
This equation can be written in the form r.df / D A.df /, where A is the map from
1-tensors to symmetric 2-tensors given by
A./ D . ˝ / 12 h; i2g g C P;
or in components,
A./ij D i j 12 m m gij C Pij : (7.53)
We will solve this equation by first looking for a smooth 1-form that satisfies
r D A./. In any local coordinates, if we write D j dx j , this becomes an
overdetermined system of first-order partial differential equations for the n unknown
functions 1 ; : : : ; n . (A system of differential equations is said to be overdetermined
if there are more equations than unknown functions.) If is a solution to r D A./
on an open subset of M , then the Ricci identity (7.21) shows that A./ satisfies
Lemma 7.38 below shows that this condition is actually sufficient: more precisely,
r D A./ has a smooth solution in a neighborhood of each point, provided that for
Curvatures of Conformally Related Metrics 221
Aij Ik AikIj
D iIk j C i j Ik mIk m gij iIj k i kIj C mIj m gik C Cij k :
Now substitute the right-hand side of (7.53) (with appropriate index substitutions)
for iIj wherever it appears, subtract Rj ki l l , and use the relation Rm D W CP g.
After extensive cancellations, we obtain
Our hypotheses guarantee that the right-hand side is identically zero, so there is a
solution to r D A./ in a neighborhood of each point.
Because A./ij is symmetric in i and j , it follows that the ordinary deriva-
tives @j i D iIj C ijk k are also symmetric, and thus is a closed 1-form. By
the Poincaré lemma, in some (possibly smaller) neighborhood of each point, there
is a smooth function f such that D df D rf ; this f is the function we seek. u t
Here is the lemma that was used in the proof of the preceding theorem.
r D A./; (7.56)
Then for every p 2 M and every covector 0 2 Tp M , there is a smooth solution to
(7.56) on a neighborhood of p satisfying p D 0 .
Proof. Let p 2 M be given. In smooth local coordinates x i on a neighborhood of
p, (7.56) is equivalent to the overdetermined system
@i .x/
D aij .x; .x//;
@x j
where
aij .x; / D ijk .x/k C A./ij :
An application of the Frobenius theorem [LeeSM, Prop. 19.29] shows that there is
a smooth solution to this overdetermined system in a neighborhood of p with p
arbitrary, provided that
222 7 Curvature
Problems
7-1. Complete the proof of Proposition 7.3 by showing that R.X; Y /.f Z/ D
fR.X; Y /Z for all smooth vector fields X; Y; Z and smooth real-valued
functions f .
7-2. Prove Proposition 7.4 (the formula for the curvature tensor in coordinates).
7-3. Show that the curvature tensor of a Riemannian locally symmetric space is
parallel: rRm 0. (Used on pp. 297, 351.)
7-4. Let M be a Riemannian or pseudo-Riemannian manifold, and let x i be
normal coordinates centered at p 2 M . Show that the following holds at p:
1
Rij kl D @j @l gik C @i @k gj l @i @l gj k @j @k gil :
2
7-5. Let r be the Levi-Civita connection on a Riemannian or pseudo-Riemannian
manifold .M; g/, and let !i j be its connection 1-forms with respect to a local
frame .Ei / (Problem 4-14). Define a matrix of 2-forms
i j , called the cur-
vature 2-forms, by
1
i j D Rkli j "k ^ "l ;
2
i
where " is the coframe dual to .Ei /. Show that the curvature 2-forms
satisfy Cartan’s second structure equation:
Problems 223
i j D d!i j !i k ^ !k j :
Rm D
1 Rm1 C
2 Rm2 ;
Rc D
1 Rc1 C
2 Rc2 ;
S D
1 S1 C
1 S2 ;
where Rmi , Rci , and Si are the Riemann, Ricci, and scalar curvatures of
.Mi ; gi /, and
i W M1 M2 ! Mi is the projection. (Used on pp. 257, 261.)
7-7. Suppose .M; g/ is a Riemannian manifold and u 2 C 1 .M /. Use (5.29) and
(7.21) to prove Bochner’s formula:
1 ˇ ˇ2 ˝ ˛
2
j grad uj2 D ˇr 2 uˇ C grad.u/; grad u C Rc .grad u; grad u/ :
4.n 1/
Lu D u C S u;
.n 2/
where is the Laplace–Beltrami operator of g and S is its scalar curvature.
Prove that if gz D e 2f g for some f 2 C 1 .M /, and Lz denotes the conformal
Laplacian with respect to gz, then for every u 2 C 1 .M /,
nC2
z D L e n2
e 2 f Lu 2 fu :
Rm.W; X; Y; Z/ı
D Rm W e
; X;
z 1 W
z Yz ; Z
2
D
; Xz V ; Yz ; Z
E
z V
1 D z
V z z
V E 1 D z
V z z
V E
W ; Y ; X; Z C W ; Z ; X; Y :
4 4
(Used on p. 258.)
7-15. Suppose .M; g/ is a 2-dimensional pseudo-Riemannian manifold of signature
.1; 1/, and p 2 M .
(a) Show that there is a smooth local frame .E1 ; E2 / in a neighborhood of
p such that g.E1 ; E1 / D g.E2 ; E2 / D 0.
(b) Show that there are smooth coordinates .x; y/ in a neighborhood of
p such that .dx/2 .dy/2 D fg for some smooth, positive real-valued
function f . [Hint: Use Prop. A.45 to show that there exist coordinates
.t; u/ in which E1 D @=@t , and coordinates .v; w/ in which E2 D @=@v,
and set x D u C w, y D u w.]
(c) Show that .M; g/ is locally conformally flat.
Chapter 8
Riemannian Submanifolds
This chapter has a dual purpose: first to apply the theory of curvature to Riemann-
ian submanifolds, and then to use these concepts to derive a precise quantitative
interpretation of the curvature tensor.
After introducing some basic definitions and terminology concerning Riemann-
ian submanifolds, we define a vector-valued bilinear form called the second fun-
damental form, which measures the way a submanifold curves within the ambient
manifold. We then prove the fundamental relationships between the intrinsic and
extrinsic geometries of a submanifold, including the Gauss formula, which relates
the Riemannian connection on the submanifold to that of the ambient manifold, and
the Gauss equation, which relates their curvatures. We then show how the second
fundamental form measures the extrinsic curvature of submanifold geodesics.
Using these tools, we focus on the special case of hypersurfaces, and use the
second fundamental form to define some real-valued curvature quantities called the
principal curvatures, mean curvature, and Gaussian curvature. Specializing even
more to hypersurfaces in Euclidean space, we describe various concrete geomet-
ric interpretations of these quantities. Then we prove Gauss’s Theorema Egregium,
which shows that the Gaussian curvature of a surface in R3 can be computed intrin-
sically from the curvature tensor of the induced metric.
In the last section, we introduce the promised quantitative geometric interpreta-
tion of the curvature tensor. It allows us to compute sectional curvatures, which are
just the Gaussian curvatures of 2-dimensional submanifolds swept out by geodesics
tangent to 2-planes in the tangent space. Finally, we compute the sectional curva-
tures of our frame-homogeneous model Riemannian manifolds—Euclidean spaces,
spheres, and hyperbolic spaces.
jM ! TM;
> W T M
?
W TM jM ! NM :
so both projections are smooth bundle homomorphisms (i.e., they are linear on fibers
and map smooth sections to smooth sections). If X is a section of T M jM , we often
use the shorthand notations X > D > X and X ? D ? X for its tangential and
normal projections.
If X; Y are vector fields in X.M /, we can extend them to vector fields on an
open subset of M (still denoted by X and Y ), apply the ambient covariant derivative
z and then decompose at points of M to get
operator r,
zX Y D r
r zX Y > C rzX Y ?: (8.1)
We wish to interpret the two terms on the right-hand side of this decomposition.
Let us focus first on the normal component. We define the second fundamental
form of M to be the map II W X.M / X.M / ! .NM / (read “two”) given by
II.X; Y / D rzX Y ?;
Since X and Y are tangent to M at all points of M , so is their Lie bracket (Cor.
A.40). Therefore ŒX; Y ? D 0, so II is symmetric.
Because rz X Y jp depends only on Xp , it follows that the value of II.X; Y / at p
depends only on Xp , and in particular
is independent of the extension chosen for X .
Because rz X Y is linear over C 1 M
in X , and every f 2 C 1 .M / can be extended
to a smooth function on a neighborhood of M in M , it follows that II.X; Y / is linear
over C 1 .M / in X . By symmetry, the same claims hold for Y . t
u
As a consequence of the preceding proposition, for every p 2 M and all vectors
v; w 2 Tp M , it makes sense to interpret II.v; w/ as the value of II.V; W / at p,
where V and W are any vector fields on M such that Vp D v and Wp D W , and we
will do so from now on without further comment.
We have not yet identified the tangential term in the decomposition of r zXY .
Proposition 5.12(b) showed that in the special case of a submanifold of a Euclidean
or pseudo-Euclidean space, it is none other than the covariant derivative with respect
to the Levi-Civita connection of the induced metric on M . The following theorem
shows that the same is true in the general case. Therefore, we can interpret the
second fundamental form as a measure of the difference between the intrinsic Levi-
Civita connection on M and the ambient Levi-Civita connection on M .
z X Y D rX Y C II.X; Y /:
r
Proof. Because of the decomposition (8.1)and the definition of the second funda-
z X Y > D rX Y at all points of M .
mental form, it suffices to show that r
>
Define a map r W X.M / X.M / ! X.M / by
>
rX Y D rz X Y >;
The Gauss formula can also be used to compare intrinsic and extrinsic covariant
derivatives along curves. If W I ! M is a smooth curve and X is a vector field
along that is everywhere tangent to M , then we can regard X as either a vector
field along in M or a vector field along in M . We let Dz t X and Dt X denote
its covariant derivatives along as a curve in M and as a curve in M , respectively.
The next corollary shows how the two covariant derivatives are related.
Corollary 8.3 (The Gauss Formula Along a Curve). Suppose .M; g/ is an em-
bedded
Riemannian submanifold of a Riemannian or pseudo-Riemannian manifold
M ; gz , and W I ! M is a smooth curve. If X is a smooth vector field along that
is everywhere tangent to M , then
Dz t X D Dt X C II 0 ; X : (8.2)
The map WN is called the Weingarten map in the direction of N. Because the
second fundamental form is bilinear over C 1 .M /, it follows that WN is linear over
C 1 .M / and thus defines a smooth bundle homomorphism from TM to itself.
Proposition 8.4 (The Weingarten Equation). Suppose .M; g/ is an embedded
Rie-
mannian submanifold of a Riemannian or pseudo-Riemannian manifold M ; gz . For
every X 2 X.M / and N 2 .NM /, the following equation holds:
230 8 Riemannian Submanifolds
rz X N > D WN .X /; (8.5)
.
when N is extended arbitrarily to an open subset of M
z X N is independent of the
Proof. Note that at points of M , the covariant derivative r
choice of extensions of X and N by Proposition 4.26. Let Y 2 X.M / be arbitrary,
extended to a vector field on an open subset of M . Since hN; Y i vanishes identically
along M and X is tangent to M , the following holds at points of M :
0 D X hN; Y i
˝ ˛ ˝ ˛
D r zXY
z X N; Y C N; r
˝ ˛ ˝ ˛
D r z X N; Y C N; rX Y C II.X; Y /
˝ ˛ ˝ ˛
D r z X N; Y C N; II.X; Y /
˝ ˛ ˝ ˛
D r z X N; Y C WN .X /; Y :
f
Rm.W; X; Y; Z/ D Rm.W; X; Y; Z/ hII.W; Z/; II.X; Y /i C hII.W; Y /; II.X; Z/i:
Apply the Weingarten equation to each of the terms involving II (with II.X; Y / or
II.W; Y / playing the role of N ) to get
˝ ˛ ˝ ˛
f
Rm.W; X; Y; Z/ D r z W rX Y; Z II.X; Y /; II.W; Z/
˝ ˛ ˝ ˛ ˝ ˛
r z X rW Y; Z C II.W; Y /; II.X; Z/ rz ŒW;X Y; Z :
The Second Fundamental Form 231
Decomposing each term involving r z into its tangential and normal components, we
see that only the tangential component survives, because Z is tangent to M . The
Gauss formula allows each such term to be rewritten in terms of r, giving
f
Rm.W; X; Y; Z/ D hrW rX Y; Zi hrX rW Y; Zi hrŒW;X Y; Zi
hII.X; Y /; II.W; Z/i C hII.W; Y /; II.X; Z/i
D hR.W; X /Y; Zi
hII.X; Y /; II.W; Z/i C hII.W; Y /; II.X; Z/i: t
u
There is one other fundamental submanifold equation, which relates the normal
part of the ambient curvature endomorphism to derivatives of the second fundamen-
tal form. We will not have need for it, but we include it here for completeness. To
state it, we need to introduce a connection on the normal bundle of a Riemannian
submanifold.
If .M; g/
is aRiemannian submanifold of a?Riemannian or pseudo-Riemannian
manifold M ; gz , the normal connection r W X.M / .NM / ! .NM / is
defined by
rX?
ND r zX N ?;
.
where N is extended to a smooth vector field on a neighborhood of M in M
Proposition 8.6. If .M; g/ is an embedded
Riemannian
submanifold of a Riemann-
ian or pseudo-Riemannian manifold M ; gz , then r ? is a well-defined connection
in NM , which is compatible with gz in the sense that for any two sections N1 ; N2
of NM and every X 2 X.M /, we have
˝ ? ˛ ˝ ?
˛
X hN1 ; N2 i D rX N1 ; N2 C N1 ; rX N2 :
Now we are ready to state the last of the fundamental equations for submanifolds.
This equation was independently discovered (in the special case of surfaces in R3 )
by Karl M. Peterson (1853), Gaspare Mainardi (1856), and Delfino Codazzi (1868–
1869), and is sometimes designated by various combinations of these three names.
232 8 Riemannian Submanifolds
For the sake of simplicity we use the traditional but historically inaccurate name
Codazzi equation.
Proof. It suffices to show that both sides of (8.6) give the same result when we take
their inner products with an arbitrary smooth normal vector field N along M :
˝ ˛ ˝ F ˛ ˝ F ˛
z
R.W; X /Y; N D rW II .X; Y /; N rX II .W; Y /; N : (8.7)
We begin as in the proof of the Gauss equation: after extending the vector fields
to a neighborhood of M and applying the Gauss formula, we obtain
˝
f
Rm.W; X; Y; N / D r z W rX Y C II.X; Y / r z X rW Y C II.W; Y /
˛
r z ŒW;X Y; N :
Now when we expand the covariant derivatives, we need only pay attention to the
normal components. This yields
f .W; X; Y; N /
Rm
˝ F ˛
D II .W; rX Y / C rW II .X; Y / C II .rW X; Y / C II .X; rW Y / ; N
˝ F ˛
II .X; rW Y / C rX II .W; Y / C II .rX W; Y / C II .W; rX Y / ; N
˝ ˛
II .ŒW; X ; Y / ; N :
The terms involving rX Y and rW Y cancel each other in pairs, and three other
terms sum to zero because rW X rX W ŒW; X D 0. What remains is (8.7). u
t
Curvature of a Curve
manifold, the same approach works, but we have to restrict the definition to curves
such that j 0 .t /j is everywhere nonzero. Problem 8-6 gives a formula that can
be used in the Riemannian case to compute the geodesic curvature directly without
explicitly finding a unit-speed reparametrization.
From the definition, it follows that a smooth unit-speed curve has vanishing
geodesic curvature if and only if it is a geodesic, so the geodesic curvature of a
curve can be regarded as a quantitative measure of how far it deviates from being a
geodesic. If M D Rn with the Euclidean metric, the geodesic curvature agrees with
the notion of curvature
introduced
in advanced calculus courses.
Now suppose M ; gz is a Riemannian or pseudo-Riemannian manifold and
.M; g/ is a Riemannian submanifold. Every regular curve W I ! M has two dis-
tinct geodesic curvatures: its intrinsic curvature as a curve in M , and its extrinsic
curvature z as a curve in M . The second fundamental form can be used to compute
the relationship between the two.
Proof. Suppose W ."; "/ ! M is any regular curve with .0/ D p and 0 .0/ D v.
Applying the Gauss formula (Corollary 8.3) to the vector field 0 along , we obtain
z t 0 D Dt 0 C II. 0 ; 0 /:
D
Proof. Every vector v 2 V can be written v D y v for some unit vector vy, so the
bilinearity of B and B 0 implies B.v; v/ D B 0 .v; v/ for every v, not just unit vectors.
The result then follows from the following polarization identity, which is proved in
exactly the same way as its counterpart (2.2) for inner products:
B.v; w/ D 14 B.v C w; v C w/ B.v w; v w/ : t
u
234 8 Riemannian Submanifolds
Because the intrinsic and extrinsic accelerations of a curve are usually different,
it is generally not the case that a gz-geodesic that starts tangent to M stays in M ;
just think of a sphere
in Euclidean space, for example. A Riemannian submanifold
.M; g/ of M ; gz is said to be totally geodesic if every gz-geodesic that is tangent to
M at some time t0 stays in M for all t in some interval .t0 "; t0 C "/.
Proposition 8.12. Suppose .M; g/ is an embedded
Riemannian submanifold of a
Riemannian or pseudo-Riemannian manifold M ; gz , The following are equivalent:
.
(a) M is totally geodesic in M
(b) Every g-geodesic in M is also a gz-geodesic in M .
(c) The second fundamental form of M vanishes identically.
Proof. We will prove (a) ) (b) ) (c) ) (a). First assume that M is totally
geodesic. Let W I ! M be a g-geodesic. For each t0 2 I , let z W Iz ! M be the gz-
0 0
geodesic with z.t0 / D .t0 / and z .t0 / D .t0 /. The hypothesis implies that there
is some open interval I0 containing t0 such that z.t / 2 M for t 2 I0 . On I0 , the
Gauss formula (8.2) for z0 reads
0DD z t z0 D Dt z0 C II z0 ; z0 :
Because the first term on the right is tangent to M and the second is normal, the two
terms must vanish individually. In particular, Dt z0 0 on I0 , which means that z
is also a g-geodesic there. By uniqueness of geodesics, therefore, D z on I0 , so it
follows in turn that is a gz-geodesic there. Since the same is true in a neighborhood
of every t0 2 I , it follows that is a gz-geodesic on its whole domain.
Next assume that every g-geodesic is a gz-geodesic. Let p 2 M and v 2 Tp M be
arbitrary, and let D v W I ! M be the g-geodesic with .0/ D p and 0 .0/ D v.
The hypothesis implies that is also a gz-geodesic. Thus D z t 0 D Dt 0 D 0, so the
0 0
Gauss formula yields II. ; / D 0 along . In particular, II.v; v/ D 0. By Lemma
8.11, this implies that II is identically zero.
Finally, assume that II 0, and let z W I ! M be a gz-geodesic such that z.t0 / 2
0
M and z .t0 / 2 TM for some t0 2 I . Let W I ! M be the g-geodesic with the same
initial conditions: .t0 / D z.t0 / and 0 .t0 / D z0 .t0 /. The Gauss formula together
with the hypothesis II 0 implies that D z t D Dt D 0, so is also a gz-geodesic.
By uniqueness of geodesics, therefore, z D on the intersection of their domains,
which implies that z.t/ lies in M for t in some open interval around t0 . t
u
Hypersurfaces
Now we specialize the preceding considerations to the case in which M is a hyper-
surface (i.e., a submanifold of codimension 1) in M . Throughout this section, our
default assumption is that .M; g/ is an embedded n-dimensional
Riemannian sub-
manifold of an .n C 1/-dimensional Riemannian manifold M ; gz . (The analogous
formulas in the pseudo-Riemannian case are a little different; see Problem 8-19.)
Hypersurfaces 235
In this situation, at each point of M there are exactly two unit normal vectors.
In terms of any local adapted orthonormal frame .E1 ; : : : ; EnC1 /, the two choices
are ˙EnC1 . In a small enough neighborhood of each point of M , therefore, we can
always choose a smooth unit normal vector field along M .
If both M and M are orientable, we can use an orientation to pick out a global
smooth unit normal vector field along all of M . In general, though, this might or
might not be possible. Since all of our computations in this chapter are local, we
will always assume that we are working in a small enough neighborhood that a
smooth unit normal field exists. We will address as we go along the question of how
various quantities depend on the choice of normal vector field.
Having chosen a distinguished smooth unit normal vector field N on the hypersur-
face M M , we can replace the vector-valued second fundamental form II by a
simpler scalar-valued form. The scalar second fundamental
form of M is the sym-
metric covariant 2-tensor field h 2 †2 T M defined by h D IIN (see (8.3)); in
other words,
h.X; Y / D hN; II.X; Y /i: (8.8)
Using the Gauss formula r z X Y D rX Y C II.X; Y / and noting that rX Y is orthog-
onal to N , we can rewrite the definition as
˝ ˛
zXY :
h.X; Y / D N; r (8.9)
e
Rm.W; X; Y; Z/ D Rm.W; X; Y; Z/ 12 .h h/.W; X; Y; Z/:
e
Rm.W; X; Y; N / D .Dh/.Y; W; X /: (8.12)
Proof. Parts (a), (b), and (d) follow immediately from substituting (8.10) into the
general versions of the Gauss formula and Gauss equation. To prove (c), note first
that the general version of the Weingarten equation can be written rz X N > D sX .
˝ ˛
Since r z X N; N D 1 X jN j2 D 0, it follows that r z X N is tangent to M , so (c)
2
follows.
To prove the hypersurface Codazzi equation, note that the fact that N is a unit
vector field implies ˝ ? ˛
0 D X jN j2gz D 2 rX N; N gz :
Since N spans the normal bundle, this implies that N is parallel with respect to the
normal connection. Moreover,
Hypersurfaces 237
F ?
rW II .X; Y / D rW .II.X; Y // II.rW X; Y / II.X; rW Y /
?
D rW .h.X; Y /N / II.rW X; Y / II.X; rW Y /
D W .h.X; Y // h.rW X; Y / h.X; rW Y / N
D rW .h/.X; Y /N:
Inserting this into the general form (8.6) of the Codazzi equation and using the fact
that rh is symmetric in its first two indices yields
f
Rm.W; X; Y; N / D rW .h/.X; Y / rX .h/.W; Y /
D .rh/.X; Y; W / .rh/.W; Y; X /
D .rh/.Y; X; W / .rh/.Y; W; X /;
Principal Curvatures
At every point p 2 M , we have seen that the shape operator s is a self-adjoint linear
endomorphism of the tangent space Tp M . To analyze such an operator, we recall
some linear-algebraic facts about self-adjoint endomorphisms.
this basis, both h and s are represented by diagonal matrices, and h has the expres-
sion
h.v; w/ D 1 v 1 w 1 C C n v n w n :
The eigenvalues of s at a point p 2 M are called the principal curvatures
of M at p, and the corresponding eigenspaces are called the principal directions.
The principal curvatures all change sign if we reverse the normal vector, but the
principal directions and principal curvatures are otherwise independent of the choice
of coordinates or bases.
There are two combinations of the principal curvatures that play particularly
important roles for hypersurfaces. The Gaussian curvature is defined as K D det.s/,
and the mean curvature as H D .1=n/ tr.s/ D .1=n/ trg .h/. Since the determinant
and trace of a linear endomorphism are basis-independent, these are well defined
once a unit normal is chosen. In terms of the principal curvatures, they are
1
K D 1 2 n ; HD .1 C C n /;
n
as can be seen by expressing s in terms of an orthonormal basis of eigenvectors. If
N is replaced by N , then H changes sign, while K is multiplied by .1/n .
The formulas for sa and Ha follow by using .g ˛ / (the inverse matrix of .g˛ /) to
raise an index and then taking the trace. t
u
Minimal Hypersurfaces
A natural question that has received a great deal of attention over the past century is
this: Given a simple closed curve C in R3 , is there an embedded or immersed sur-
face M with @M D C that has least area among all surfaces with the same bound-
ary? If so, what is it? Such surfaces are models of the soap films that are produced
when a closed loop of wire is dipped in soapy water.
More generally, we can consider the analogous question for hypersurfaces in
Riemannian manifolds. Suppose M is a compact codimension-1 submanifold with
nonempty boundary in an .n C 1/-dimensional Riemannian manifold M ; gz . By
analogy with the case of surfaces in R3 , it is traditional to use the term area to
refer to the n-dimensional volume of M with its induced Riemannian metric, and
to say that M is area-minimizing if it has the smallest area among all compact
embedded hypersurfaces in M with the same boundary. One key observation is the
following theorem, which is an analogue for hypersurfaces of Theorem 6.4 about
length-minimizing curves.
Proof. Let g denote the induced metric on M . The fact that M minimizes area
among hypersurfaces with the same boundary means, in particular, that it minimizes
area among small perturbations of M in a neighborhood of a single point. We will
exploit this idea to prove that M must
have zero mean curvature everywhere.
Let p 2 Int M be arbitrary, let x 1 ; : : : ; x n ; v be Fermi coordinates for M on an
open set Uz M containing p (see Example 6.43), and let U D Uz \ M . By taking
240 8 Riemannian Submanifolds
For each t , let gyt D Mt gz denote the induced Riemannian metric on Mt , and
let gt D Ft gyt D Ft gz denote the pulled-back metric on M . When t D 0, we have
M0 D M , and both g0 and gy0 are equal to the induced metric g on M . Since gz
is given by (8.13) in Fermi coordinates, a simple computation shows that in U ,
gt D Ft gz has the coordinate expression gt D .gt /˛ˇ dx ˛ dx ˇ , where
@' @'
.gt /˛ˇ D t 2 ˛
.x/ ˇ .x/ C g˛ˇ x; t '.x/ ; (8.15)
@x @x
while on M X U , gt is equal to g and thus is independent of t .
Since each Mt is a smooth hypersurface with the same boundary as M , our
hypothesis guarantees that Area .Mt ; gyt / achieves a minimum at t D 0. Because Ft
is an isometry from .M; gt / to .Mt ; gyt /, we can express this area as follows:
The first term on the right is independent of t , and we can compute the second term
explicitly in coordinates x 1 ; : : : ; x n on U :
Z p
Area .U; gt / D det gt dx 1 dx n ;
U
where det gt denotes the determinant of the matrix 1 .gt /˛ˇ defined by (8.15). The
integrand above is a smooth function of t and x ; : : : ; x n , so the area is a smooth
function of t . We have
ˇ Z ˇ
d ˇˇ @ ˇˇ p
ˇ Area .Mt ; gyt / D ˇ det gt dx 1 dx n
dt tD0 U @t tD0
Z ˇ (8.16)
1 1=2 @ ˇ
ˇ 1 n
D .det g/ .det gt / dx dx :
U 2 @t ˇtD0
(The differentiation under the integral sign is justified because the integrand is
smooth and has compact support in U .)
To compute the derivative of the determinant, note that the expansion by minors
along, say, row ˛ shows that the partial derivative of det with respect to the matrix
entry in position .˛; ˇ/ is equal to the cofactor cof ˛ˇ , and thus by the chain rule,
ˇ ˇ
@ ˇˇ ˛ˇ @ ˇ
ˇ
ˇ .det gt / D cof .gt /˛ˇ : (8.17)
@t tD0 @t ˇtD0
On the other hand, Cramer’s rule shows that the .˛; ˇ/ component of the inverse
matrix is given by g ˛ˇ D .det g/1 cof ˛ˇ . Thus from (8.17) and (8.15) we obtain
ˇ ˇ
@ ˇˇ ˛ˇ @ ˇ
ˇ @g˛ˇ
ˇ .det g / D .det g/g .gt /˛ˇ D .det g/g ˛ˇ ':
@t ˇtD0
t
@t tD0 @v
Inserting this into (8.16) and using the result of Proposition 8.17, we conclude that
ˇ Z
d ˇˇ 1 @g˛ˇ
ˇ Area .M t ; y
g t / D .det g/1=2 g ˛ˇ ' dx 1 dx n
dt tD0 U 2 @v
Z (8.18)
D n H ' d Vg ;
U
and let Ms;t D @Ds;t , so D0;0 D D and M0;0 D M (see Fig. 8.3). For sufficiently
small s and t , the set Ds;t is a regular domain and Ms;t is a compact smooth hy-
persurface, and Vol.Ds;t / and Area .Ms;t / are both smooth functions of .s; t /. For
convenience, write V .s; t / D Vol.Ds;t / and A.s; t / D Area .Ms;t /.
The same argument that led to (8.18) shows that
Z Z
@A @A
.0; 0/ D n H ' d Vg ; .0; 0/ D n H d Vg :
@s U @t W
Hypersurfaces 243
To compute the partial derivatives of the volume, we just note that if we hold t D 0
fixed and let s vary, the only change in volume occurs in the part of Ds;t contained
in Uz , so the fundamental theorem of calculus gives
ˇ
@V d ˇˇ
.0; 0/ D Vol Ds;0 \ Uz
@s ds ˇsD0
ˇ Z Z s'.x/ p !
d ˇˇ
D det gz.x; v/ dv dx 1 dx n
ds ˇsD0 U "
Z ˇ Z s'.x/ p !
d ˇˇ
D det gz.x; v/ dv dx 1 dx n
U ds ˇsD0 "
Z p Z
D '.x/ det gz.x; 0/ dx 1 dx n D ' d Vg D 1;
U U
where the differentiation under the integral sign in the third line is justified just
like (8.16), and in the last line we used the fact that g˛ˇ .x/ D gz˛ˇ .x; 0/ in these
coordinates. Similarly, @V =@t .0; 0/ D 1.
Because V .0; 0/ D Vol.D/ and @V =@t .0; 0/ ¤ 0, the implicit function theorem
guarantees that there is a smooth function W .ı; ı/ ! R for some ı > 0 such that
V .s; .s// Vol.D/. The chain rule then implies
ˇ
d ˇˇ @V @V
0D ˇ V .s; .s// D .0; 0/ C 0 .0/ .0; 0/ D 1 C 0 .0/:
ds sD0 @s @t
point.) It is a remarkable fact that the Gauss and Codazzi equations are actually
sufficient, at least locally. A sketch of a proof of this fact, called the fundamental
theorem of hypersurface theory, can be found in [Pet16, pp. 108–109].
In the setting of a hypersurface M RnC1 , we can give some very concrete
geometric interpretations of the quantities we have defined so far. We begin with
curves. For every unit vector v 2 Tp M , let D v W I ! M be the g-geodesic in M
with initial velocity v. Then the Gauss formula shows that the ordinary Euclidean
x t 0 .0/ D h.v; v/Np (Fig. 8.4). Thus jh.v; v/j is
acceleration of at 0 is 00 .0/ D D
the Euclidean curvature of at 0, and h.v; v/ D h 00 .0/; Np i > 0 if and only if 00 .0/
points in the same direction as Np . In other words, h.v; v/ is positive if is curving
in the direction of Np , and negative if it is curving away from Np .
The next proposition shows that this Euclidean curvature can be interpreted in
terms of the radius of the “best circular approximation,” as mentioned in Chapter 1.
Proof. An easy geometric argument shows that every circle in Rm with center q and
radius R has a unit-speed parametrization of the form
t t t t
0 0
c.t / D q C R cos v C R sin w;
R R
where .v; w/ is a pair of orthonormal vectors in Rm . By direct computation, such a
parametrization satisfies
1
c.t0 / D q C Rv; c 0 .t0 / D w; c 00 .t0 / D v:
R
Thus if we put
246 8 Riemannian Submanifolds
1 1
RD D ; v D R 00 .t0 /; w D 0 .t0 /; q D .t0 / Rv;
j 00 .t0 /j .t0 /
we obtain a circle satisfying the required conditions, and its radius is equal to
1=.t0 / by construction. Uniqueness is left as an exercise. u
t
I Exercise 8.22. Complete the proof of the preceding proposition by proving uniqueness
of the osculating circle.
@ X
h Xi ; Xj D ;N : (8.23)
@ui @uj
Hypersurfaces in Euclidean Space 247
Proof. Let u0 D u10 ; : : : ; un0 be an arbitrary point of U and let p D X.u0 / 2 M . For
each i 2 f1; : : : ; ng, the curve .t / D X u10 ; : : : ; ui0 C t; : : : ; un0 is a smooth curve in
M whose initial velocity is Xi . Regarding the normal field N as a smooth map
from M to RnC1 , we have
@ x X N.X.u0 //:
N.X.u0 // D .N ı /0 .0/ D r
@ui i
@X
.u/; N.X.u// :
@uj
Differentiating with respect to ui and using the product rule for ordinary inner prod-
ucts in RnC1 yields
2
@ X @X x
0D .u/; N.X.u// C .u/; rXi .u/ N.X.u// :
@ui @uj @uj
By the Weingarten equation (8.11), the last term on the right becomes
˝ ˛
Xj .u/; s.Xi .u// D h.Xj .u/; Xi .u//:
Proof. Replacing M by its image under a translation and a rotation (which are
Euclidean isometries), we may assume that p is the origin and Tp M is equal to the
span of .@1 ; : : : ; @n /. Then after reflecting in the x 1 ; : : : ; x n -hyperplane if neces-
sary, we may assume that the chosen unit normal is .0; : : : ; 0; 1/. By an orthogonal
transformation in the first n variables, we can also arrange that the scalar second fun-
damental form at 0 is diagonal with respect to the basis .@1 ; : : : ; @n /, with diagonal
entries .1 ; : : : ; n /.
It follows from the implicit function theorem that there is some neighborhood
U of 0 such that M \ U is the graph of a smooth function of the form x nC1 D
f x 1 ; : : : ; x n with f .0/ D 0. A smooth local parametrization of M is then given by
248 8 Riemannian Submanifolds
X.u/ D u1 ; : : : ; un ; f .u/ , and the fact that T0 M is spanned by @=@x 1 ; : : : ; @=@x n
guarantees that @1 f .0/ D D @n f .0/ D 0. Because Xi D @=@x i at 0, Proposition
8.23 then yields
@ @ @2 f @2 f
h i
; j
D 0; : : : ; 0; i j
.0/ ; .0; : : : ; 0; 1/ D i j .0/:
@x @x @x @x @x @x
One common way to produce such a smooth vector field is to work with a local
defining function for M : Recall that this is a smooth real-valued function defined
on some open subset U RnC1 such that U \ M is a regular level set of F (see
Prop. A.27). The definition ensures that grad F (the gradient of F with respect to
gx) is nonzero on some neighborhood of M \ U , so a convenient choice for a unit
normal vector field along M is
grad F
ND :
j grad F j
Here is an application.
Therefore s D .1=R/Id. The principal curvatures, therefore, are all equal to 1=R,
and it follows that the mean curvature is H D 1=R and the Gaussian curvature is
.1=R/n . //
For surfaces in R3 , either of the above methods can be used. When a parametriza-
tion X is given, the normal vector field is particularly easy to compute: because X1
and X2 span the tangent space to M at each point, their cross product is a nonzero
normal vector, so one choice of unit normal is
X1 X2
ND : (8.26)
jX1 X2 j
Problems 8-1, 8-2, and 8-3 will give you practice in carrying out these computations
for surfaces presented in various ways.
Because the Gaussian and mean curvatures are defined in terms of a particular
embedding of M into RnC1 , there is little reason to suspect that they have much to
do with the intrinsic Riemannian geometry of .M; g/. The next exercise illustrates
the fact that the mean curvature has no intrinsic meaning.
I Exercise 8.26. Let M1 R3 be the plane fz D 0g, and let M2 R3 be the cylinder
fx 2 C y 2 D 1g. Show that M1 and M2 are locally isometric, but the former has mean
curvature zero, while the latter has mean curvature ˙ 12 , depending on which normal is
chosen.
The amazing discovery made by Gauss was that the Gaussian curvature of a
surface in R3 is actually an intrinsic invariant of the Riemannian manifold .M; g/.
He was so impressed with this discovery that he called it Theorema Egregium, Latin
for “excellent theorem.”
On the other hand, Corollary 7.27 shows that Rm D 14 Sg g, and thus by the defi-
nition of the Kulkarni–Nomizu product we have
250 8 Riemannian Submanifolds
Rmp .b1 ; b2 ; b2 ; b1 / D 14 S.p/ 2g11 g22 2g12 g21 D 12 S.p/: t
u
Sectional Curvatures
p
jv ^ wj D jvj2 jwj2 hv; wi2 : (8.27)
It follows from the Cauchy–Schwarz inequality that jv ^ wj 0, with equality if
and only if v and w are linearly dependent, and jv ^ wj D 1 when v and w are
orthonormal. (One can define an inner product on the space ƒ2 .V / of contravariant
alternating 2-tensors, analogous to the inner product on forms defined in Problem
2-16, and this is the associated norm; see Problem 8-33(a).)
Proposition 8.29 (Formula for the Sectional Curvature). Let .M; g/ be a Rie-
mannian manifold and p 2 M . If v; w are linearly independent vectors in Tp M ,
then the sectional curvature of the plane spanned by v and w is given by
Rmp .v; w; w; v/
sec.v; w/ D : (8.28)
jv ^ wj2
Proof. Let ˘ Tp M be the subspace spanned by .v; w/. For this proof, we de-
note the induced metric on S˘ by gy, and its associated curvature tensor by Rm. c By
y
definition, sec.v; w/ is equal to K.p/, the Gaussian curvature of gy at p.
We show first that the second fundamental form of S˘ in M vanishes at p. To
see why, let z 2 ˘ be arbitrary, and let D z be the g-geodesic with initial velocity
z, whose image lies in S˘ for t sufficiently near 0. By the Gauss formula for vector
fields along curves,
0 D Dt 0 D D y t 0 C II. 0 ; 0 /:
Since the two terms on the right-hand side are orthogonal, each must vanish iden-
tically. Evaluating at t D 0 gives II.z; z/ D 0. Since z was an arbitrary element of
˘ D Tp .S˘ / and II is symmetric, polarization shows that II D 0 at p. (We cannot in
general expect II to vanish at other points of S˘ —it is only at p that all g-geodesics
starting tangent to S˘ remain in S˘ .) The Gauss equation then tells us that the
curvature tensors of M and S˘ are related at p by
c p .u; v; w; x/
Rmp .u; v; w; x/ D Rm
whenever u; v; w; x 2 ˘ .
Now choose an orthonormal basis .b1 ; b2 / for ˘ . Based on the observations
above, we see that the sectional curvature of ˘ is
y
K.p/ y
D 12 S.p/
2
X
D 1 c p b i ; bj ; bj ; bi
Rm
2
i;j D1
1c c p .b2 ; b1 ; b1 ; b2 /
D 2
Rmp .b1 ; b2 ; b2 ; b1 / C 12 Rm
c p .b1 ; b2 ; b2 ; b1 / D Rmp .b1 ; b2 ; b2 ; b1 / :
D Rm
To see how to compute this in terms of an arbitrary basis, let .v; w/ be any basis for
˘ . The Gram–Schmidt algorithm yields an orthonormal basis as follows:
252 8 Riemannian Submanifolds
v
b1 D I
jvj
w hw; b1 ib1
b2 D :
jw hw; b1 ib1 j
Then by the preceding computation,
y
K.p/ D Rmp .b1 ; b2 ; b2 ; b1 /
v w hw; b1 ib1 w hw; b1 ib1 v
D Rmp ; ; ;
jvj jw hw; b1 ib1 j jw hw; b1 ib1 j jvj (8.29)
Rmp .v; w; w; v/
D 2 ;
jvj jw hw; b1 ib1 j2
I Exercise 8.30. Suppose .M; g/ is a Riemannian manifold and gz D g for some pos-
itive constant . Use Theorem 7.30 to prove that for every p 2 M and plane ˘ Tp M ,
z and g are related by sf
the sectional curvatures of ˘ with respect to g ec.˘ / D 1 sec.˘ /.
Proposition 8.29 shows that one important piece of quantitative information pro-
vided by the curvature tensor is that it encodes the sectional curvatures of all plane
sections. It turns out, in fact, that this is all of the information contained in the curva-
ture tensor: as the following proposition shows, the sectional curvatures completely
determine the curvature tensor.
0 D D.v C w; x; x; v C w/
D D.v; x; x; v/ C D.v; x; x; w/ C D.w; x; x; v/ C D.w; x; x; w/
D 2D.v; x; x; w/:
We can also give a geometric interpretation of the Ricci and scalar curvatures on
a Riemannian manifold. Since the Ricci tensor is symmetric and bilinear, Lemma
8.11 shows that it is completely determined by its values of the form Rc.v; v/ for
unit vectors v.
Proof. Given any unit vector v 2 Tp M , let .b1 ; : : : ; bn / be as in the hypothesis. Then
Rcp .v; v/ is given by
n
X n
X
Rcp .v; v/ D R11 .p/ D Rk11 k .p/ D Rmp .bk ; b1 ; b1 ; bk / D sec.b1 ; bk /:
kD1 kD2
For the scalar curvature, we let .b1 ; : : : ; bn / be any orthonormal basis for Tp M , and
compute
n
X n
X
S.p/ D Rj j .p/ D Rcp bj ; bj D Rmp bk ; bj ; bj ; bk
j D1 j;kD1
X
D sec bj ; bk : t
u
j ¤k
254 8 Riemannian Submanifolds
We can now compute the sectional curvatures of our three families of frame-
homogeneous model spaces. A Riemannian metric or Riemannian manifold is said
to have constant sectional curvature if the sectional curvatures are the same for all
planes at all points.
Proof. Frame homogeneity implies, in particular, that given two 2-planes at the
same or different points, there is an isometry taking one to the other. The result
follows from the isometry invariance of the curvature tensor. t
u
Thus to compute the sectional curvature of one of our model spaces, it suffices
to compute the sectional curvature for one plane at one point in each space.
Theorem 8.34 (Sectional Curvatures of the Model Spaces). The following Rie-
mannian manifolds have the indicated constant sectional curvatures:
(a) Rn ; gx has constant sectional curvature 0.
n
(b) S .R/; gV R has constant sectional curvature 1=R2 .
(c) Hn .R/; gM R has constant sectional curvature 1=R2 .
Proof. First we consider the simplest case: Euclidean space. Since the curvature
tensor of Rn is identically zero, clearly all sectional curvatures are zero. This is also
easy to see geometrically, since each plane section is actually a plane, which has
zero Gaussian curvature.
Next consider the sphere Sn .R/. We need only compute the sectional curvature
of the plane ˘ spanned by .@1 ; @2 / at the point .0; : : : ; 0; 1/. The geodesics with
initial velocities in ˘ are great circles in the .x 1 ; x 2 ; x nC1 / subspace. Therefore
S˘ is isometric to the round 2-sphere of radius R embedded in R3 . As Example
8.25 showed, S2 .R/ has Gaussian curvature 1=R2 . Therefore Sn .R/ has constant
sectional curvature equal to 1=R2 .
Finally, the proof for hyperbolic spaces is left to Problem 8-28. t
u
Problems 255
Note that for every real number c, exactly one of the model spaces listed above
has constant sectional curvature c. These spaces will play vital roles in our compar-
ison and local-to-global theorems in the last two chapters of the book.
I Exercise 8.35. Show that the metric on real projective space RP n defined in Example
2.34 has constant positive sectional curvature.
Since the sectional curvatures determine the curvature tensor, one would expect
to have an explicit formula for the full curvature tensor when the sectional curvature
is constant. Such a formula is provided in the following proposition.
Proposition 8.36. A Riemannian metric g has constant sectional curvature c if and
only if its curvature tensor satisfies
Rm D 12 cg g:
In this case, the Ricci tensor and scalar curvature of g are given by the formulas
Problems
8-1. Suppose U is an open set in Rn and f W U ! R is a smooth function. Let
M D f.x; f .x// W x 2 U g RnC1 be the graph of f , endowed with the
induced Riemannian metric and upward unit normal (see Example 2.19).
(a) Compute the components of the shape operator in graph coordinates, in
terms of f and its partial derivatives.
(b) Let M RnC1 be the n-dimensional paraboloid defined as the graph
of f .x/ D jxj2 . Compute the principal curvatures of M .
(Used on p. 361.)
8-2. Let .M; g/ be an embedded Riemannian hypersurface in a Riemannian
manifold M ; gz , let F be a local defining function for M , and let N D
grad F=j grad F j.
256 8 Riemannian Submanifolds
(a) Show that the scalar second fundamental form of M with respect to the
unit normal N is given by
z 2 F .X; Y /
r
h.X; Y / D
j grad F j
these curves as ˘ varies. [Remark: This justifies the informal recipe for
computing principal
curvatures given in Chapter 1.]
8-12. Suppose W M ; gz ! .M; g/ is a Riemannian submersion, and M ; gz has
all sectional curvatures bounded below by a constant c. Use O’Neill’s for-
mula (Problem 7-14) to show that the sectional curvatures of .M; g/ are
bounded below by the same constant.
8-13. Let p W S2nC1 ! CP n be the Riemannian submersion described in Exam-
ple 2.30. In this problem, we identify C nC1 with R2nC2 by means of coor-
dinates .x 1 ; y 1 ; : : : ; x nC1 ; y nC1 / defined by z j D x j C iy j .
(a) Show that the vector field
@ @
S D xj j
yj j
@y @x
(e) Show that for n 2, the sectional curvatures at each point of CP n take
on all values between 1 and 4, inclusive, and conclude that CP n is not
frame-homogeneous.
(f) Compute the Gaussian curvature of CP 1 .
(Used on pp. 56, 262, 367.)
8-14. Show that a Riemannian 3-manifold is Einstein if and only if it has constant
sectional curvature.
8-15. Suppose .M; g/ is a 3-dimensional Riemannian manifold that is homoge-
neous and isotropic. Show that g has constant sectional curvature. Show that
the analogous result in dimension 4 is not true. [Hint: See Problem 8-13.]
8-16. For each a > 0, let ga be the Berger metric on SU.2/ defined in Problem 3-10.
Compute the sectional curvatures with respect to ga of the planes spanned by
.X; Y /, .Y; Z/, and .Z; X /. Prove that if a ¤ 1, then .SU.2/; ga / is homoge-
neous but not isotropic anywhere. (Used on p. 56.)
8-17. Let G be a Lie group with a bi-invariant metric g (see Problem 7-13).
(a) Suppose X and Y are orthonormal elements of Lie.G/. Show that
ˇ ˇ2
sec.Xp ; Yp / D 14 ˇŒX; Y ˇ for each p 2 G, and conclude that the sec-
tional curvatures of .G; g/ are all nonnegative.
(b) Show that every Lie subgroup of G is totally geodesic in G.
(c) Now suppose G is connected. Show that G is flat if and only if it is
abelian.
(Used on p. 282.)
8-18. Suppose
.M; g/ is a Riemannian hypersurface in a Riemannian manifold
M ; gz , and N is a unit normal vector field along M . We say that M is
convex (with respect to N ) if its scalar second fundamental form satisfies
h.v; v/ 0 for all v 2 TM . Show that if M is convex and M has sectional
curvatures bounded below by a constant c, then all sectional curvatures of M
are bounded below by c.
8-19. Suppose .M; g/ is a Riemannian
hypersurface in an .n C 1/-dimensional
Lorentz manifold M ; gz , and N is a smooth unit normal vector field along
M . Define the scalar second fundamental form h and the shape operator s
by requiring that II.X; Y / D h.X; Y /N and hsX; Y i D h.X; Y / for all X; Y 2
X.M /. Prove the following Lorentz analogues of the formulas of Theorem
8.13 (with notation as in that theorem):
(a) G AUSS F ORMULA : r z X Y D rX Y C h.X; Y /N .
(b) G AUSS F ORMULA FOR A C URVE : D z t X D Dt X C h. 0 ; X /N .
(c) W EINGARTEN E QUATION : r z X N D sX .
(d) f
G AUSS E QUATION : Rm.W; X; Y; Z/ D Rm C 12 h h .W; X; Y; Z/.
(e) f
C ODAZZI E QUATION : Rm.W; X; Y; N / D .Dh/.Y; W; X /.
8-20. Suppose M ; gz is an .n C 1/-dimensional Lorentz manifold, and assume
that gz satisfies the Einstein field equation with a cosmological constant:
260 8 Riemannian Submanifolds
e 1 Szgz C z
Rc g D T , where is a constant and T is a smooth symmetric
2
,
2-tensor field (see p. 211). Let .M; g/ be a Riemannian hypersurface in M
and let h be its scalar second fundamental form as defined in Problem 8-19.
Use the results of Problem 8-19 to show that g and h satisfy the following
Einstein constraint equations on M :
2
S 2 jhj2g C trg h D 2;
div h d trg h D J;
.x 1 ; : : : ; x i ; : : : ; x j ; : : : ; x n / 7! .x 1 ; : : : ; x j ; : : : ; x i ; : : : ; x n /;
R 2 P i 2
and compute Sn1 x i d VgV using the fact that i x D 1 on the
sphere.] [Remark: It is a standard fact of linear algebra that the trace of A
is independent of the choice of basis. This gives a geometric interpretation to
the trace as n times the average value of the associated quadratic form Q.]
(Used on p. 313.)
8-22. Let .M; g/ be a Riemannian n-manifold and p 2 M . Proposition 8.32 gave
a geometric interpretation of the Ricci curvature at p based on a choice of
orthonormal basis. This problem describes an interpretation that does not
refer to a basis. For each unit vector v 2 Tp M , prove that
Z
n1
Rcp .v; v/ D sec.v; w/ d Vgy ;
Vol Sn2 w2Sv?
where Sv? denotes the set of unit vectors in Tp M that are orthogonal to v and
gy denotes the Riemannian metric on Sv? induced from the flat metric gp on
Tp M . [Hint: Use Problem 8-21.]
Problems 261
1 1
W1221 D .k12 C k34 / .k13 C k14 C k23 C k24 /;
3 6
where kij is the sectional curvature of the plane spanned by bi ; bj .
8-27. Prove that the Fubini–Study metric on CP 2 is not locally conformally flat.
[Hint: Use Problems 8-13 and 8-26.]
8-28. Complete the proof of Theorem 8.34 by showing in two ways that the hy-
perbolic space of radius R has constant sectional curvature equal to 1=R2 .
(a) In the hyperboloid model, compute the second fundamental form of
Hn .R/ Rn;1 at the point .0; : : : ; 0; R/, and use either the general form
of the Gauss equation (Thm. 8.5) or the formulas of Problem 8-19.
(b) In the Poincaré ball model, use formula (7.44) for the conformal trans-
formation of the curvature to compute the Riemann curvature tensor at
the origin.
8-29. Prove Proposition 8.36 (curvature tensors on constant-curvature spaces).
8-30. Suppose M RnC1 is a hypersurface with the induced Riemannian metric.
Show that the Ricci tensor of M satisfies
262 8 Riemannian Submanifolds
˝ ˛
Rc.v; w/ D nH sv s 2 v; w ;
where H and s are the mean curvature and shape operator of M , respec-
tively, and s 2 v D s.sv/.
8-31. For 1 < k n, show that any k points in the hyperbolic space Hn .R/ lie
in a totally geodesic .k 1/-dimensional submanifold, which is isometric to
Hk1 .R/.
8-32. Suppose .M; g/ is a Riemannian manifold and G is a Lie group acting
smoothly and isometrically on M . Let S M be the fixed point set of G,
that is, the set of points p 2 M such that '.p/ D p for all ' 2 G. Show
that each connected component of S is a smoothly embedded totally geodesic
submanifold of M . (The reason for restricting to a single connected comp-
onent is that different components may have different dimensions.)
8-33. Suppose .M; g/ is a Riemannian manifold. Let ƒ2 .TM / denote the bundle
of 2-vectors (alternating contravariant 2-tensors) on M .
(a) Show that there is a unique fiber metric on ƒ2 .TM / whose associated
norm satisfies
jw ^ xj2 D jwj2 jxj2 hw; xi2
for all tangent vectors w; x at every point q 2 M .
(b) Show that there is a unique bundle endomorphism R W ƒ2 .TM / !
ƒ2 .TM /, called the curvature operator of g, that satisfies
˝ ˛
R.w ^ x/; y ^ z D Rm.w; x; y; z/ (8.30)
All the work we have done so far has been focused on purely local properties of
Riemannian manifolds and their submanifolds. We are finally in a position to prove
our first major local-to-global theorem in Riemannian geometry: the Gauss–Bonnet
theorem. The grandfather of all such theorems in Riemannian geometry, it is a local-
to-global theorem par excellence, because it asserts the equality of two very differ-
ently defined quantities on a compact Riemannian 2-manifold: the integral of the
Gaussian curvature, which is determined by the local geometry, and 2 times the
Euler characteristic, which is a global topological invariant. Although in this form
it applies only in two dimensions, it has provided a model and an inspiration for
innumerable local-to-global results in higher-dimensional geometry, some of which
we will prove in Chapter 12.
This chapter begins with some not-so-elementary notions from plane geome-
try, leading up to a proof of Hopf’s rotation index theorem, which expresses the
intuitive idea that the velocity vector of a simple closed plane curve, or more gen-
erally of a “curved polygon,” makes a net rotation through an angle of exactly 2
as one traverses the curve counterclockwise. Then we investigate curved polygons
on Riemannian 2-manifolds, leading to a far-reaching generalization of the rota-
tion index theorem called the Gauss–Bonnet formula, which relates the exterior an-
gles, geodesic curvature of the boundary, and Gaussian curvature in the interior of
a curved polygon. Finally, we use the Gauss–Bonnet formula to prove the global
statement of the Gauss–Bonnet theorem.
curve, the velocity vector makes a net rotation through an angle of exactly 2. Our
task in the first part of this chapter is to make these notions precise.
Throughout this section, W Œa; b ! R2 is an admissible curve in the plane. We
say that is a simple closed curve if .a/ D .b/ but is injective on Œa; b/. We do not
assume that has unit speed; instead, we define the unit tangent vector field of as
the vector field T along each smooth segment of given by
0 .t /
T .t / D :
j 0 .t /j
Since each tangent space to R2 is naturally identified with R2 itself, we can think
of T as a map into R2 , and since T has unit length, it takes its values in S1 .
If is smooth (or at least continuously differentiable), we define a tangent an-
gle function for to be a continuous function W Œa; b ! R such that T .t / D
.cos .t /; sin .t // for all t 2 Œa; b. It follows from the theory of covering spaces
that such a function exists: the map q W R ! S1 given by q.s/ D .cos s; sin s/ is
a smooth covering map, and the path-lifting property of covering maps (Prop.
A.54(b)) ensures that there is a continuous function W Œa; b ! R that satisfies
q..t // D T .t /. By the unique lifting property (Prop. A.54(a)), a lift is uniquely
determined once its value at any single point is determined, and thus any two lifts
differ by a constant integral multiple of 2.
If is a continuously differentiable simple closed curve such that 0 .a/ D 0 .b/
(Fig. 9.1), then .cos .a/; sin .a// D .cos .b/; sin .b//, so .b/ .a/ must be an
integral multiple of 2. For such a curve, we define the rotation index of to be
the following integer:
1
. / D .b/ .a/ ;
2
where is any tangent angle function for . For any other choice of tangent angle
function, .a/ and .b/ would change by addition of the same constant, so the
rotation index is independent of the choice of .
We would also like to extend the definition of the rotation index to certain piece-
wise regular closed curves. For this purpose, we have to take into account the
“jumps” in the tangent angle function at corners. To do so, suppose W Œa; b ! R2
Some Plane Geometry 265
Fig. 9.5: Tangent angle at a vertex Fig. 9.6: Tangent angle function
where "k is the exterior angle at .b/. (See Figs. 9.5 and 9.6.) Such a function always
exists: start by defining .t / for t 2 Œa; a1 / to be any lift of T on that interval; then on
Œa1 ; a2 / define .t / to be the unique lift that satisfies (9.1), and continue by induc-
tion, ending with .b/ defined by (9.2). Once again, the difference between any two
such functions is aconstant integral
multiple of 2. We define the rotation index of
1
to be . / D 2 .b/ .a/ just as in the smooth case. As before, . / is an in-
teger, because the definition ensures that .cos .b/; sin .b// D .cos .a/; sin .a//.
The following theorem was first proved by Heinz Hopf in 1935. (For a readable
version of Hopf’s proof, see [Hop89, p. 42].) It is frequently referred to by the
German name given to it by Hopf, the Umlaufsatz.
Theorem 9.1 (Rotation Index Theorem). The rotation index of a positively ori-
ented curved polygon in the plane is C1.
Some Plane Geometry 267
Fig. 9.7: The curve after changing the parameter interval and translating .a/ to the origin
Proof. Let W Œa; b ! R2 be such a curved polygon. Assume first that all the ver-
tices of are flat. This means, in particular, that 0 is continuous and 0 .a/ D 0 .b/.
Since .a/ D .b/, we can extend to a continuous map from R to R2 by requiring
it to be periodic of period b a, and our hypothesis 0 .a/ D 0 .b/ guarantees that
the extended map still has continuous first derivatives. Define T .t / D 0 .t /=j 0 .t /j
as before.
Let W R ! R be any lift of T W R ! S1 . Then jŒa;b is a tangent angle function
for , and thus .b/ D .a/ C 2. /. If we set z.t / D .t C b a/ 2. /, then
cos z.t /; sin z.t / D cos .t C b a/; sin .t C b a/ D T .t C b a/ D T .t /;
so jŒa1 ;b1 has the same rotation index as jŒa;b . Thus we obtain the same result by
restricting to any closed interval of length b a.
Using this freedom, we can assume that the parameter interval Œa; b has been
chosen so that the y-coordinate of achieves its minimum at t D a. Moreover, by a
translation in the xy-plane (which does not change 0 or ), we may as well assume
that .a/ is the origin. With these adjustments, the image of remains in the closed
upper half-plane, and T .a/ D T .b/ D .1; 0/ (Fig. 9.7). By adding a constant integral
multiple of 2 to if necessary, we can also assume that .a/ D 0.
Next, we define a continuous secant angle function, denoted by '.t1 ; t2 /, repre-
senting the angle between the positive x-direction and the vector from .t1 / to .t2 /.
To be precise, let R2 be the triangular region D f.t1 ; t2 / W a t1 t2 bg
268 9 The Gauss–Bonnet Theorem
Fig. 9.8: The domain of ' Fig. 9.9: The secant angle function
The function V is continuous at points where t1 < t2 and .t1 ; t2 / ¤ .a; b/, because
is continuous and injective there. To see that it is continuous elsewhere, note that
for t1 < t2 , the fundamental theorem of calculus gives
Z 1 Z 1
d
.t2 / .t1 / D t1 C s.t2 t1 / ds D 0 t1 C s.t2 t1 / .t2 t1 / ds;
0 ds 0
and thus
ˇ ˇ Z 1
ˇ .t2 / .t1 / ˇ ˇ 0 ˇ
ˇ .t /ˇˇ
0 ˇ t1 C s.t2 t1 / 0 .t /ˇ ds:
ˇ t t
2 1 0
Because 0 is uniformly continuous on the compact set Œa; b, this last expression
can be made as small as desired by taking .t1 ; t2 / close to .t; t /. It follows that
.t2 / .t1 /
lim D 0 .t /;
.t1 ;t2 /!.t;t/ t2 t1
t1 <t2
and therefore
Some Plane Geometry 269
.t2 / .t1 /
lim V .t1 ; t2 / D lim
.t1 ;t2 /!.t;t/ .t1 ;t2 /!.t;t/ j.t2 / .t1 /j
t1 <t2 t1 <t2
,ˇ ˇ
.t2 / .t1 / ˇ .t2 / .t1 / ˇ
D lim ˇ ˇ
.t1 ;t2 /!.t;t/ t2 t1 ˇ t t ˇ
2 1
t1 <t2
0
.t /
D D T .t / D V .t; t /:
j 0 .t /j
Similarly, because T is continuous,
.t2 / .t1 C b a/
lim V .t1 ; t2 / D lim
.t1 ;t2 /!.a;b/ .t1 ;t2 /!.a;b/ j.t2 / .t1 C b a/j
t1 <t2 t1 <t2
.s2 / .s1 /
D lim D T .b/ D V .a; b/:
.s1 ;s2 /!.b;b/ j.s2 / .s1 /j
s1 >s2
Thus V is continuous.
Since is simply connected, Corollary A.57 guarantees that V W ! S1 has a
continuous lift ' W ! R, which is unique if we require '.a; a/ D 0 (Fig. 9.9). This
is our secant angle function.
We can express the rotation index in terms of the secant angle function as follows:
1 1 1
. / D .b/ .a/ D '.b; b/ '.a; a/ D '.b; b/:
2 2 2
Observe that along the side of where t1 D a and t2 2 Œa; b, the vector V .a; t2 /
has its tail at the origin and its head in the upper half-plane. Since we stipulate that
'.a; a/ D 0, we must have '.a; t2 / 2 Œ0; on this segment. By continuity, there-
fore, '.a; b/ D (since '.a; b/ represents the tangent angle of T .b/ D .1; 0/).
Similarly, on the side where t2 D b, the vector V .t1 ; b/ has its head at the origin and
its tail in the upper half-plane, so '.t1 ; b/ 2 Œ; 2. Therefore, since '.b; b/ repre-
sents the tangent angle of T .b/ D .1; 0/, we must have '.b; b/ D 2 and therefore
. / D 1. This completes the proof for the case in which 0 is continuous.
Now suppose has one or more ordinary vertices. It suffices to show there is a
curve with a continuous velocity vector field that has the same rotation index as .
We will construct such a curve by “rounding the corners” of . It will simplify the
proof somewhat if we choose the parameter interval Œa; b so that .a/ D .b/ is not
one of the ordinary vertices.
270 9 The Gauss–Bonnet Theorem
Let .ai / be any ordinary vertex, let "i be its exterior angle, and let ˛ be a
positive number less than 12 . j"i j/. Recall that is continuous from the right
at ai and limt%ai .t / D .ai / "i . Therefore, we can choose ı small enough that
j.t / ..ai / "i /j < ˛ when t 2 .ai ı; ai /, and j.t / .ai /j < ˛ when t 2
.ai ; ai C ı/.
The image under of Œa; bX.ai ı; ai Cı/ is a compact set that does not contain
.ai /, so we can choose r small enough that does not enter Bxr ..ai // except when
t 2 .ai ı; ai C ı/. Let t1 2 .ai ı; ai / denote a time when enters Bxr ..ai //, and
t2 2 .ai ; ai C ı/ a time when it leaves (Fig. 9.10). By our choice of ı, the total
change in .t / is not more than ˛ when t 2 Œt1 ; ai /, and again not more than ˛ when
t 2 .ai ; t2 . Therefore, the total change in .t / during the time interval Œt1 ; t2 is
between "i 2˛ and "i C 2˛, which implies < < .
Now we simply replace jŒt1 ;t2 with a smooth curve segment that has the same
velocity as at .t1 / and .t2 /, and whose tangent angle increases or decreases
monotonically from .t1 / to .t2 /; an arc of a hyperbola will do (Fig. 9.11). Since
The Gauss–Bonnet Formula 271
the change in tangent angle of is between and and represents the angle
between T .t1 / and T .t2 /, it must be exactly . Repeating this process for each
vertex, we obtain a new curved polygon with a continuous velocity vector field
whose rotation index is the same as that of , thus proving the theorem. t
u
From the rotation index theorem, it is not hard to deduce the three local-to-global
theorems mentioned at the beginning of the chapter as corollaries. (The angle-sum
theorem is immediate; for the total curvature theorem, the trick is to show that 0 .t /
is equal to the signed curvature of ; the circumference theorem follows from the to-
tal curvature theorem as mentioned in Chapter 1.) However, instead of proving them
directly, we will prove a general formula, called the Gauss–Bonnet formula, from
which these results and more follow easily. You will easily see how the statement
and proof of Theorem 9.3 below can be simplified in case the metric is Euclidean.
at each t where 0 is continuous, and that is continuous from the right and satisfies
(9.1) and (9.2) at vertices. The existence of such a function follows as in the planar
case, using the fact that
Proof. If we use the given oriented coordinate chart to regard as a curved polygon
in the plane, we can compute its tangent angle function either with respect to g or
with respect to the Euclidean metric gx. In either case, . / is an integer because
.a/ and .b/ both represent the angle between the same two vectors, calculated
with respect to some inner product. Now for 0 s 1, let gs D sg C .1 s/x g . By
the same reasoning, the rotation index gs . / with respect to gs is also an integer
for each s, so the function f .s/ D gs . / is integer-valued.
In fact, the function f is continuous in s, as can be deduced easily from the
following observations: (1) Our preferred gs -orthonormal frame E1.s/ ; E2.s/
depends continuously on s, as can be seen from the formulas (2.5)–(2.6) used to
implement the Gram–Schmidt algorithm. (2) On every interval Œai1 ; ai where is
smooth, the functions u1 and u2 satisfying the gs -analogue of (9.5) can be expressed
˝ ˛
as uj .t; s/ D Ts .t /; Ej.s/ j.t/ gs , where Ts .t / D 0 .t /=j 0 .t /jgs . Thus u1 and u2 de-
pend continuously on .t; s/ 2 Œai1 ; ai Œ0; 1, so the function .u1 ; u2 / W Œai1 ; ai
The Gauss–Bonnet Formula 273
Proof. Let .a0 ; : : : ; ak / be an admissible partition of Œa; b, and let .x; y/ be oriented
smooth coordinates on an open set U containing x̋ . Let W Œa; b ! R be a tangent
angle function for . Using the rotation index theorem and the fundamental theorem
of calculus, we can write
274 9 The Gauss–Bonnet Theorem
k
X k Z
X ai
2 D .b/ .a/ D "i C 0 .t / dt: (9.7)
iD1 iD1 ai 1
Differentiating 0 (and omitting the t dependence from the notation for simplicity),
we get
0 D rv jE1 j2 D 2hrv E1 ; E1 i;
0 D rv jE2 j2 D 2hrv E2 ; E2 i;
0 D rv hE1 ; E2 i D hrv E1 ; E2 i C hE1 ; rv E2 i:
It follows that the covariant derivatives of the basis vectors are given by
rv E1 D !.v/E2 I
(9.9)
rv E2 D !.v/E1 :
Thus the 1-form ! completely determines the connection in U . (In fact, when the
connection is expressed in terms of the local frame .E1 ; E2 / as in Problem 4-14, this
computation shows that the connection 1-forms are just !2 1 D !1 2 D !, !1 1 D
!2 2 D 0; but it is simpler in this case just to derive the result directly as we have
done.)
Using (9.8) and (9.9), we can compute
The Gauss–Bonnet Formula 275
N D hDt 0 ; N i
D h 0 N; N i C cos hr 0 E1 ; N i C sin hr 0 E2 ; N i
D 0 cos h!. 0 /E2 ; N i C sin h!. 0 /E1 ; N i
D 0 cos2 !. 0 / sin2 !. 0 /
D 0 !. 0 /:
K dA.E1 ; E2 / D K D Rm.E1 ; E2 ; E2 ; E1 /
˝ ˛
D rE1 rE2 E2 rE2 rE1 E2 rŒE1 ;E2 E2 ; E1
˝ ˛
D rE1 .!.E2 /E1 / rE2 .!.E1 /E1 / !.ŒE1 ; E2 /E1 ; E1
˝
D E1 .!.E2 //E1 C !.E2 /rE1 E1 E2 .!.E1 //E1
˛
!.E1 /rE2 E1 !.ŒE1 ; E2 /E1 ; E1
D E1 .!.E2 // E2 .!.E1 // !.ŒE1 ; E2 /
D d!.E1 ; E2 /:
(a) (b)
.M / D V E C F;
Proof. We may as well assume that M is connected, because if not we can prove
the theorem for each connected component and add up the results.
First consider the case in which
R M is orientable. In this case, we can choose an
orientation for M , and then M K dA gives the same result whether we interpret
dA as the Riemannian density or as the Riemannian volume form, so we will use
the latter interpretation for the proof. Let f˝i W i D 1; : : : ; F g denote the faces of the
triangulation, and for each i , let fij W j D 1; 2; 3g be the edges of ˝i and fij W j D
1; 2; 3g its interior angles. Since each exterior angle is minus the corresponding
interior angle, applying the Gauss–Bonnet formula to each triangle and summing
over i gives
F Z
X X 3 Z
F X F X
X 3 F
X
K dA C
N ds C . ij / D 2: (9.11)
iD1 ˝i iD1 j D1 ij iD1 j D1 iD1
Note that each edge integral appears exactly twice in the above sum, with oppo-
site orientations, so the integrals of
N all cancel out. Thus (9.11) becomes
Z F X
X 3
K dA C 3F ij D 2F: (9.12)
M iD1 j D1
Note also that each interior angle ij appears exactly once. At each vertex, the angles
that touch that vertex must have interior measures that add up to 2 (Fig. 9.15); thus
the angle sum can be rearranged to give exactly 2V . Equation (9.12) thus can be
written Z
K dA D 2V F: (9.13)
M
Finally, since each edge appears in exactly two triangles, and each triangle has
exactly three edges, the total number of edges counted with multiplicity is 2E D 3F ,
where we count each edge once for each triangle in which it appears. This means
that F D 2E 2F , so (9.13) finally becomes
Z
K dA D 2V 2E C 2F D 2.M /:
M
278 9 The Gauss–Bonnet Theorem
Now suppose M is nonorientable. Then Proposition B.18 shows that there is an ori-
entable connected smooth manifold M that admits a 2-sheeted smooth covering
yW M
! M , and Exercise A.62 shows that M is compact. If we endow M with the
metric gy D c
y g, then the Riemannian density of gy is given by dA D y dA, and
its Gaussian curvature is Ky D
y K, so c
y .K dA/ D Ky dA. The result of Problem
R R
2-14 shows that M Ky c
dA D 2 K dA.
M
To compare Euler characteristics, we will show that the given triangulation of M
“lifts” to a smooth triangulation of M . To see this, let be any curved triangle in
M and let ˝ be its interior. By definition, this means that there exists a smooth chart
x̋ 2
.U;
'/ whose domain contains and whose image is a disk D R , and such that
' x̋ D x̋0 , where ˝0 is the interior of a curved triangle 0 in R2 . Then ' 1 is an
embedding of D into M , which restricts to a diffeomorphism F W x̋0 ! x̋ . Because
D is simply connected, it follows from Corollary A.57 that ' 1 (and therefore also
F ) has a lift to M , which is smooth because y is a local diffeomorphism; and
because the covering is two-sheeted, there are exactly two such lifts F1 ; F2 . Each
lift is injective because y ı Fi D F , which is injective, and their images are disjoint
because if two lifts agree at a point, they must be identical. From this it is straightfor-
of M with twice
ward to verify that the lifted curved triangles form a triangulation
as many vertices, edges, and faces as that of M , and thus M D 2.M /. Substi-
tuting these relations into the Gauss–Bonnet theorem for M and dividing through
by 2, we obtain the analogous relation for M . t
u
The significance of this theorem cannot be overstated. Together with the classi-
fication theorem for compact surfaces, it gives us very detailed information about
the possible Gaussian curvatures for metrics on compact surfaces. The classification
theorem [LeeTM, Thms. 6.15 and 10.22] says that every compact, connected, ori-
entable 2-manifold M is homeomorphic to a sphere or a connected sum of n tori,
and every nonorientable one is homeomorphic to a connected sum of n copies of the
real projective plane RP 2 ; the number n is called the genus of M . (The sphere is
said to have genus zero.) By constructing simple triangulations, one can show that
The Gauss–Bonnet Theorem 279
Corollary 9.8 Let .M; g/ be a compact Riemannian 2-manifold and let K be its
Gaussian curvature.
(a) If M is homeomorphic to the sphere or the projective plane, then K > 0 some-
where.
(b) If M is homeomorphic to the torus or the Klein bottle, then either K 0 or K
takes on both positive and negative values.
(c) If M is any other compact surface, then K < 0 somewhere. t
u
This corollary has a remarkable converse, proved in the mid-1970s by Jerry Kaz-
dan and Frank Warner: If K is any smooth function on a compact 2-manifold M
satisfying the necessary sign condition of Corollary 9.8, then there exists a Rie-
mannian metric on M for which K is the Gaussian curvature. The proof is a deep
application of the theory of nonlinear partial differential equations. (See [Kaz85] for
a nice expository account.)
In Corollary 9.8 we assumed we knew the topology of M and drew conclusions
about the possible curvatures it could support. In the following corollary we reverse
our point of view, and use assumptions about the curvature to draw conclusions
about the topology of the manifold.
Corollary 9.9 Let .M; g/ be a compact Riemannian 2-manifold and K its Gaussian
curvature.
(a) If K > 0 everywhere on M , then the universal covering manifold of M is
homeomorphic to S2 , and 1 .M / is either trivial or isomorphic to the two-
element group Z=2.
(b) If K 0 everywhere on M , then the universal covering manifold of M is
homeomorphic to R2 , and 1 .M / is infinite.
Proof. Suppose first that M has positive Gaussian curvature. From the Gauss–
Bonnet theorem, M has positive Euler characteristic. The classification theorem
for compact surfaces shows that the only such surfaces are the sphere (with trivial
fundamental group) and the projective plane (with fundamental group isomorphic
to Z=2), both of which are covered by the sphere.
On the other hand, suppose M has nonpositive Gaussian curvature. Then its Euler
characteristic is nonpositive, so it is either an orientable surface of genus n 1 or
a nonorientable one of genus n 2. Thus the universal covering space of M is
R2 if M is the torus or the Klein bottle, and B2 in all other cases (see [LeeTM,
Thm. 12.29]), both of which are homeomorphic to R2 . The fact that the universal
covering space is noncompact implies that the universal covering map has infinitely
many sheets by the result of Exercise A.62, and then Proposition A.61 shows that
1 .M / is infinite. t
u
280 9 The Gauss–Bonnet Theorem
Pf W R.V / ! ƒ2n .V /;
called the Pfaffian, with the property that for every oriented compact Riemannian
2n-manifold M , Z
Pf .Rm/ D .2/n .M /:
M
(Depending on how the Pfaffian is defined, you will see different choices of nor-
malization constants on the right-hand side of this equation.) For example, in four
dimensions, the theorem can be written in terms of familiar curvature quantities as
follows: Z
jRmj2 4jRcj2 C S 2 d Vg D 32 2 .M /: (9.14)
M
In a certain sense, this might be considered a very satisfactory generalization of
Gauss–Bonnet. The only problem with this result is that the relationship between the
Pfaffian and sectional curvatures is obscure in higher dimensions, so it is very hard
to interpret the theorem geometrically. For example, after he proved the version
of the theorem for Euclidean hypersurfaces, Hopf conjectured in the 1930s that
every compact even-dimensional manifold that admits a metric with strictly positive
sectional curvature must have positive Euler characteristic; to date, the conjecture is
known to be true in dimensions 2 and 4, but it is still open in higher dimensions (see
[Pet16, p. 320]).
Problems 281
Problems
where
N is the signed geodesic curvature of @M with respect to the inward-
pointing normal N .
9-11. Suppose g is a Riemannian metric on the cylinder S1 Œ0; 1 such that both
boundary curves are totally geodesic. Prove that the Gaussian curvature of
g either is identically zero or attains both positive and negative values. Give
examples of both possibilities.
9-12. Prove the plane curve classification theorem (Theorem 1.5). [Hint: Show that
every smooth unit-speed plane curve .t / D .x.t /; y.t // satisfies the second-
order ODE 00 .t / D
N .t /N.t /, where N is the unit normal vector field given
by N.t / D .y 0 .t /; x 0 .t //.] (Used on p. 4.)
9-13. Use the four-dimensional Chern–Gauss–Bonnet formula (9.14) to prove that
a compact 4-dimensional Einstein manifold must have positive Euler char-
acteristic unless it is flat.
Chapter 10
Jacobi Fields
Our goal for the remainder of this book is to generalize to higher dimensions some
of the geometric and topological consequences of the Gauss–Bonnet theorem. We
need to develop a new approach: instead of using Stokes’s theorem and differential
forms to relate the curvature to global topology as in the proof of the Gauss–Bonnet
theorem, we study the way that curvature affects the behavior of nearby geodesics.
Roughly speaking, positive curvature causes nearby geodesics to converge, while
negative curvature causes them to spread out (Fig. 10.1). In order to draw topological
consequences from this fact, we need a quantitative way to measure the effect of
curvature on a one-parameter family of geodesics.
We begin by deriving the Jacobi equation, which is an ordinary differential equa-
tion satisfied by the variation field of any one-parameter family of geodesics. A
vector field satisfying this equation along a geodesic is called a Jacobi field. We
then introduce conjugate points, which are pairs of points along a geodesic where
some nontrivial Jacobi field vanishes. Intuitively, if p and q are conjugate along a
geodesic, one expects to find a one-parameter family of geodesic segments that start
at p and end (almost) at q.
After defining conjugate points, we prove a simple but essential fact: the points
conjugate to p are exactly the points where expp fails to be a local diffeomorphism.
We then derive an expression for the second derivative of the length functional
with respect to proper variations of a geodesic, called the second variation formula.
Using this formula, we prove another essential fact about conjugate points: once a
geodesic passes its first conjugate point, it is no longer minimizing. The converse of
this statement, however, is untrue: a geodesic can cease to be minimizing before it
reaches its first conjugate point. In the last section of the chapter, we study the set
of points where geodesics starting at a given point p cease to minimize, called the
cut locus of p.
In the next two chapters, we will derive geometric and topological consequences
of these facts.
Fig. 10.1: Geodesics converge in positive curvature, and spread out in negative curvature
0 D Ds D t T
D Dt Ds T C R.S; T /T
D Dt Dt S C R.S; T /T;
where the last step follows from the symmetry lemma. Evaluating at s D 0, where
S.0; t / D J.t / and T .0; t / D 0 .t /, we get (10.1). t
u
A smooth vector field along a geodesic that satisfies the Jacobi equation is called
a Jacobi field. As the following proposition shows, the Jacobi equation can be writ-
The Jacobi Equation 285
J.a/ D v; Dt J .a/ D w:
Proof. Choose a parallel orthonormal frame .Ei / along , and write v D v i Ei .a/,
w D w i Ei .a/, and 0 .t / D y i .t /Ei .t / in terms of this frame. Writing an unknown
vector field J 2 X. / as J.t / D J i .t /Ei .t /, we can express the Jacobi equation as
JR i .t / C Rj kl i ..t //J j .t /y k .t /y l .t / D 0:
JP i .t / D W i .t /;
WP i .t / D Rj kl i ..t //J j .t /y k .t /y l .t /:
Then Theorem 4.31 guarantees the existence and uniqueness of a smooth solution
on the whole interval I with arbitrary initial conditions J i .a/ D v i , W i .a/ D w i .
Since Dt J.a/ D JP i .a/Ei .a/ D W i .a/Ei .a/ D w, it follows that J.t / D J i .t /Ei .t /
is the desired Jacobi field. t
u
Given a geodesic , let J. / X. / denote the set of Jacobi fields along .
Proof. Because the Jacobi equation is linear, J. / is a linear subspace of X. /. Let
p D .a/ be any point on , and consider the linear map from J. / to Tp M ˚ Tp M
by sending J to .J.a/; Dt J.a//. The preceding proposition says precisely that this
map is bijective. t
u
The following proposition is a converse to Theorem 10.1; it shows that each
Jacobi field along a geodesic segment tells us how some family of geodesics
behaves, at least to first order along .
Proof. Let J be a Jacobi field along . After applying a translation in t (which does
not affect either the fact that is a geodesic or the fact that J is a Jacobi field), we
can assume that the interval I contains 0, and write p D .0/ and v D 0 .0/. Note
that this implies .t / D expp .t v/ for all t 2 I .
Choose a smooth curve W ."; "/ ! M and a smooth vector field V along
satisfying
.0/ D p; V .0/ D v;
0
.0/ D J.0/; Ds V .0/ D Dt J.0/;
If we can show that Dt W .0/ D Dt J.0/ as well, it then follows from the uniqueness
of Jacobi fields that W J , and the proposition is proved.
Formula (10.2) shows that each main curve s .t / is a geodesic whose initial
velocity is V .s/, so
ˇ
@ ˇˇ
@t .s; 0/ D ˇ s .t / D V .s/:
@t tD0
Proposition
10.5 (Local Isometry Invariance of Jacobi Fields). Suppose .M; g/
and M ; gz are Riemannian or pseudo-Riemannian manifolds and ' W M ! M is
a local isometry. Let W I !
M and z W I ! M be geodesics related by z D ' ı ,
and let J 2 X. /, Jz 2 X z be related by d'.t/ .J.t // D Jz.t / for all t 2 I . Then
J is a Jacobi field if and only if Jz is.
Along every geodesic W I ! M , there are always two Jacobi fields that
we can
write down immediately (see Fig. 10.2). Because Dt 0 0 and R 0 ; 0 0 0 by
antisymmetry of R, the vector fields J0 .t / D 0 .t / and J1 .t / D t 0 .t / both satisfy
the Jacobi equation by direct computation. If I is compact or M is complete, the
vector field J0 is the variation field of the variation .s; t / D .s C t /, while J1 is
the variation field of .s; t / D ..1 C s/t /. Therefore, these two Jacobi fields just
reflect the possible reparametrizations of , and do not tell us anything about the
behavior of geodesics other than itself.
To distinguish these trivial cases from more informative ones, we make the
following definitions. Given a regular curve W I ! M , for each t 2 I we let
>
T.t/ M T.t/ M denote the one-dimensional subspace spanned by 0 .t /, and
?
T.t/ M its orthogonal complement. A tangential vector field along is a vector
>
field V 2 X. / such that V .t / 2 T.t/ M for all t , and a normal vector field along
?
is one such that V .t / 2 T.t/ M for all t . Thus a normal Jacobi field along is
a Jacobi field J satisfying J.t / ? 0 .t / for all t . Let X? . / and X> . / denote the
spaces of smooth normal and tangential vector fields along , respectively. When
is a geodesic, J ? . / and J > . / denote the spaces of normal and tangential Jacobi
fields along , respectively.
288 10 Jacobi Fields
only if f and its first derivative vanish at a, which happens if and only if f 0.
Similarly, J is orthogonal to 0 at two points if and only if f vanishes at two points,
which happens if and only if f is identically zero. If this is the case, then f 0 0 as
well, which implies that both J and Dt J are orthogonal to 0 for all t . t
u
Proof. As we noted in the proof of Corollary 10.3, for every a 2 I , the map from
J. / to T.a/ M ˚ T.a/ M given by J 7! .J.a/; Dt J.a// is an isomorphism, and
Proposition 10.7 shows that J ? . / is exactly the preimage of the subspace consist-
ing of all pairs .v; w/ 2 T.a/ M ˚ T.a/ M such that hv; 0 .a/i D hw; 0 .a/i D 0.
Because this subspace has dimension 2n 2, it follows that J ? . / has the same
dimension.
On the other hand, J > . / contains J0 .t / D 0 .t / and J1 .t / D t 0 .t /, which are
linearly independent over R because 0 .t / never vanishes, so it is a subspace of
dimension at least 2. Because J ? . / \ J > . / D f0g, the dimension of J > . / must
be exactly 2, and we have a direct sum decomposition J. / D J ? . / ˚ J > . /.
This implies that every J 2 J. / has a unique decomposition J D J ? C J > , with
J ? 2 J ? . / and J > 2 J > . /. t
u
Basic Computations with Jacobi Fields 289
Proof. The proof of Proposition 10.4 showed that J is the variation field of a vari-
ation of the form (10.2), with any smooth curve satisfying .0/ D p and
0 .0/ D 0, and V a smooth vector field along with V .0/ D v and Ds V .0/ D w. In
this case, we can take .s/ p and V .s/ D v C sw 2 Tp M , yielding (10.5). t
u
This result leads to some explicit formulas for all of the Jacobi fields vanishing
at a point.
where v D 0 .0/, and we regard t w as an element of Ttv .Tp M / by means of the
canonical identification Ttv .Tp M / Š Tp M . If x i are normal coordinates on a
normal neighborhood of p containing the image of , then J is given by the formula
ˇ
J.t / D t w i @i ˇ.t/ ; (10.7)
ˇ
where w i @i ˇ0 is the coordinate representation of w.
Proof. Under the given hypotheses, Lemma 10.9 showed that the restriction of J
to any compact interval containing 0 is the variation field of a variation through
geodesics of the form (10.5). Using the chain rule to compute J.t / D @s .0; t /,
we arrive at (10.6). Because every t in the domain of is contained in some such
compact interval, the formula holds for all such t .
In normal coordinates, the coordinate representation of the exponential map is
the identity, so can be written explicitly in coordinates as
.s; t / D t .v 1 C sw 1 /; : : : ; t .v n C sw n / :
(See Fig. 10.4.) Differentiating .s; t / with respect to s and setting s D 0 shows that
its variation field J is given by (10.7). t
u
1
Proof. Let x i be normal coordinates on 1U . Given q
D q ; : : : ; q n
2 U X fpg and
i n
w D w @i jq 2 Tq M , the curve .t / D t q ; : : : ; t q is a radial geodesic satisfying
.0/ D p and .1/ D q. The previous proposition showed that J.t / D t w i @i j.t/ is
a Jacobi field along . Because J.0/ D 0 and J.1/ D w, the result follows. t
u
where E is any parallel unit normal vector field along , k is an arbitrary constant,
and sc is defined by (10.8). The initial derivative of such a Jacobi field is
Proof. Since g has constant curvature, its curvature endomorphism is given by the
formula of Proposition 8.36:
R.v; w/x D c hw; xiv hv; xiw :
Substituting this into the Jacobi equation, we find that a normal Jacobi field J satis-
fies
0 D Dt2 J C c h 0 ; 0 iJ hJ; 0 i 0
(10.12)
D Dt2 J C cJ;
where we have used the facts that j 0 j2 D 1 and hJ; 0 i D 0.
Since (10.12) says that the second covariant derivative of J is a multiple of J
itself, it is reasonable to try to construct a solution by choosing an arbitrary parallel
unit normal vector field E along and setting J.t / D u.t /E.t / for some function u
to be determined. Plugging this into (10.12), we find that J is a Jacobi field if and
only if u is a solution to the differential equation
u00 .t / C cu.t / D 0:
It is an easy matter to solve this ODE explicitly. In particular, the solutions satisfying
u.0/ D 0 are constant multiples of sc . This construction yields all the normal Jacobi
fields vanishing at 0, since there is an .n 1/-dimensional space of them, and the
space of parallel normal vector fields has the same dimension.
To prove the last two statements, suppose J is given by (10.9), and compute
since sc0 .0/ D 1 in every case. Because E is a unit vector field, jDt J.0/j D jkj, and
(10.11) follows. t
u
Here is our first significant application of Jacobi fields. Because every tangent
vector in a normal neighborhood is the value of a Jacobi field vanishing at the origin
by Corollary 10.11, Proposition 10.12 yields explicit formulas for constant-
curvature metrics in normal coordinates. To set the stage, we will rewrite the
Euclidean metric on Rn in a form that is somewhat more convenient for these comp-
utations.
Let W Rn X f0g ! Sn1 be the radial projection
x
.x/ D ; (10.13)
jxj
Basic Computations with Jacobi Fields 293
gy D g;
V (10.14)
Lemma 10.13. On Rn X f0g, the metric gy defined by (10.14) and the Euclidean
metric gx are related by
gx D dr 2 C r 2 gy; (10.15)
where r.x/ D jxj is the Euclidean distance from the origin.
Proof. Example 2.24 observed that the map ˚ W RC Sn1 ! Rn X f0g given by
˚.; !/ D ! (10.16)
is an isometry when RC Sn1 has the warped product metric d2 ˚ 2 gV and
Rn X f0g has the Euclidean metric (see
1 also2 Problem
2-4).2 Because ˚ 1 .x/ D
2 2
.r.x/; .x//, this means that gx D ˚ d ˚ gV D dr C r gy. t
u
Proof. Let gx denote the Euclidean metric in x-coordinates, and let gc denote the
metric defined by the formula on the right-hand side of (10.17). By the properties
of normal coordinates, at points of U X fpg, all three metrics g, gx, and gc make the
radial vector field @r a unit vector orthogonal to the level sets of r. Thus we need
only show that g.w1 ; w2 / D gc .w1 ; w2 / when w1 ; w2 are tangent to a level set of r,
and by polarization it suffices to show that g.w; w/ D gc .w; w/ for every such
vector w. Note that if w is tangent to a level set r D b, then formulas (10.17) and
(10.15) imply
sc .b/2
gc .w; w/ D sc .b/2 gy.w; w/ D gx.w; w/:
b2
Let q 2 U X fpg and w 2 Tq M , and assume that w is tangent to the r-level
set containing q. Let b D dg .p; q/, and let W Œ0; b ! U be the unit-speed radial
geodesic from p to q, so the coordinate representation of is
t 1 t n
.t / D q ;:::; q ;
b b
1
where q ; : : : ; q n is the coordinate representation of q. Let J 2 X. / be the vector
field along given by
294 10 Jacobi Fields
t i ˇˇ
J.t / D w @i .t/ ; (10.18)
b
where w i @i jq is the coordinate representation for w. By Proposition 10.10, J is a
Jacobi field satisfying Dt J.0/ D .1=b/w i @i jp , and it follows from the definition
that J.b/ D w. Because J is orthogonal to 0 at p and q, it is normal by Proposition
10.7. Thus by Proposition 10.12,
product of the form .0; b/ sc Sn1 , where .0; b/ R has the Euclidean metric, and
Sn1 is the unit sphere with the round metric g.V
Proof. By virtue of Theorem 10.14, we may consider g to be a metric on the ball
of radius b in Rn given by formula (10.17). Let ˚ W .0; b/ Sn1 ! U X fpg and
W Rn X f0g ! Sn1 be the maps defined by (10.16) and (10.13). Because ı
˚ restricts to the identity on fg Sn1 for each fixed , it follows that ˚ gy D
˚ gV D g,
V and thus
˚ g D d2 ˚ sc ./2 g:
V t
u
The next corollary describes a formula for integration in polar coordinates with
respect to a constant-curvature metric. It will be useful in our proofs of volume
comparison theorems in the next chapter.
Corollary 10.17 (Polar Decomposition of Integrals). Suppose .M; g/ is a Rie-
mannian manifold with constant sectional curvature c, and U is an open or closed
geodesic ball of radius b around a point p 2 M . If f W U ! R is any bounded
integrable function, then the integral of f over U can be expressed as
Z Z Z b
f d Vg D f ı ˚.; !/sc ./n1 d d VgV ;
U Sn1 0
Here is another application of the theory of Jacobi fields. Recall that a Riemannian
manifold is a locally symmetric space if every p 2 M has a neighborhood that
admits a point reflection at p. Problem 7-3 showed that every locally symmetric
space has parallel curvature tensor. Now we can prove the converse. The key is the
following lemma due to Élie Cartan. We will use the lemma again in Chapter 12 to
prove a more global result (see Thm. 12.6).
Lemma 10.18. Suppose .M; g/ is a Riemannian manifold with parallel curvature
tensor, and for some points p; py 2 M we are given a linear map A W Tp M ! Tpy M
satisfying A .gpy / D gp and A .Rmpy / D Rmp . Then there exist a neighborhood U
of p and a local isometry ' W U ! M such that '.p/ D py and d'p D A. If M is
complete, then U can be taken to be any normal neighborhood of p.
296 10 Jacobi Fields
Proof. The hypothesis means that rRm 0, and because covariant differentia-
tion commutes with raising and lowering indices, the curvature endomorphism is
also parallel. If M is complete, let U be any normal neighborhood of p; other-
wise choose U D B" .p/, where " > 0 is chosen small enough that both B" .p/
and B" .p/ y are geodesic balls. Our choice guarantees that ' D exppy ıA ı exp1 p is
a smooth map from U into M , and it satisfies '.p/ D py and d'p D A. We will
show that jd'q .x/jg D jxjg for every tangent vector x at every point q 2 U . It then
follows by polarization that hd'q .x/; d'q .y/ig D hx; yig for all x; y 2 Tq M , and
thus ' g D g.
Because d'p D A is a linear isometry, we have jd'p .x/jg D jxjg for x 2 Tp M ,
so we need only consider points q ¤ p. Let q 2 U X fpg and x 2 Tq M be arbitrary.
Let W Œ0; 1 ! U be the radial geodesic from p to q, given explicitly by .t / D
expp .t v/ for some v 2 Tp M . It follows from Corollary 10.11 that there is a Jacobi
field J along such that J.0/D 0 and J.1/ D x; and Proposition 10.10 shows
that it is of the form J.t / D d expp tv .t w/ for some w 2 Tp M . Let vy D A.v/
and w y D A.w/ 2 Tpy M , and define y.t / D exppy .t vy/ and Jy.t / D d.exppy /t vy .t w/
y for
y
t 2 Œ0; 1. Then y is a geodesic from py D '.p/ to '.q/, and J is a Jacobi field along
' and the chain rule (using the fact that dAv D A
y. It follows from the definition of
because A is linear) that d'q ı d expp v D d.exppy /vy ı A, and thus
for some smooth functions Rij k l W Œ0; 1 ! R. The parallel curvature hypothesis and
the fact that each Ei is parallel imply
0
0 D .Dt R/ Ei .t /; Ej .t / Ek .t / D Dt R.Ei .t /; Ej .t //Ek .t / D Rij k l .t / El .t /;
so in fact the coefficients Rij k l .t / are constant in t . The same argument shows
that the curvature
endomorphism has constant coefficients along y in terms of the
frame Eyi . Because our hypotheses guarantee that the coefficients of the two cur-
vature endomorphisms agree at t D 0, they are in fact the same constants along both
geodesics; we write those constants henceforth as Rij k l .
Also, we can write J.t / D J i .t /Ei .t / and Jy.t / D Jyi .t /Eyi .t / for some smooth
functions J i ; Jyi W Œ0; 1 ! R. The Jacobi equations for J and Jy, written in terms of
our parallel frames, read
Conjugate Points 297
l 00
J .t / C Rij k l J i .t /v j v k D 0;
l 00
Jy .t / C Rij k l Jyi .t /v j v k D 0:
Proof. One direction is taken care of by Problem 7-3. To prove the converse, sup-
pose .M; g/ is a Riemannian manifold with rRm 0. Let p 2 M be arbitrary, and
let U be a geodesic ball centered at p. The linear map A D Id W Tp M ! Tp M
satisfies A gp D gp , and for all w; x; y; z 2 Tp M , we have
.A Rmp /.w; x; y; z/ D Rmp .w; x; y; z/ D Rmp .w; x; y; z/:
Conjugate Points
Our next application of Jacobi fields is to study the question of when the exponential
map is a local diffeomorphism.
Suppose .M; g/ is a Riemannian or pseudo-Riemannian manifold and p 2 M .
The restricted exponential map expp is defined on an open subset Ep Tp M , and
because it is a smooth map between manifolds of the same dimension, the inverse
function theorem guarantees that it is a local diffeomorphism
in a neighborhood of
each of its regular points (points v 2 Tp M where d expp v is surjective and thus
invertible). To see where this fails, we need to identify the critical points of expp
(the points where its differential is singular). Proposition 5.19(d) guarantees that 0
is a regular point, but it may well happen that it has critical points elsewhere in Ep .
298 10 Jacobi Fields
order of conjugacy of two points along can be at most n 1. This bound is sharp:
Proposition 10.12 shows that if is a geodesic joining antipodal points p and q on
Sn .R/, then there is a Jacobi field vanishing at p and q for each parallel normal
vector field along ; thus in that case p and q are conjugate to order exactly n 1.
The most important fact about conjugate points is that they are the images of
critical points of the exponential map, as the following proposition shows.
Proof. Suppose first that v is a critical point of expp . Then there is a nonzero vector
w 2 Tv .Tp M / such that d expp v .w/ D 0. Because Tp M is a vector space, we
can identify Tv .Tp M / with Tp M as usual and regard w as a vector in Tp M . Let
be the variation of defined by (10.5), and let J.t / D @s .0; t / be its variation
field. We can compute J.1/ as follows:
ˇ
@ ˇˇ
J.1/ D @s .0; 1/ D ˇ expp .v C sw/ D d expp v .w/ D 0:
@s sD0
Proposition 10.21 (The Two-Point Boundary Problem for Jacobi Fields). Sup-
pose .M; g/ is a Riemannian or pseudo-Riemannian manifold, and W Œa; b ! M
is a geodesic segment. The two-point boundary problem for Jacobi fields along is
uniquely solvable for every pair of vectors v 2 T.a/ M and w 2 T.b/ M if and only
if .a/ and .b/ are not conjugate along .
Our next task is to study the question of which geodesic segments are minimizing.
In the remainder of the chapter, because of the complications involved in studying
lengths on pseudo-Riemannian manifolds, we restrict our attention to the Riemann-
ian case.
In our proof that every minimizing curve is a geodesic, we imitated the first-
derivative test of elementary calculus: if a geodesic is minimizing, then the first
derivative of the length functional must vanish for every proper variation of . Now
we imitate the second-derivative test: if is minimizing, the second derivative must
be nonnegative. First, we must compute this second derivative. In keeping with clas-
sical terminology, we call it the second variation of the length functional.
Differentiating again with respect to s, and using the symmetry lemma and Propo-
sition 7.5, we obtain
d2
2
Lg s jŒai 1 ;ai
ds Z
ai
hDs Dt S; T i hDt S; Ds T i 1 hDt S; T i2hDs T; T i
D C dt
ai 1 hT; T i1=2 hT; T i1=2 2 hT; T i3=2
Z ai
hDt Ds S C R.S; T /S; T i hDt S; Dt S i hDt S; T i2
D C dt:
ai 1 jT j jT j jT j3
Therefore,
Rm.V; 0 ; 0 ; V / D Rm.V ? ; 0 ; 0 ; V ? /:
You should think of I.V; W / as a sort of “Hessian” or second derivative of the length
functional. Because every proper normal vector field along is the variation field
of some proper normal variation, the preceding theorem can be rephrased in terms
of the index form in the following way.
The next proposition gives another expression for I , which makes the role of the
Jacobi equation more evident.
d ˝ ˛
hDt V ; W i D Dt2 V ; W C hDt V ; Dt W i :
dt
Thus, by the fundamental theorem of calculus,
Z ai Z ai ˇai
˝ 2 ˛ ˇ
hDt V ; Dt W i dt D Dt V ; W dt C hDt V ; W iˇˇ :
ai 1 ai 1 ai 1
We can use the second variation formula to prove another extremely important fact
about conjugate points: no geodesic is minimizing past its first conjugate point. The
geometric intuition is as follows. Suppose W Œa; c ! M is a minimizing geodesic
segment, and .b/ is conjugate to .a/ along for some a < b < c. If J is a Jacobi
field along that vanishes at t D a and t D b, then there is a variation of through
geodesics, all of which start at .a/. Since J.b/ D 0, we can expect them to end
“almost” at .b/. If they really did all end at .b/, we could construct a broken
geodesic by following some s from .a/ to .b/ and then following from .b/
to .c/, which would have the same length and thus would also be a minimizing
curve. But this is impossible: as the proof of Theorem 6.4 shows, a broken geodesic
can always be shortened by rounding the corner.
The problem with this heuristic argument is that there is no guarantee that we
can construct a variation through geodesics that actually end at .b/. The proof of
the following theorem is based on an “infinitesimal” version of rounding the corner
to obtain a shorter curve.
Given a geodesic segment W Œa; c ! M , we say that has a conjugate point
if there is some b 2 .a; c such that .b/ is conjugate to .a/ along , and has an
interior conjugate point if there is such a b 2 .a; c/.
Since V satisfies the Jacobi equation on each subinterval Œa; b and Œb; c, and
V .b/ D 0, (10.24) gives
304 10 Jacobi Fields
Similarly,
I.V; W / D hDt V ; W .b/i D jW .b/j2 :
Thus
I.X" ; X" / D 2" jW .b/j2 C "2 I.W; W /:
If we choose " small enough, this is strictly negative. t
u
There is a partial converse to the preceding theorem, which says that a geodesic
without conjugate points has the shortest length among all nearby curves in any
proper variation. Before we prove it, we need the following technical lemma.
Lemma 10.27. Let W Œa; b ! M be a geodesic segment, and suppose J1 and
J2 are Jacobi fields along . Then hDt J1 .t /; J2 .t /i hJ1 .t /; Dt J2 .t /i is constant
along .
Proof. Let f .t / D hDt J1 .t /; J2 .t /i hJ1 .t /; Dt J2 .t /i. Using the Jacobi equation,
we compute
f 0 .t / D hDt2 J1 .t /; J2 .t /i C hDt J1 .t /; Dt J2 .t /i
hDt J1 .t /; Dt J2 .t /i hJ1 .t /; Dt2 J2 .t /i
D Rm J1 .t /; 0 .t /; 0 .t /; J2 .t / C Rm J2 .t /; 0 .t /; 0 .t /; J1 .t / D 0;
where the last equality follows from the symmetries of the curvature tensor. t
u
Theorem 10.28. Let .M; g/ be a Riemannian manifold. Suppose W Œa; b ! M is
a unit-speed geodesic segment without interior conjugate points. If V is any proper
normal piecewise smooth vector field along , then I.V; V / 0, with equality if and
only if V is a Jacobi field. In particular, if .b/ is not conjugate to .a/ along ,
then I.V; V / > 0.
Proof. To simplify the notation, we can assume (after replacing t by t a) that
a D 0. Let p D .0/, and let .w1 ; : : : ; wn / be an orthonormal basis for Tp M , chosen
The Second Variation Formula 305
so that w1 D 0 .0/. For each ˛ D 2; : : : ; n, let J˛ be the unique normal Jacobi field
along satisfying J˛ .0/ D 0 and Dt J˛ .0/ D w˛ .
Our assumption that has no interior conjugate points guarantees that no linear
combination of the J˛ .t /’s can vanish for any t 2 .0; b/, and thus .J˛ .t // forms a
basis for the orthogonal complement of 0 .t / in T.t/ M for each such t . Thus, given
V as in the statement of the theorem, for t 2 .0; b/ we can write
V .t / D v ˛ .t /J˛ .t / (10.25)
for some piecewise smooth functions v ˛ W .0; b/ ! M . (Here and in the remainder
of this proof, the summation convention is in effect, with Greek indices running
from 2 to n.)
˛
In each function v has a piecewise smooth extension to Œ0; b. To see why,
ifact,
let x be the normal coordinates centered at p determined by the basis .wi /. For
sufficiently small t > 0, we can express J˛ .t / and V .t / in normal coordinates as
ˇ
@ ˇˇ
J˛ .t / D t ; ˛ D 2; : : : ; n;
@x˛ ˇ.t/
ˇ
@ ˇˇ
V .t / D v ˛ .t /J˛ .t / D t v ˛ .t / :
@x ˇ
˛ .t/
(The formula for J˛ .t / follows from Prop. 10.10.) Because V is smooth on Œ0; ı/
for some ı > 0 and V .0/ D 0, it follows from Taylor’s theorem that the components
of V .t /=t extend smoothly to Œ0; ı/, which shows that v ˛ is smooth there. Because
V .b/ D 0, it follows similarly that v ˛ extends smoothly to t D b as well. (If J˛ .b/ D
0, the argument is the same as for t D 0, while if not, it is even easier.)
Let .a0 ; : : : ; ak / be an admissible partition for V . On each subinterval .ai1 ; ai /
where V is smooth, define vector fields X and Y along by
X D v ˛ Dt J˛ ; Y D vP ˛ J˛ :
Then Dt V D X C Y on each such interval, and the fact that V is piecewise smooth
implies that Dt V , X , and Y extend smoothly to Œai1 ; ai for each i , with one-sided
derivatives at the endpoints.
To compute I.V; V /, we will use the following identity, which holds on each
subinterval Œai1 ; ai :
d
jDt V j2 Rm.V; 0 ; 0 ; V / D hV; X i C jY j2 : (10.26)
dt
Granting this for now, we use the fundamental theorem of calculus to compute
306 10 Jacobi Fields
k Z
X ai
I.V; V / D jDt V j2 Rm.V; 0 ; 0 ; V / dt
iD1 ai 1
k
X ˇtDai Z b
ˇ
D hV; X iˇ C jY j2 dt;
tDai 1 0
iD1
where the boundary terms are to be interpreted as limits from above and below.
Because X and V are both continuous on Œ0; b, the boundary terms at t Da1 ; : : : ; ak1
all cancel, and because V .0/ D V .b/ D 0, the boundary terms at t D 0 and t D b
Rb
are both zero. It follows that I.V; V / D 0 jY j2 dt 0. If I.V; V / D 0, then Y is
identically zero on .0; b/. Since the J˛ ’s are linearly independent there, this implies
that vP ˛ 0 for each ˛, so each v ˛ is constant. Thus V is a linear combination of
Jacobi fields with constant coefficients, so it is a Jacobi field.
It remains only to prove (10.26). Note that
d
hV; X i D hDt V; X i C hV; Dt X i D hX C Y; X i C hV; Dt X i: (10.27)
dt
The Jacobi equation gives
Therefore,
hDt X; V i D hvP ˛ Dt J˛ ; v ˇ Jˇ i Rm.V; 0 ; 0 ; V /: (10.28)
Because hDt J˛ ; Jˇ i hJ˛ ; Dt Jˇ i D 0 at t D 0, it follows from Lemma 10.27 that
hDt J˛ ; Jˇ i D hJ˛ ; Dt Jˇ i all along . Thus we can simplify the first term in (10.28)
as follows:
Inserting this into (10.28), and then inserting the latter into (10.27) yields
d
hV; X i D hX C Y; X i C hY; X i Rm.V; 0 ; 0 ; V /
dt
D jX C Y j2 jY j2 Rm.V; 0 ; 0 ; V /;
(b) If .a/ and .b/ are conjugate but has no interior conjugate points, then
for every proper normal variation of , the curve s is strictly longer
than for all sufficiently small nonzero s unless the variation field of is
a Jacobi field.
(c) If has no conjugate points, then for every proper normal variation of ,
the curve s is strictly longer than for all sufficiently small nonzero s. u
t
Cut Points
Theorem 10.26 showed that once a geodesic passes its first conjugate point, it ceases
to be minimizing. The converse, however, is not true: a geodesic can cease to be
minimizing without reaching a conjugate point. For example, on the cylinder S1 R
with the product metric, there are no conjugate points along any geodesic; but no
geodesic segment that wraps more than halfway around the cylinder is minimizing
(Fig. 10.10).
Therefore it is useful to make the following definitions. Suppose .M; g/ is a
complete, connected Riemannian manifold, p is a point of M , and v 2 Tp M . Define
the cut time of .p; v/ by
˚
tcut .p; v/ D sup b > 0 W the restriction of v to Œ0; b is minimizing ;
308 10 Jacobi Fields
Proof. Suppose first that 0 < b < c. By definition of tcut .p; v/, there is a time b 0
such that b < b 0 < c and v jŒ0;b 0 is minimizing. Then v .t / cannot be conjugate
to p along v for any 0 < t b (Thm. 10.26), and v jŒ0;b is minimizing because
a shorter admissible curve from p to v .b/ could be combined with v jŒb;b 0 to
produce a shorter admissible curve from p to v .b 0 /, contradicting the fact that
v jŒ0;b 0 is minimizing.
To see that v jŒ0;b is the unique unit-speed minimizing curve between its end-
points, suppose for the sake of contradiction that W Œ0; b ! M is another. Note
that 0 .b/ ¤ v0 .b/, since otherwise and v would agree on Œ0; b by uniqueness of
geodesics. Define a new unit-speed admissible curve z W Œ0; b 0 ! M that is equal to
.t / for t 2 Œ0; b and equal to v .t / for t 2 Œb; b 0 . Then z has length b 0 , so it is also
a minimizing curve from p to v .b 0 /; but it is not smooth at t D b, contradicting the
fact that minimizing curves are smooth geodesics. This completes the proof of (a).
Cut Points 309
Now suppose c < 1. By definition of tcut .p; v/, there is a sequence of times
bi % c such that the restriction of v to Œ0; bi is minimizing. By continuity of the
distance function, therefore,
which shows that v is minimizing on Œ0; c. To prove that one of the options in
(b) must hold, assume that v .c/ is not conjugate to p along v . We will prove the
existence of a second unit-speed minimizing geodesic from p to v .c/.
Let .bi / be a sequence of real numbers such that bi & c. By definition of cut time,
v jŒ0;bi is not minimizing, so for each i there is a unit-speed minimizing geodesic
i W Œ0; ai ! M such that i .0/ D p, i .ai / D v .bi /, and ai < bi . Set wi D i0 .0/ 2
Tp M , so each wi is a unit vector. By compactness of the unit sphere, after passing
to a subsequence we may assume that wi converges to some unit vector w. Since the
ai ’s are all positive and bounded above by b1 , by passing to a further subsequence,
we may also assume that ai converges to some number a. Then by continuity of the
exponential map, i .ai / D expp .ai wi / converges to expp .aw/. But we also know
that i .ai / D v .bi /, which converges to v .c/, so expp .aw/ D v .c/. Moreover,
by continuity of the distance function,
We will show that c tcut .p; v/ b, which implies ci ! tcut .p; v/.
To show that c tcut .p; v/, suppose first that c < 1. By passing to a subse-
quence, we may assume that ci is finite for each i and ci ! c. Proposition 10.32
shows that vi is minimizing on Œ0; ci . By continuity of the exponential map,
exp.pi ; ci vi / ! exp.p; cv/ as i ! 1, and therefore by continuity of the distance
function we have
This shows that v is minimizing on Œ0; c, and therefore tcut .p; v/ c, as claimed.
Now suppose c D 1. Again, by passing to a subsequence, we may assume
ci ! 1. It follows that for every positive number c0 , the geodesic vi is mini-
mizing on Œ0; c0 for i sufficiently large, and it follows by continuity as above that
v is minimizing on Œ0; c0 . Since c0 was arbitrary, this means that tcut .p; v/ D 1.
Next we show that tcut .p; v/ b. If b D 1, there is nothing to prove, so assume
b < 1. Again by passing to a subsequence, we may assume that ci is finite for each
i and ci ! b. By virtue of Proposition 10.32, either there are infinitely many indices
i for which vi .ci / is conjugate to pi along vi , or there are infinitely many i for
which there exists a second minimizing unit-speed geodesic i from pi to vi .ci /.
In the first case, because conjugate points are critical values of the restricted
exponential map, which can be detected in coordinates by the vanishing of a deter-
minant of a matrix of first derivatives, it follows by continuity that v .b/ is also a
critical value, and thus v .b/ is conjugate to p along v . Then Theorem 10.26 shows
that tcut .p; v/ b.
In the second case, let wi be the unit vector in Tpi M such that i D wi . Because
the components of wi with respect to a local orthonormal frame lie in Sn1 , by
passing to a subsequence we may assume .pi :wi / ! .p; w/. If v .b/ is conjugate
to p along v , then tcut .p; v/ b as above, so we may assume that v .b/ is not a
conjugate point. This means that bv is a regular point of the restricted exponential
map expp . Since the set of such regular points is an open subset of TM , there is
some " > 0 such that exppi is injective on the "-neighborhood of ci wi for all i
sufficiently large. This implies that jci wi ci vi jg " for all such i , and therefore
the limits bw and bv are distinct. Thus w jŒ0;b is another minimizing geodesic from
p to v .b/, which by Proposition 10.32 implies that tcut .p; v/ b. t
u
Given p 2 M , we define two subsets of Tp M as follows: the tangent cut locus
of p is the set ˚
TCL.p/ D v 2 Tp M W jvj D tcut .p; v=jvj/ ;
and the injectivity domain of p is
˚
ID.p/ D v 2 Tp M W jvj < tcut .p; v=jvj/ :
It is immediate that TCL.p/ D @ ID.p/, and Cut.p/ D expp TCL.p/ . Further
properties of Cut.p/ and ID.p/ are described in the following theorem.
Cut Points 311
Problems
where Z
n D .sin r/n dr:
0
and .s; b/ 2 M2 for all s. Prove the following generalization of the second
variation formula:
ˇ Z b
d 2 ˇˇ 2
0 0
L . / D jD V j Rm V; ; ; V dt
ds 2 ˇsD0
g s t
a
˝ ˛ ˝ ˛
C II2 .V .b/; V .b//; 0 .b/ II1 .V .a/; V .a//; 0 .a/ ;
where V is the variation field of , and IIi is the second fundamental form
of Mi for i D 1; 2. (Used on p. 365.)
10-10. Prove the following theorem of Theodore Frankel [Fra61], generalizing the
well-known fact that any two great circles on S2 must intersect: Suppose
.M; g/ is a complete, connected Riemannian manifold with positive sec-
tional curvature. If M1 ; M2 M are compact, totally geodesic subman-
ifolds such that dim M1 C dim M2 dim M , then M1 \ M2 ¤ ¿. [Hint:
Assuming that the intersection is empty, show that there exist a shortest
geodesic segment connecting M1 and M2 and a parallel vector field along
that is tangent to M1 and M2 at the endpoints; then apply the second
variation formula of Problem 10-9 to derive a contradiction.]
10-11. Prove Corollary 10.25 (I.V; W / D 0 for all W if and only if V is a Jacobi
field). [Hint: Adapt the proof of Theorem 6.4.]
10-12. Let .M; g/ be a Riemannian manifold. Suppose W Œa; b ! M is a unit-
speed geodesic segment with no interior conjugate points, J is a normal
Jacobi field along , and V is any other piecewise smooth normal vector
field along such that V .a/ D J.a/ and V .b/ D J.b/.
(a) Show that I.V; V / I.J; J /.
(b) Now assume in addition that .b/ is not conjugate to .a/ along .
Show that I.V; V / D I.J; J / if and only if V D J .
10-13. Suppose .M; g/ is a Riemannian manifold and X 2 X.M / is a Killing vec-
tor field (see Problems 5-22 and 6-24). Show that if W Œa; b ! M is any
geodesic segment, then X restricts to a Jacobi field along .
10-14. Suppose P is an embedded submanifold of a Riemannian manifold M , and
W I ! M is a geodesic that meets P orthogonally at t D a for some a 2 I .
A Jacobi field J along is said to be transverse to P if its restriction to each
compact subinterval of I containing a is the variation field of a variation of
through geodesics that all meet P orthogonally at t D a.
(a) Prove that the tangential Jacobi field J.t / D .t a/ 0 .t / along is
transverse to P .
(b) Prove that a normal Jacobi field J along is transverse to P if and
only if
Dt J.a/ D W 0 .a/ .J.a//;
where W 0 .a/ is the Weingarten map of P in the direction of 0 .a/.
316 10 Jacobi Fields
(c) When M has dimension n, prove that the set of transverse Jacobi fields
along is an n-dimensional linear subspace of J. /, and the set of
transverse normal Jacobi fields is an .n 1/-dimensional subspace of
that.
(Used on p. 342.)
10-15. Let x 1 ; : : : ; x n / be any semigeodesic coordinates on an open subset U in a
Riemannian n-manifold .M; g/ (see Prop. 6.41 and Examples 6.43–6.46),
and let .t / D x 1 ; : : : ; x n1 ; t be an xn -coordinate curve defined on some
interval I . Prove that for all constants a1 ; : : : ; an1 , the following vector
field along is a normal Jacobi field along that is transverse to each of
the level sets of x n (in the sense defined in Problem 10-14):
n1
X ˇ
i @ ˇˇ
J.t / D a :
@x i ˇ.t/
iD1
The purpose of this chapter is to show how upper or lower bounds on curvature can
be used to derive bounds on other geometric quantities such as lengths of tangent
vectors, distances, and volumes. The intuition behind all the comparison theorems
is that negative curvature forces geodesics to spread apart faster as you move away
from a point, and positive curvature forces them to spread slower and eventually to
begin converging.
One of the most useful comparison theorems is the Jacobi field comparison
theorem (see Thm. 11.9 below), which gives bounds on the sizes of Jacobi fields
based on curvature bounds. Its importance is based on four observations: first, in a
normal neighborhood of a point p, every tangent vector can be represented as the
value of a Jacobi field that vanishes at p (by Cor. 10.11); second, zeros of Jacobi
fields correspond to conjugate points, beyond which geodesics cannot be minimizing;
third, Jacobi fields represent the first-order behavior of families of geodesics; and
fourth, each Jacobi field satisfies a differential equation that directly involves the
curvature.
In the first section of the chapter, we set the stage for the comparison theorems
by showing how the growth of Jacobi fields in a normal neighborhood is controlled
by the Hessian of the radial distance function, which satisfies a first-order differ-
ential equation called a Riccati equation. We then state and prove a fundamental
comparison theorem for Riccati equations.
Next we proceed to derive some of the most important geometric comparison the-
orems that follow from the Riccati comparison theorem. The first few comparison
theorems are all based on upper or lower bounds on sectional curvatures. Then we
explain how some comparison theorems can also be proved based on lower bounds
for the Ricci curvature. In the next chapter, we will see how these comparison the-
orems can be used to prove significant local-to-global theorems in Riemannian ge-
ometry.
Since all of the results in this chapter are deeply intertwined with lengths and
distances, we restrict attention throughout the chapter to the Riemannian case.
Our main aim in this chapter is to use curvature inequalities to derive consequences
about how fast the metric grows or shrinks, based primarily on size estimates for
Jacobi fields. But first, we need to make one last stop along the way.
The Jacobi equation is a second-order differential equation, but comparison the-
ory for differential equations generally works much more smoothly for first-order
equations. In order to get the sharpest results about Jacobi fields and other geomet-
ric quantities, we will derive a first-order equation, called a Riccati equation, that is
closely related to the Jacobi equation.
Let .M; g/ be an n-dimensional Riemannian manifold, let U be a normal neigh-
borhood of a point p 2 M , and let r W U ! R be the radial distance function as
defined by (6.4). The Gauss lemma shows that the gradient of r on U X fpg is the
radial vector field @r .
On U X fpg, we can form the symmetric covariant 2-tensor field r 2 r (the covari-
ant Hessian of r) and the .1; 1/-tensor field Hr D r.@r /. Because @r D grad r D
.rr/] and r commutes with the musical isomorphisms (Prop. 5.17), we have
]
Hr D r.@r / D r .rr/] D r 2 r :
The Hessian operator also has a close relationship with Jacobi fields.
Proof. Let v D 0 .0/, so jvjg D 1 and .t / D expp .t v/. Proposition 10.10 shows
that
J.t / D d expp tv .t w/;
where Dt J.0/ D w D w i @i jp . Because we are assuming that J is normal, it follows
from Proposition 10.7 that w ? v.
Because w ? v ensures that w is tangent to the unit sphere in Tp M at v, we
can choose a smooth curve W ."; "/ ! Tp M that satisfies j .s/jg D 1 for all
s 2 ."; "/, with initial conditions .0/ D v and 0 .0/ D w. (As always, we are
using the canonical identification between Tv .Tp M / and Tp M .) Define a smooth
family of curves W ."; "/ Œ0; b ! M by .s; t / D expp .t .s//. Then .0; t / D
expp .t v/ D .t /, so is a variation of . The stipulation that j .s/jg 1 ensures
that each main curve s .t / D .s; t / is a unit-speed radial geodesic, so its velocity
satisfies the following identity for all .s; t /:
ˇ
@t .s; t / D .s /0 .t / D @r ˇ .s;t/ : (11.4)
This last expression is the covariant derivative of @t .s; t / along the curve .t/ .s/ D
.s; t / evaluated at s D 0. Since the velocity of this curve at s D 0 is @s .0; t / D
J.t / and @t .s; t / D @r is an extendible vector field by (11.4), we obtain
constant sectional curvature c on U if and only if the following formula holds at all
points of U X fpg:
s 0 .r/
Hr D c r ; (11.6)
sc .r/
where sc is defined by (10.8), and for each q 2 U X fpg, r W Tq M ! Tq M is the
orthogonal projection onto the tangent space of the level set of r (equivalently, onto
the orthogonal complement of @r jq ).
Dt Ji .t / D sc0 .t /Ei .t /:
because r .En .b// D r .@r jq / D 0. Since .Ei .b// is a basis for Tq M , this proves
(11.6).
Conversely, suppose Hr is given by (11.6). Let be a radial geodesic starting at
p, and let J be a normal Jacobi field along that vanishes at t D 0. By Proposi-
tion 11.2, Dt J.t / D Hr J.t / D sc0 .t /J.t /=sc .t /. A straightforward computation then
shows that sc .t /1 J.t / is parallel along . Thus we can write every such Jacobi field
in the form J.t / D ksc .t /E.t / for some constant k and some parallel unit normal
vector field E along . Proceeding exactly as in the proof of Theorem 10.14, we
conclude that g is given by formula (10.17) in these coordinates, and therefore has
constant sectional curvature c. t
u
Jacobi Fields, Hessians, and Riccati Equations 323
˚
For convenience, we record the exact formulas for the quotient sc0 =sc that ap-
peared in the previous proposition (see Fig. 11.1):
1
; if c D 0I
t
sc0 .t / 1 t 1
D cot ; if c D > 0I
sc .t / R R R2
1 t 1
coth ; if c D 2 < 0:
R R R
Now we are in a position to derive the first-order equation mentioned at the begin-
ning of this section. (Problem 11-3 asks you to show, with a different argument, that
the conclusion of the next theorem holds for the Hessian operator of every smooth
local distance function, not just the radial distance function in a normal neighbor-
hood.) This theorem concerns the covariant derivative of the endomorphism field
Hr along a curve . We can compute the action of Dt Hr on every V 2 X. / by
noting that Hr .V .t // is a contraction of Hr ˝ V .t /, so the product rule implies
Dt .Hr .V // D .Dt Hr /V C Hr .Dt V /.
Theorem 11.4 (The Riccati Equation). Let .M; g/ be a Riemannian manifold; let
U be a normal neighborhood of a point p 2 M ; let r W U ! R be the radial distance
function; and let W Œ0; b ! U be a unit-speed radial geodesic. The Hessian oper-
ator Hr satisfies the following equation along j.0;b , called a Riccati equation:
Dt Hr C Hr2 C R 0 D 0; (11.7)
324 11 Comparison Theory
are self-adjoint endomorphism fields along j.a;b that satisfy the following Riccati
equations:
Dt C 2 C D 0; Dt z C z2 C z D 0; (11.8)
where and z are continuous self-adjoint endomorphism fields along satisfying
Then
z.t / .t / for all t 2 .a; b:
To prove this theorem, we will express the endomorphism fields , z, , and z in
terms of a parallel orthonormal frame along . In this frame, they become symmetric
matrix-valued functions, and then the Riccati equations for and z become ordinary
differential equations for these matrix-valued functions. The crux of the matter is the
following comparison theorem for solutions to such matrix-valued equations.
Let M.n; R/ be the space of all n n real matrices, viewed as linear endomor-
phisms of Rn , and let S.n; R/ M.n; R/ be the subspace of symmetric matrices,
corresponding to self-adjoint endomorphisms of Rn with respect to the standard
inner product.
z W .a; b !
Theorem 11.6 (Matrix Riccati Comparison Theorem). Suppose H; H
S.n; R/ satisfy the following matrix Riccati equations:
H 0 C H 2 C S D 0; z0 CH
H z 2 C Sz D 0; (11.10)
Then
z .t / H.t / for all t 2 .a; b:
H (11.13)
z .t / H.t /; 1 z
A.t / D H B.t / D H .t / C H.t / :
2
The hypothesis implies that A extends to a continuous matrix-valued function (still
denoted by A) on Œa; b satisfying A.a/ 0. We need to show that A.t / 0 for all
t 2 .a; b.
326 11 Comparison Theory
Simple computations show that the following equalities hold on .a; b:
A0 D BA C AB Sz C S;
1 z2 z
B0 D H CS CH2 CS :
2
Our hypotheses applied to these formulas imply
A0 BA C AB; (11.14)
B 0 k Id; (11.15)
where the last inequality holds for some (possibly negative) real number k because
z 2 are positive semidefinite and S and Sz are continuous on all of Œa; b and
H 2 and H
thus bounded below. Therefore, for every t 2 .a; b,
Z b
B.t / D B.b/ B 0 .u/ du
t (11.16)
B.b/ .b t /k Id B.b/ C jb aj jkj Id;
which shows that B.t / is bounded above on .a; b. Let K be a constant large enough
that B.t / K Id is negative definite for all such t .
Define a continuous function f W Œa; b Sn1 ! R by
To prove the theorem, we need to show that f .t; x/ 0 for all .t; x/ 2 Œa; b Sn1 .
Suppose this is not the case; then by compactness of Œa; b Sn1 , f takes on a
positive maximum at some .t0 ; x0 / 2 Œa; b Sn1 . Since f .a; x/ D 0 for all x,
we must have a < t0 b. Because hA.t0 /x; xi hA.t0 /x0 ; x0 i for all x 2 Sn1 ,
it follows from Lemma 8.14 that x0 is an eigenvector of A.t0 / with eigenvalue
0 D hA.t0 /x0 ; x0 i > 0.
Since f is differentiable at .t0 ; x0 / and f .t; x0 / f .t0 ; x0 / for a < t < t0 , we
have
@f f .t; x0 / f .t0 ; x0 /
.t0 ; x0 / D lim 0:
@t t%t0 t t0
(We have to take a one-sided limit here to accommodate the fact that t0 might equal
b.) On the other hand, from (11.14) and the fact that A.t0 / and B.t0 / are self-adjoint,
we have
@f ˝ ˛
.t0 ; x0 / D e 2Kt0 A0 .t0 /x0 ; x0 2K hA.t0 /x0 ; x0 i
@t
e 2Kt0 .hB.t0 /A.t0 /x0 ; x0 i C hA.t0 /B.t0 /x0 ; x0 i 2K hA.t0 /x0 ; x0 i/
D e 2Kt0 .2 hA.t0 /x0 ; B.t0 /x0 i 2K hA.t0 /x0 ; x0 i/
D e 2Kt0 .2 0 hx0 ; .B.t0 / K Id/x0 i/ < 0:
These two inequalities contradict each other, thus completing the proof. t
u
Comparisons Based on Sectional Curvature 327
Proof of Theorem 11.5. Suppose , z, , and z are self-adjoint endomorphism fields
along satisfying the hypotheses of the theorem. Let .E1 ; : : : ; En / be a parallel
orthonormal frame along , and let H; H z W .a; b ! S.n; R/ and S; Sz W Œa; b !
S.n; R/ be the symmetric matrix-valued functions defined by
sc0 .r/
Hr r ; (11.17)
sc .r/
where sc and r are defined as in Proposition 11.3, and U0 D U if c 0,
while U0 D fq 2 U W r.q/ < Rg if c D 1=R2 > 0.
(b) If all sectional curvatures of M are bounded below by a constant c, then the
following inequality holds in all of U X fpg:
sc0 .r/
Hr r : (11.18)
sc .r/
Proof. Let x 1 ; : : : ; x n be Riemannian normal coordinates on U centered at p, let
r be the radial distance function on U , and let sc be the function defined by (10.8).
Let U0 U be the subset on which sc .r/ > 0; when c 0, this is all of U , but when
c D 1=R2 > 0, it is the subset where r < R. Let Hrc be the endomorphism field on
U0 X fpg given by
328 11 Comparison Theory
sc0 .r/
Hrc D r :
sc .r/
Let q 2 U0 X fpg be arbitrary, and let W Œ0; b ! U0 be the unit-speed radial
geodesic from p to q. Note that at every point .t / for 0 < t b, the endomorphism
field r can be expressed as r .w/ D w hw; @r i@r D w hw; 0 i 0 , and in this
form it extends smoothly to an endomorphism field along all of . Moreover, since
Dt 0 D 0 along , it follows that Dt r D 0 along as well. Therefore, direct com-
putation using the facts that sc00 D csc and r2 D r shows that Hrc satisfies the
following Riccati equation along j.0;b :
Dt Hr C Hr2 C R 0 D 0:
On the other hand, if the sectional curvatures are bounded below by c, the same
argument with the roles of Hr and Hrc reversed shows that Hr Hrc on U0 . It
remains only to show that U0 D U in this case. If c 0, this is automatic. If c D
1=R2 > 0, then sc0 .r/=sc .r/ ! 1 as r % R; since Hr is defined and smooth in
all of U X fpg and bounded above by Hrc , it must be the case that r < R in U ,
which implies that U0 D U . t
u
for all t 2 Œ0; b2 , where b2 is chosen so that .b2 / is the first conjugate point
to .0/ along if there is one, and otherwise b2 D b.
330 11 Comparison Theory
Under hypothesis (a), it follows from the Hessian comparison theorem that f 0 .t /
0 for all t 2 .0; b0 /, so f .t / is nondecreasing, and thus so is sc .t /1 jJ.t /j. Simi-
larly, under hypothesis (b), we get f 0 .t / 0, which implies that sc .t /1 jJ.t /j is
nonincreasing.
Next we consider the limit of sc .t /1 jJ.t /j as t & 0. Two applications of
l’Hôpital’s rule yield
provided the last limit exists. Since J ! 0, sc ! 0, and sc0 ! 1 as t & 0, this last
limit does exist and is equal to jDt J.0/j2 . Combined with the derivative estimates
above, this shows that the appropriate conclusion
(11.21) or (11.22) holds on .0; b0 /,
and thus by continuity on Œ0; b0 , when Œ0; b0 / is contained in a normal neighbor-
hood of p.
Now suppose is an arbitrary geodesic segment, not assumed to be contained in
a normal neighborhood of p. Let v D 0 .0/, so that .t / D expp .t v/ for t 2 Œ0; b,
and define b0 as above. The definition ensures that .t / is not conjugate to .0/ for
t 2 .0; b0 /, and therefore expp is a local diffeomorphism on some neighborhood of
the set L D ft v W 0 t < b0 g. Let W Tp M be a convex open set containing L on
which expp is a local diffeomorphism, and let gz D expp g, which is a Riemannian
metric on W that satisfies the same curvature estimates as g (Fig. 11.2). By construc-
tion, W is a normal neighborhood of 0, and expp is a local isometry from W; gz
to .M; g/. The curve z.t / D t v for t 2 Œ0; b0 / is a radial geodesic in W , and Propo-
1
sition 10.5 shows that the vector field Jz.t / D d expp tv .J.t // is a Jacobi field
along z thatˇ vanishes
ˇ at t Dˇ 0. Therefore,
ˇ for t 2 .0; ˇb0 /, the
ˇ preceding
ˇ argumentˇ
z t Jz.0/ˇ in case (a) and ˇJz.t /ˇ sc .t /ˇD
implies that ˇJz.t /ˇgz sc .t /ˇD g
z g
z
z t Jz.0/ˇ in
g
z
case (b). This implies that the conclusions of the theorem hold for J on the interval
.0; b0 / and thus by continuity on Œ0; b0 .
To complete the proof, we need to show that b0 b1 in case (a) and b0 b2 in
case (b). Assuming the hypothesis of (a), suppose for contradiction that b0 <b1 . The
Comparisons Based on Sectional Curvature 331
only way this can occur is if .b0 / is conjugate to .0/ along , while sc .b0 />0.
This means that there is a nontrivial normal Jacobi field J 2 J. / satisfying J.0/ D
0 D J.b0 /. But the argument above showed that every such Jacobi field satisfies
jJ.t /j sc .t /jDt J.0/j for t 2 Œ0; b0 and thus jJ.b0 /j sc .b0 /jDt J.0/j > 0, which
is a contradiction. Similarly, in case (b), suppose b0 < b2 . Then sc .b0 / D 0, but .b0 /
is not conjugate to .0/ along . If J is any nontrivial normal Jacobi field along
that vanishes at t D 0, the argument above shows that jJ.t /j sc .t /jDt J.0/j for
t 2 Œ0; b0 , so jJ.b0 /j sc .b0 /jDt J.0/j D 0; but this is impossible because .b0 / is
not conjugate to .0/. t
u
There is a generalization of the preceding theorem, called the Rauch comparison
theorem, that allows for comparison of Jacobi fields in two different Riemannian
manifolds when neither is assumed to have constant curvature. The statement and
proof can be found in [CE08, Kli95].
Because all tangent vectors in a normal neighborhood are values of Jacobi fields
along radial geodesics, the Jacobi field comparison theorem leads directly to the
following comparison theorem for metrics.
Theorem
11.10
(Metric Comparison). Let .M; g/ be a Riemannian manifold, and
let U; x i be any normal coordinate chart for g centered at p 2 M . For each
c 2 R, let gc denote the constant-curvature metric on U X fpg given in the same
coordinates by formula (10.17).
(a) Suppose all sectional curvatures of g are bounded above by a constant c. If
c 0, then for all q 2 U X fpg and all w 2 Tq M , we have g.w; w/ gc .w; w/.
If c D 1=R2 > 0, then the same holds, provided that dg .p; q/ < R.
332 11 Comparison Theory
(b) If all sectional curvatures of g are bounded below by a constant c, then for all
q 2 U X fpg and all w 2 Tq M , we have g.w; w/ gc .w; w/.
Proof. Let q 2 U Xfpg, satisfying the restriction dg .p; q/ < R if we are in case (a)
and c D 1=R2 > 0, but otherwise arbitrary; and let b D dg .p; q/. Given w 2 Tq M ,
we can decompose w as a sum w D y C z, where y is a multiple of the radial
vector field @r and z is tangent to the level set r D b. Then gc .y; z/ D 0 by direct
computation, and the Gauss lemma shows that g.y; z/ D 0, so
Because @r is a unit vector with respect to both g and gc , it follows that g.y; y/ D
gc .y; y/. So it suffices to prove the comparison for z.
There is a radial geodesic W Œ0; b ! U satisfying .0/ D p and .b/ D q, and
Corollary 10.11 shows that z D J.b/ for some Jacobi field J along vanishing at
t D 0, which is normal because it is orthogonal to 0 at t D 0 and t D b. Proposition
i
10.10
1 shows n
iformula J.t / D t a @i j.t/ for some constants
that J has the coordinate
a ; : : : ; a . Since the coordinates x / are normal coordinates for both g and gc ,
it follows that is also a radial geodesic for gc , and the same vector field J is also
a normal Jacobi field for gc along . Therefore, g.z; z/ D jJ.b/j2g and gc .z; z/ D
jJ.b/j2gc . In case (a), our hypothesis guarantees that sc .b/ > 0, so
g.z; z/ D jJ.b/j2g jsc .b/j2 jDt J.0/j2g (Jacobi field comparison theorem)
2
D jsc .b/j jDt J.0/j2gc (since g and gc agree at p)
D jJ.b/j2gc D gc .z; z/ (by Prop. 10:12):
In case (b), the same argument with the inequalities reversed shows that g.z; z/
gc .z; z/. t
u
The next three comparison theorems (Laplacian, conjugate point, and volume
comparisons) can be proved equally easily under the assumption of either an upper
bound or a lower bound for the sectional curvature, just like the preceding theorems.
However, we state these only for the case of an upper bound, because we will prove
stronger theorems later in the chapter based on lower bounds for the Ricci curvature
(see Thms. 11.15, 11.16, and 11.19).
The first of the three is a comparison of the Laplacian of the radial distance func-
tion with its constant-curvature counterpart. Our primary interest in the Laplacian of
the distance function stems from its role in volume and conjugate point comparisons
(see Thms. 11.14, 11.16, and 11.19 below); but it also plays an important role in the
study of various partial differential equations on Riemannian manifolds.
sc0 .r/
r .n 1/ ;
sc .r/
Proof. The case c 0 is covered by Problem 10-7, so assume c D 1=R2 > 0. Let
W Œ0; b ! M be a unit-speed geodesic segment, and suppose J is a nontrivial
normal Jacobi field along that vanishes at t D 0. The Jacobi field comparison
theorem implies that jJ.t /j .constant/ sin.t =R/ > 0 as long as 0 < t < R. t
u
The last of our sectional curvature comparison theorems is a comparison of vol-
ume growth of geodesic balls. Before proving it, we need the following lemma,
which shows how the Riemannian volume form is related to the Laplacian of the
radial distance function.
Lemma 11.13. Suppose .M; g/ is a Riemannian manifold and x i are Riemannian
normal coordinates on a normal neighborhood U of p 2 M . Let det g denote the
determinant of the matrix .gij / in these coordinates, let r be the radial distance
function, and let @r be the unit radial vector field. The following identity holds on
U X fpg: p
r D @r log r n1 det g : (11.23)
Proof. Corollary 6.10 to the Gauss lemma shows that the vector fields grad r and
@r are equal on U X fpg. Comparing the components of these two vector fields in
normal coordinates, we conclude (using the summation convention as usual) that
xi
g ij @j r D :
r
Based on the formula for r from Proposition 2.46, we compute
334 11 Comparison Theory
1 p
r D p @i g ij det g @j r
det g
i
1 x p
Dp @i det g
det g r
i
x 1 xi p
D @i Cp @i det g
r det g r
n1 1 p
D Cp @r det g;
r det g
where the first term in the last line follows by direct computation using r D
P i 2 1=2
i x . This is equivalent to (11.23). u
t
The following result was proved by Paul Günther in 1960 [Gün60]. (Günther also
proved an analogous result in the case of a lower sectional curvature bound, but that
result has been superseded by the Bishop–Gromov theorem, Thm. 11.19 below.)
Proof. The volume estimate (11.24) follows easily from the metric comparison the-
orem, which implies
p the determinants of the metrics g and gc in normal coor-
that p
dinates satisfy det g det gc . If that were all we needed, we could stop here;
but to prove the other statements, we need a more involved argument, which inci-
dentally provides another proof of (11.24) that does not rely directly on the metric
comparison theorem, and therefore can be adapted more easily to the case in which
we have only
an estimate of the Ricci curvature (see Thm. 11.19 below).
Let x i be normal coordinates on Bı0 .p/ (interpreted as all of M if ı0 D 1).
Using these coordinates, we might as well consider g to be a Riemannian metric
on an open subset of Rn and p to be the origin. Let gx denote the Euclidean metric
in these coordinates, and let gc denote the constant-curvature metric in the same
coordinates, given on the complement of the origin by (10.17).
The Laplacian comparison theorem together with Lemma 11.13 shows that
p s 0 .r/
@r log r n1 det g D r .n 1/ c D @r log sc .r/n1 : (11.25)
sc .r/
Comparisons Based on Sectional Curvature 335
p
Thus log r n1 det g=sc .r/n1 p is a nondecreasing function of r along each radial
geodesic, and so is the ratio r n1 p det g=sc .r/n1 . To compute the limit as r & 0,
note that gij D ıij at the origin, so det g convergesp uniformly to 1 as r & 0. Also,
n1
for every c, we have sc .r/=r
p ! 1 as r & 0, so r det g=sc .r/n1 ! 1.
We can write d Vg D det g d Vgx . Corollary 10.17 in the case c D 0 shows that
Z Z ı p
Vg .ı/ D det g ı ˚.; !/n1 d d VgV ;
Sn1 0
where ˚.; !/ D ! for 2 .0; ı0 / and ! 2 Sn1 . The same corollary shows that
Z Z Z ! Z
ı ı
Vc .ı/ D sc ./n1 d d VgV D sc ./n1 d d VgV :
Sn1 0 0 Sn1
Therefore,
Z Z ı p
det g ı ˚.; !/n1 d d VgV
Vg .ı/ Sn1 0
D Z ! Z
Vc .ı/ ı
sc ./n1 d d VgV
0
˙Z ı
Sn1
n1
Z .; !/sc ./ d
1 0
D Z ı d VgV ; (11.26)
Vol.Sn1 / Sn1
sc ./n1 d
0
The argument above (together with the fact that r ı ˚ D ) shows that .; !/ is a
nondecreasing function of for each !, which approaches 1 uniformly as & 0.
We need to show that the quotient in parentheses in the last line of (11.26) is also
nondecreasing as ı increases. Suppose 0 < ı1 < ı2 . Because is nondecreasing, we
have
Z ı1 Z ı2
n1
.; !/sc ./ d sc ./n1 d
0 ı1
Z ! Z !
ı1 ı2
sc ./n1 d .ı1 ; !/ sc ./n1 d
0 ı1
Z ı1 Z ı2
sc ./n1 d .; !/sc ./n1 d;
0 ı1
and therefore (suppressing the dependence of the integrands on and ! for brevity),
336 11 Comparison Theory
Z ı1 Z ı2
scn1 d scn1 d
0 0
Z Z Z !
ı1 ı1 ı2
D scn1 d scn1 d C scn1 d
0 0 ı1
Z ı1 Z ı1 Z ı1 Z ı2
scn1 d scn1 d C scn1 d scn1 d
0 0 0 ı1
Z ı1 Z ı2
D scn1 d scn1 d:
0 0
Evaluating (11.26) at ı1 and ı2 and inserting the inequality above shows that the
ratio Volg .ı/= Volc .ı/ is nondecreasing as a function of ı, and it approaches 1 as
ı & 0 because .; !/ ! 1 as & 0. It follows that Vg .ı/ Vc .ı/ for all ı 2 .0; ı0 .
It remains only to consider the case in which the volume ratio is equal to 1 for
some ı. If .; !/ is not identically 1 on the set where 0 < < ı, then it is strictly
greater than 1 on a nonempty open subset, which implies that the volume ratio
in (11.26) is strictly greater than 1; so Volg .ı/ D Volc .ı/ implies
p .; !/ 1 on
.0; ı/ Sn1 , and pulling back to U via ˚ 1 shows that r n1 det g sc .r/n1 on
Bı .p/. By virtue of (11.25), we have r .n 1/sc0 .r/=sc .r/, or in other words,
tr.Hr / D tr..sc0 .r/=sc .r//r /.It follows from
the Hessian comparison theorem that
the endomorphism field Hr sc0 .r/=sc .r/ r is positive semidefinite, so its eigen-
values are all nonnegative. Since its trace is zero, the eigenvalues must all be zero.
In other words, Hr .sc0 .r/=sc .r//r on the geodesic ball Bı .p/. It then follows
from Proposition 11.3 that g has constant sectional curvature c on that ball. t
u
All of our comparison theorems so far have been based on assuming an upper or
lower bound for the sectional curvature. It is natural to wonder whether anything can be
said if we weaken the hypotheses and assume only bounds on other curvature quan-
tities such as Ricci or scalar curvature.
It should be noted that except in very low dimensions, assuming a bound on Ricci
or scalar curvature is a strictly weaker hypothesis than assuming one on sectional
curvature. Recall Proposition 8.32, which says that on an n-dimensional Rieman-
nian manifold, the Ricci curvature evaluated on a unit vector is a sum of n 1 sec-
tional curvatures, and the scalar curvature is a sum of n.n 1/ sectional curvatures.
Thus if .M; g/ has sectional curvatures bounded below by c, then its Ricci curvature
satisfies Rc.v; v/ .n 1/c for all unit vectors v, and its scalar curvature satisfies
S n.n 1/c, with analogous inequalities if the sectional curvature is bounded
above. However, the converse is not true: an upper or lower bound on the Ricci cur-
vature implies nothing about individual sectional curvatures, except in dimensions
2 and 3, where the entire curvature tensor is determined by the Ricci curvature (see
Comparisons Based on Ricci Curvature 337
Cors. 7.26 and 7.27). For example, in every even dimension greater than or equal to
4, there are compact Riemannian manifolds called Calabi–Yau manifolds that have
zero Ricci curvature but nonzero sectional curvatures (see, for example, [Bes87,
Chap. 11]).
In this section we investigate the extent to which bounds on the Ricci curva-
ture lead to useful comparison theorems. The strongest theorems of the preceding
section, such as the Hessian, Jacobi field, and metric comparison theorems, do not
generalize to the case in which we merely have bounds on Ricci curvature. How-
ever, it is a remarkable fact that Laplacian, conjugate point, and volume comparison
theorems can still be proved assuming only a lower (but not upper) bound on the
Ricci curvature. (The problem of drawing global conclusions from scalar curvature
bounds is far more subtle, and we do not pursue it here. A good starting point for
learning about that problem is [Bes87].)
The next theorem is the analogue of Theorem 11.11.
sc0 .r/
r .n 1/ ; (11.27)
sc .r/
Proof. Let q 2 U0 X fpg be arbitrary, and let W Œ0; b ! U0 be the unit-speed radial
geodesic from p to q. We will show that (11.27) holds at .t/ for 0 < t b.
Because covariant differentiation commutes with the trace operator (since it is
just a particular kind of contraction), the Riccati
equation
(11.7) for Hr implies the
following scalar equation along for r D tr Hr :
d
r C tr Hr2 C tr R 0 D 0: (11.28)
dt
We need to analyze the last two terms on the left-hand side. We begin with the
last term. In terms of any local orthonormal frame .Ei /, we have
n
X n n
˝ ˛ X ˝ ˛ X
tr R 0 D R 0 .Ei /; Ei D R.Ei ; 0 / 0 ; Ei D Rm.Ei ; 0 ; 0 ; Ei /
iD1 iD1 iD1
D Rc. 0 ; 0 /:
r
HV r D Hr r : (11.29)
n1
338 11 Comparison Theory
We compute
r r .r/2 2
tr HV 2r D tr Hr2 tr.Hr ı r / tr.r ı Hr / C tr r :
n1 n1 .n 1/2
To simplify this, note that r2 D r because r is a projection, and thus tr r2 D
tr.r / D n 1. Also, Hr .@r / 0 implies that Hr ı r D Hr ; and since Hr is self-
adjoint, hHr v; @r ig D hv; Hr @r ig D 0 for all v, so the image of Hr is contained in
the orthogonal complement of @r , and it follows
that r ı Hr D Hr as well. Thus
the last three terms in the formula for tr HV 2r combine to yield
.r/2
tr HV 2r D tr Hr2 :
n1
Solving this for tr Hr2 , substituting into (11.28), and dividing by n 1, we obtain
d r r 2 tr HV 2r C Rc. 0 ; 0 /
C C D 0: (11.30)
dt n1 n1 n1
H 0 .t / C H.t /2 C c 0:
We wish to apply the 1 1 case of the matrix Riccati comparison theorem (Thm.
11.6) with H.t / as above, S.t / c,
ˇ ˇ
.r/ˇ.t/ tr HV 2r ˇ.t/ C Rc. 0 .t /; 0 .t //
Hz .t / D ; and S.t/z D :
n1 n1
Note that HV 2r is positive semidefinite, which means that all of its eigenvalues are
nonnegative, so its trace (which is the sum of the eigenvalues) is also nonnegative.
(This is the step that does not work in the case of an upper bound on Ricci curvature.)
Thus our hypothesis on the Ricci curvature guarantees that S.t/ z c for all t 2 .0; b.
To apply Theorem 11.6, we need to verify that Sz has a continuous exten-
sion to Œ0; b and that H z .t / H.t / has a nonnegative limit as t & 0. Recall
that we showed in (11.19) and (11.20) that sc0 .r/=sc .r/ D 1=r C O.r/ and Hr D
.1=r/r C O.r/ as r & 0. This implies that r D tr.Hr / D .n 1/=r C O.r/,
and therefore both HV r j.t/ and H z .t / H.t / approach 0 and Sz.t / approaches
Rc. 0 .0/; 0 .0//=.n 1/ c as t & 0. Therefore, we can apply Theorem 11.6 to
conclude that H z .t / H.t / for t 2 .0; b. Since .b/ D q was arbitrary, this com-
pletes the proof. t
u
The next theorem and its two corollaries will be crucial ingredients in the proofs
of our theorems in the next chapter about manifolds with positive Ricci curvature
(see Thms. 12.28 and 12.24).
Comparisons Based on Ricci Curvature 339
Suppose U contains a point q where r R, and let W Œ0; b ! U be the unit-speed
radial geodesic from p to q. Because sc .R/ D 0, (11.31) shows that det g..t // !
0 as t % R, and therefore by continuity det g D 0 at .b/, which contradicts the
fact that det g > 0 in every coordinate neighborhood. The upshot is that no normal
neighborhood can include points where r R.
Now suppose W Œ0; b ! M is a unit-speed geodesic with b R, and assume
for the sake of contradiction that has no conjugate points. Let pD.0/ and vD 0 .0/,
so .t / D expp .t v/ for t 2 Œ0; b. As in the proof of Theorem 11.9, because has no
conjugate points, we can choose a star-shaped open subset W Tp M containing
the set L D ft v W 0 t < b0 g Tp M on which expp is a local diffeomorphism,
and let gz be the pulled-back metric expp g on W , which satisfies the same curvature
estimates as g. Then z.t / D t v is a radial gz-geodesic in W of length greater than or
equal to R, which contradicts the argument in the preceding paragraph. t
u
Proof. Every radial geodesic segment in a geodesic ball is minimizing, but the pre-
ceding theorem shows that no geodesic segment of length R or greater is minimiz-
ing. Thus no geodesic ball has radius greater than R. t
u
Proof. This follows from the fact that any two points of M can be connected by a
minimizing geodesic segment, and the conjugate point comparison theorem implies
that no such segment can have length greater than R. t
u
340 11 Comparison Theory
Our final comparison theorem is a powerful volume estimate under the assump-
tion of a lower bound on the Ricci curvature. We will use it in the proof of Theorem
12.28 in the next chapter, and it plays a central role in many of the more advanced
results of Riemannian geometry.
A weaker version of this result was proved by Paul Günther in 1960 [Gün60]
for balls within the injectivity radius under the assumption of a lower bound on
sectional curvature; it was improved by Richard L. Bishop in 1963 (announced in
[Bis63], with a proof in [BC64]) to require only a lower Ricci curvature bound; and
then it was extended by Misha Gromov in 1981 [Gro07] to cover all metric balls in
the complete case, not just those inside the injectivity radius.
where det g denotes the determinant of the matrix of g in the normal coordinates
determined by the choice of basis.
Let ˚ W RC Sn1 ! Tp M X f0g Š Rn X f0g be the map ˚.; !/ D ! as in
Corollary 10.17, and define zsc W RC ! Œ0; 1/ by zsc ./ D sc ./ if c 0, while in
Comparisons Based on Ricci Curvature 341
Corollary 11.18 shows that the cut time of every unit vector in Tp M is less than or
equal to R. Thus zsc ./ > 0 whenever ˚.; !/ 2 ID.p/, and we can define z W RC
n1
S ! R by
p
det g ı ˚.; !/n1
z !/ D ; ˚.; !/ 2 ID.p/;
.; zsc ./n1
0; ˚.; !/ … ID.p/;
The arguments of Theorem 11.14 (with inequalities reversed) show that for each
z !/ is nonincreasing in for all positive , and it follows
! 2 Sn1 , the function .;
just as in that proof that Vg .ı/=Vc .ı/ (now interpreted as a ratio of volumes of
metric balls) is nonincreasing for all ı > 0 and approaches 1 as ı & 0, and (11.32)
follows.
Finally, suppose Vg .ı/ D Vc .ı/ for some ı > 0, and assume first that ı inj.p/.
An argument exactly analogous to the one at the end of the proof of Theorem 11.14
shows that .; !/ 1 everywhere on the set where 0 < < ı. Combined with
Lemma 11.13, this implies that
sc0 .r/
r D .n 1/ (11.33)
sc .r/
everywhere on Bı .p/ X fpg. This means that along each unit-speed radial geodesic
, the function u.t / D .r/j.t/ =.n 1/ D sc0 .t /=sc .t / satisfies u0 C u2 C c 0 by
direct computation. Comparing this to (11.30), we conclude that
tr HV 2r C Rc. 0 ; 0 /
c
n1
on Bı .p/Xfpg. Since
tr HV 2r 0 and Rc. 0 ; 0 / .n1/c everywhere, this is pos-
sible only if tr HV 2r vanishes identically there. Because HV 2r is positive semidefinite
and its trace is zero, it must vanish identically, which by definition of HV r means
342 11 Comparison Theory
r s 0 .r/
Hr D r D c r :
n1 sc .r/
Proposition 11.3 then shows that g has constant sectional curvature c on Bı .p/.
Now suppose .M; g/ is complete. The argument of Theorem 11.14 then shows
z !/ 1 everywhere on the set where 0 < < ı and zsc ./ > 0. In view of
that .;
z this implies in particular that ID.p/ \ Bı .0/ contains all of the
the definition of ,
points in Bı .0/ where zsc .r/ > 0. In case c 0, zsc .r/ D sc .r/ > 0 everywhere, so
ID.p/ \ Bı .0/ D Bı .0/ and therefore the metric ball of radius ı around p is actually
a geodesic ball, and the argument above applies.
In case c D 1=R2 > 0, if ı R, then zsc .r/ D sc .r/ > 0 on Bı .0/, and once
again we conclude that the metric ı-ball is a geodesic ball. On the other hand, if
ı > R, then the diameter comparison theorem (Cor. 11.18) shows that the metric
ball of radius ı is actually the entire manifold. The fact that the volume ratio is
nonincreasing implies that Vg .R/ D Vc .R/, and the argument above shows that
g has constant sectional curvature c on the metric ball of radius R. Since the
closure of that ball is all of M , the result follows by continuity. t
u
The next corollary is immediate.
(For explicit formulas for the volumes of n-spheres, see Problem 10-4.)
Problems
11-3. Let .M; g/ be a Riemannian manifold, and let f be any smooth local dis-
tance function defined on an open subset U M . Let F D grad f (so the
integral curves of F are unit-speed geodesics), and let Hf D rF (the Hes-
sian operator of f ). Show that Hf satisfies the following Riccati equation
along each integral curve of F :
Dt Hf C Hf2 C R 0 D 0; (11.34)
u00 .t / C cu.t / D 0;
u.0/ D 1; (11.35)
u0 .0/ D :
In this final chapter, we bring together most of the tools we have developed so far to
prove some significant local-to-global theorems relating curvature and topology of
Riemannian manifolds.
We focus first on constant-curvature manifolds. The main result here is the
Killing–Hopf theorem, which shows that every complete, simply connected mani-
fold with constant sectional curvature is isometric to one of our frame-homogeneous
model spaces. A corollary of the theorem then shows that the ones that are not sim-
ply connected are just quotients of the models by discrete groups of isometries. The
technique used to prove this result also leads to a global characterization of sym-
metric spaces in terms of parallel curvature tensors.
Next we turn to nonpositively curved manifolds. The primary result is the
Cartan–Hadamard theorem, which topologically characterizes complete, simply
connected manifolds with nonpositive sectional curvature: they are all diffeomor-
phic to Rn . After proving the main result, we prove two other theorems, due to
Cartan and Preissman, that place severe restrictions on the fundamental groups of
complete manifolds with nonpositive or negative curvature.
Finally, we address the case of positive curvature. The main theorem is Myers’s
theorem, which says that a complete manifold with Ricci curvature bounded below
by a positive constant must be compact and have a finite fundamental group and
diameter no larger than that of the sphere with the same Ricci curvature. The bor-
derline case is addressed by Cheng’s maximal diameter theorem, which says that the
diameter bound is achieved only by a round sphere. These results are supplemented
by theorems of Milnor and Synge, which further restrict the topology of manifolds
with nonnegative Ricci and positive sectional curvature, respectively.
10.15 shows that each point of M has a neighborhood that is isometric to an open
subset of one of the constant-curvature model spaces of Chapter 3. In order to turn
that into a global result, we use a technique modeled on the theory of analytic con-
tinuation in complex analysis.
Suppose .M; g/ and M ; gy are Riemannian manifolds of the same dimension
and ' W U ! M is a local isometry defined on some connected open subset U M .
If W Œ0; 1 ! M is a continuous path such that .0/ 2 U , then an analytic continua-
tion of ' along is a family of pairs f.Ut ; 't / W t 2 Œ0; 1g, where Ut is a connected
neighborhood of .t / and 't W Ut ! M is a local isometry, such that '0 D ' on
U0 \ U , and for each t 2 Œ0; 1 there exists ı > 0 such that jt t1 j < ı implies that
.t1 / 2 Ut and that 't agrees with 't1 on Ut \ Ut1 (see Fig. 12.1). Note that we
require 't and 't1 to agree where they overlap only if t and t1 are sufficiently close;
in particular, if is not injective, the values of the analytic continuation may dif-
fer at different times at which returns to the same point. However, as the next
lemma shows, all analytic continuations along the same path will end up with the same
value.
Lemma 12.1 (Uniqueness of Analytic Continuation). With .M; g/, M ; gy , and '
as above, if f.Ut ; 't /g and f.Ut0 ; 't0 /g are two analytic continuations of ' along the
same path W Œ0; 1 ! M , then '1 D '10 on a neighborhood of .1/.
Proof. Let T be the set of all t 2 Œ0; 1 such that 't D 't0 on a neighborhood of .t /.
Then 0 2 T , because both '0 and '00 agree with ' on a neighborhood of .0/. We
will show that T is open and closed in Œ0; 1, from which it follows that 1 2 T , which
proves the lemma.
To see that it is open, suppose t 2 T . By the definition of analytic continuation,
there is some ı > 0 such that if t1 2 Œ0; 1 and jt1 t j < ı, then 't1 and 't01 both agree
with 't on a neighborhood of .t /, so Œ0; 1 \ .t ı; t C ı/ T . Thus T is open.
To see that it is closed, suppose t is a limit point of T . There is a sequence ti ! t
such that ti 2 T , which means that 'ti D 't0i on a neighborhood of ti for each i .
By definition of analytic continuation, for i large enough, we have .ti / 2 Ut \ Ut0
Manifolds of Constant Curvature 347
and 't D 'ti D 't0i D 't0 on a neighborhood of .ti /. In particular, this means that
't ..ti // D 't0 ..ti // and d.'t /.ti / D d.'t0 /.ti / for each such i . By continuity,
therefore, 't ..t // D 't0 ..t // and d.'t /.t/ D d.'t0 /.t/ . Proposition 5.22 shows
that 't D 't0 on a neighborhood of .t /, so t 2 T . t
u
The previous lemma shows that all analytic continuations along the same path
end up with the same value. But when we consider analytic continuations along dif-
ferent paths, this may not be the case. The next theorem gives a sufficient condition
for analytic continuations along two different paths to result in the same value.
Theorem
12.2 (Monodromy Theorem for Local Isometries). Let .M; g/ and
M ; gy be connected Riemannian manifolds. Suppose U is a neighborhood of a
point p 2 M and ' W U ! M is a local isometry that can be analytically continued
along every path starting at p. If 0 ; 1 W Œ0; 1 ! M are path-homotopic paths from
p to a point q 2 M , then the analytic continuations along 0 and 1 agree in a
neighborhood of q.
The next theorem, due to Wilhelm Killing and Heinz Hopf, is the main result of
this section.
12-2), the situation in odd dimensions is far more complicated. Already in dimen-
sion 3 there are many interesting examples: some notable ones are the lens spaces
obtained as quotients of S3 C 2 by cyclic groups rotating the two complex coordi-
nates through different angles, and the quotients of SO.3/ (which is diffeomor-
phic to RP 3 and is therefore already a quotient of S3 ) by the dihedral groups,
the symmetry groups of regular 3-dimensional polyhedra.
The complete classification of Euclidean space forms is known only in low
dimensions. Problem 12-3 shows that every compact 2-dimensional Euclidean space
form is diffeomorphic to the torus or the Klein bottle. It turns out that there are 10
classes of nondiffeomorphic compact Euclidean space forms of dimension 3, and 75
classes in dimension 4. The fundamental groups of compact Euclidean space forms
are examples of crystallographic groups—discrete groups of Euclidean isometries
with compact quotients, which have been studied extensively by physicists as well
as geometers. (A quotient of Rn by a crystallographic group is a space form, pro-
vided it is a manifold, which is true whenever the group is torsion-free.) In higher
dimensions, the classification is still elusive. The main things that are known are
two results proved by Ludwig Bieberbach in the early twentieth century: (1) every
Euclidean space form is a quotient of a flat torus by a finite group of isometries,
and (2) in each dimension, there are only finitely many diffeomorphism classes of
Euclidean space forms. See [Wol11] for proofs of the Bieberbach theorems and a
complete survey of the state of the art.
In addition to the question of classifying Euclidean space forms up to diffeomor-
phism, there is also the question of classifying the different flat Riemannian metrics
on a given manifold up to isometry. Problem 12-5 shows, for example, that there is
a three-parameter family of nonisometric flat Riemannian metrics on the 2-torus.
Finally, the study of hyperbolic space forms is a vast and rich subject in its own
right. A good introduction is the book [Rat06].
The technique of analytic continuation of local isometries can also be used to derive
a fundamental global result about symmetric spaces. Theorem 10.19 showed that
locally symmetric spaces are characterized by having parallel curvature. The next
theorem is a global version of that result.
Proof. Theorem 10.19 shows that (a) , (b). To show that (a) ) (c), assume that
M has parallel curvature tensor, and let M ; gz be its universal covering manifold
with the pullback metric. Since the covering map W M ! M is a local isometry,
M also has parallel curvature tensor, and Corollary 6.24 shows that it is complete.
Then Theorem 12.6 shows that M is a symmetric space. If we let denote the covering
automorphism group with the discrete topology, then Proposition C.20 shows that
is a discrete Lie group acting smoothly, freely, and properly on M , and it acts
isometrically because the pullback metric is invariant under all covering automor-
phisms. Finally, Corollary 2.33 shows that M is isometric to M = .
Finally, we show that (c) ) (a). If M is isometric to a Riemannian quotient of the
form M = as in (c), then the quotient map M !M = is a Riemannian covering
by Proposition 2.32. Since M has parallel curvature tensor by the result of Problem
7-3, so does M . t
u
An extensive treatment of the structure and classification of symmetric spaces
can be found in [Hel01].
The analytic continuation technique used in this section was introduced in 1926
by Élie Cartan, in the same paper [Car26] in which he first defined symmetric spaces
and proved many of their properties. The technique was generalized in 1956 by War-
ren Ambrose [Amb56] to allow for isometries between more general Riemannian
manifolds with varying curvature; and then in 1959 it was generalized further by
Noel Hicks [Hic59] to manifolds with connections in their tangent bundles, not nec-
essarily Riemannian ones. The resulting theorem is known as the Cartan–Ambrose–
Hicks theorem. We do not pursue it further, but a statement and proof can be found
in [CE08].
352 12 Curvature and Topology
Proof. These properties follow from Lemma 6.8, Proposition 6.11, and Corollaries
6.12 and 6.13, together with the fact that M is a normal neighborhood of each of its
points. t
u
Manifolds of Nonpositive Curvature 353
With a little more work, we can obtain a much more powerful result.
Proof. Suppose for the sake of contradiction that M is simply connected and g is a
metricon M N with nonpositive sectional curvature. If Nz is the universal covering
manifold of N , then there is a universal covering map W M Nz ! M N , and
g is a complete metric of nonpositive sectional curvature on M Nz . The Cartan–
Hadamard theorem shows that M Nz is diffeomorphic to a Euclidean space.
Since M is simply connected, it is orientable (Cor. B.19). Let be a smooth
orientation m-form for M , where m D dim M . Then is closed because there are
no nontrivial
R .m C 1/-forms on M , but Stokes’s theorem shows that is not exact,
because M > 0. On the other hand, if we let p W M Nz ! M denote the projec-
tion on the first factor, then the form p is exact on M Nz , because every closed
m-form on a Euclidean space is exact.
Choose an arbitrary point y0 2 Nz and define W M ! M Nz by .x/ D .x; y0 /;
it is a diffeomorphism onto its image .M / D M fy0 g, which is a compact em-
bedded submanifold of M Nz . By diffeomorphism invariance of the integral,
354 12 Curvature and Topology
Z Z Z Z
p D .p / D .p ı / D > 0;
.M / M M M
R
because p ı D Id. But p is exact on M Nz , which implies .M / p
D 0, a
contradiction. t
u
Example 12.13. By the preceding corollary, every metric on Sm Sn must have posi-
tive sectional curvature somewhere, provided m 2 and n 1, because both spheres
are compact and Sm is simply connected. //
With a little more algebraic topology, one can obtain more information. A topo-
logical space whose higher homotopy groups k .M / all vanish for k > 1 is called
aspherical.
Proof. We can write f .x/ D 12 r.x/2 , where r is the radial distance function from
q with respect to any normal coordinates, and Proposition 12.9 shows that f is
smooth on all of M . The Hessian comparison theorem implies that Hr .1=r/r
on M X fqg, and therefore
The expression on the right-hand side above is actually equal to the identity operator,
as can be checked by applying it separately to @r D grad r and an arbitrary vector
Manifolds of Nonpositive Curvature 355
perpendicular to @r . Therefore hHf .v/; vi jvj2 for all tangent vectors to M X fqg,
and by continuity, the same holds at q as well.
Now suppose W Œ0; 1 ! M is a geodesic segment. By the chain rule,
d
f ..t // D df.t/ . 0 .t // D hgrad f j.t/ ; 0 .t /i;
dt
and thus (using the fact that 0 is parallel along )
d2
f ..t // D hr 0 .t/ grad f j.t/ ; 0 .t /i D hHf . 0 .t //; 0 .t /i > 0:
dt 2
Therefore, the function f ı is strictly convex on Œ0; 1, and (12.1) follows. t
u
Since " was arbitrary, this shows that dg .q; p0 / c0 , and thus S Bxc0 .p0 /.
Now we need to show that Bxc0 .p0 / is the unique closed ball of radius c0 contain-
ing S . Suppose to the contrary that Bxc0 .p00 / is another such
ball. Let W Œ0; 1 ! M
be the geodesic segment from p0 to p00 , and let m D 12 be its midpoint. Let q 2 S
be arbitrary, and let f W M ! Œ0; 1/ be the function f .x/ D 12 dg .q; x/2 . Lemma
12.15 implies that
1
d .q; m/2 D f .m/ < 12 f .p0 / C 12 f .p00 / 12 12 c02 C 12 12 c02 D 12 c02 :
2 g
Thus for every q 2 S , we have dg .q; m/ < c0 . Since S is compact, the contin-
uous function dg .; m/ takes on a maximum value b0 < c0 on S , showing that
S Bxb0 .m/. This contradicts the fact that c0 is the minimum radius of a ball con-
taining S . t
u
Let us call the center of the smallest enclosing ball the 1-center of the set S .
(The term is borrowed from optimization theory, where the “1-center problem” is
the problem of finding a location for a single production facility that minimizes
356 12 Curvature and Topology
the maximum distance to any client. Some Riemannian geometry texts refer to the
1-center of a set S M as its “circumcenter” or “center of gravity,” but these terms
seem inappropriate, because the 1-center does not coincide with the classical mean-
ing of either term for finite subsets of the Euclidean plane.)
Preissman’s Theorem
Taken together, the Cartan–Hadamard theorem and Cartan’s torsion theorem put
severe restrictions on the possible topologies of manifolds that can carry nonposi-
tively curved metrics. If we make the stronger assumption of strictly negative sec-
tional curvature, we can restrict the topology even further. The following theorem
was first proved in 1943 by Alexandre Preissman [Pre43].
Theorem 12.19 (Preissman). If .M; g/ is a compact, connected Riemannian man-
ifold with strictly negative sectional curvature, then every nontrivial abelian sub-
group of 1 .M / is isomorphic to Z.
Before proving Preissman’s theorem, we need two more lemmas. Suppose .M; g/
is a complete Riemannian manifold and ' W M ! M is an isometry. A geodesic
W R ! M is called an axis for ' if ' restricts to a nontrivial translation along ,
that is, if there is a nonzero constant c such that '..t // D .t C c/ for all t 2 R.
An isometry with no fixed points that has an axis is said to be axial.
Example 12.20 (Axial Isometries). In R2 with its Euclidean metric, every nontrivial
translation .x; y/ 7! .x C a; y C b/ is axial, and every line parallel to the vector
a@x C b@y is an axis. In the upper half-plane model of the hyperbolic plane, every
dilation .x; y/ 7! .cx; cy/ for c ¤ 1 is axial, and the geodesic .t / D .0; e t / is an
axis. On the other hand, the horizontal translations .x; y/ 7! .x C a; y/ are not axial
for the hyperbolic plane, because no geodesic is invariant under such an isometry. //
As the next lemma shows, there is a close relation between axes and closed
geodesics. This lemma will also prove useful in the proof of Synge’s theorem later
in the chapter.
Lemma 12.21. Suppose .M; g/ is a compact, connected Riemannian manifold, and
WM ! M is its universal covering manifold endowed with the metric gz D g.
Then every covering automorphism of has an axis, which restricts to a lift of a
closed geodesic in M that is the shortest admissible path in its free homotopy class.
Proof.
Let ' be any nontrivial covering automorphism of , so it is also an isometry
of M ; gz . Every continuous path z W Œ0; 1 ! M from a point xz 2 M to its image
' xz projects to a loop D ı z in M . We begin by showing that the set of all
loops obtained from ' in this way is exactly one free homotopy class. Suppose
0 ; 1 W Œ0; 1 ! M are two loops, so that i D ı zi for each i , where zi is
such
a path in M from xzi to ' xzi (Fig. 12.3). For i D 0; 1, let xi D xzi , so i is a
at xi .Choose
loop based z from xz0 to xz1 , and let ž D ' ı ˛
a path ˛ z , which is a path
from ' xz0 to ' xz1 , and ˛ D ı ˛ z D ı ž, a path from x0 to x1 in M. Because
M is simply connected, the two paths ˛ z1 z0 ž and z1 from xz1 to ' xz1 are path-
homotopic, and therefore so are their images ˛ 1 0 ˛ and 1 . An easy argument
shows that the loops ˛ 1 0 ˛ and 0 are freely homotopic, by a homotopy that
gradually shrinks the “tail” ˛ to a point, so it follows that 0 and 1 are freely
homotopic.
358 12 Curvature and Topology
d z.1/ z00 .0/ D d z.1/ ı d'z.0/ z0 .0/ (definition of z0 /
D d z.0/ z0 .0/ (because ı ' D /
0
D .0/ (because ı z D /
D 0 .1/ . is a closed geodesic)
D d z.1/ z0 .1/ . ı z D again)
D d z.1/ z10 .0/ (definition of z1 /:
the unit sphere in TC M with its round metric, so we can apply the triangle inequality
for distances on the sphere. Similarly, †ABD †ABC C †CBD. Thus the sum
of the four angles of ABDC is strictly less than 2, which is a contradiction.
It follows that 1 and 2 must intersect at some point, say x D 1 .t1 / D 2 .t2 /.
But then '.x/ D 1 .t1 C1/ D 2 .t2 C1/ is another intersection point, so Proposition
12.9(c) shows that 2 must be a reparametrization of 1 . t
u
Proof of Preissman’s Theorem. Suppose .M; g/ satisfies the hypotheses of the the-
orem,
and let W M ! M be its universal covering, with the metric gz D g. Then
M ; gz is a Cartan–Hadamard manifold with strictly negative curvature. Because
the fundamental group of M based at any point is isomorphic to the group of cov-
ering automorphisms of (see Prop. C.22), it suffices to show that every nontrivial
abelian group of covering automorphisms is isomorphic to Z. Let H be such a
group.
Let ' be any non-identity element of H , and let W R ! M be the axis for ',
which we may give a unit-speed parametrization. Thus there is a nonzero constant c
such that '..t // D .t C c/ for all t 2 R. If is any other element of H , then for
all t 2 R we have
which shows that ı is also an axis for '. Since axes are unique by Lemma
12.22, this means that ı is a (unit-speed) reparametrization of . It cannot be a
backward reparametrization (one satisfying ' ı .t / D .t C a/), because then
would be a nontrivial covering automorphism having .a=2/ as a fixed point, while
the only covering automorphism with a fixed point is the identity. Thus there must
be some a 2 R such that ..t // D .t C a/ for all t 2 R.
Define a map F W H ! R as follows: for each 2 H , let F . / be the unique
constant a such that ..t // D .t C a/ for all t 2 R. A simple computation shows
that F . 1 ı 2 / D F . 1 / C F . 2 /, so F is a group homomorphism. It is injective
because F . / D 0 implies that fixes every point in the image of , so it must be
the identity. Thus F .H / is an additive subgroup of R isomorphic to H .
Now let b be the infimum of the set of positive elements of F .H /. If b D 0, then
there exists 2 H such that 0 < F . / < inj.M /, and then ı is a unit-speed
geodesic in M satisfying ı .a/ D ı .0/, where a D F . /, which contradicts
the fact that 0 < a < inj.M /. Thus b > 0. If b … F .H /, there is a sequence of
elements i 2 H such that F . i / & b, and then for sufficiently large i < j we
have 0 < F . i ı j1 / < inj.M /, which leads to a contradiction as before. Thus b
is the smallest positive element of F .H /, and then it is easy to verify that F .H /
consists exactly of all integral multiples of b, so it is isomorphic to Z. t
u
The most important consequence of Preissman’s theorem is the following corol-
lary.
Corollary 12.23. No product of positive-dimensional connected compact manifolds
admits a metric of strictly negative sectional curvature.
Proof. Problem 12-8. t
u
Manifolds of Positive Curvature 361
Finally, we consider manifolds with positive curvature. The most important fact
about such manifolds is the following theorem, which was first proved in 1941 by
Sumner B. Myers [Mye41], building on earlier work of Ossian Bonnet, Heinz Hopf,
and John L. Synge.
Theorem 12.24 (Myers). Let .M; g/ be a complete, connected Riemannian n-
manifold, and suppose there is a positive constant R such that the Ricci curvature of
M satisfies Rc.v; v/ .n 1/=R2 for all unit vectors v. Then M is compact, with
diameter less than or equal to R, and its fundamental group is finite.
Proof. The diameter comparison theorem (Cor. 11.18) shows that the diameter of
M is no greater than R. To show that M is compact, we just choose a base point
p and note that every point in M can be connected to p by a geodesic segment
of length at most R. Therefore, expp W BxR .0/ ! M is surjective, so M is the
continuous image of a compact set.
To prove the statement about the fundamental group, let W M ! M denote the
universal covering space of M , with the metric gz D g. Then M ; gz is complete
by Corollary 6.24, and gz also has Ricci curvature satisfying the same lower bound,
so M is compact by the argument above. By Proposition A.61, for every p 2 M
there is a one-to-one correspondence between 1 .M; p/ and the fiber 1 .p/.
If 1 .M; p/ were infinite, therefore, 1 .p/ would be an infinite discrete closed
, contradicting the compactness of M
subset of M . Thus 1 .M; p/ is finite. t
u
In case the manifold is already known to be compact, the hypothesis can be
relaxed.
Corollary 12.25. Suppose .M; g/ is a compact, connected Riemannian n-manifold
whose Ricci tensor is positive definite everywhere. Then M has finite fundamental
group.
Proof. Because the unit tangent bundle of M is compact by Proposition 2.9, the
hypothesis implies that there is a positive constant c such that Rc.v; v/ c for all
unit tangent vectors v. Thus the hypotheses of Myers’s theorem are satisfied with
.n 1/=R2 D c. t
u
On the other hand, it is possible for complete, noncompact manifolds to have
strictly positive Ricci or even sectional curvature, as long as the curvature gets arbi-
trarily close to zero, as the following example shows.
I Exercise 12.26. Let n 2, and let M RnC1 be the paraboloid f.x 1 ; : : : ; x n ; y/ W
y D jxj2 g with the induced metric (see Problem 8-1). Show that M has strictly positive
sectional and Ricci curvatures everywhere, but is not compact.
Now suppose ı1 ; ı2 are positive numbers such that ı1 C ı2 D R. The fact that
dg .p1 ; p2 / D R together with the triangle inequality implies that Bı1 .p1 / and
Bı2 .p2 / are disjoint, so
exist a finite generating set S and positive real numbers C; k such that for every
positive integer m, the number of distinct elements of that can be expressed as
products of at most m elements of S and their inverses is no larger than C mk . The
order of polynomial growth of such a group is the infimum of such values of k. It
is easy to check that Zn has polynomial growth of order n; in fact, using the funda-
mental theorem on finitely generated abelian groups [DF04, p. 158], one can show
that every finitely generated abelian group has polynomial growth of order equal to
its free rank.
The following theorem, proved in 1968 by John Milnor [Mil68], shows in effect
that fundamental groups of complete manifolds with nonnegative Ricci curva-
ture are not too far from being abelian, at least when they are finitely generated.
(Milnor conjectured in the same paper that the fundamental group of such a mani-
fold is always finitely generated. This is true in the compact case, because compact
manifolds are homotopy equivalent to finite CW complexes [Hat02, Cor. A.12],
and these have finitely generated fundamental groups, as shown, for example, in
[LeeTM, Thm. 10.15]. But the conjecture has not been proved in the noncompact
case.) By contrast, Milnor proved in the same paper that the fundamental group of
a negatively curved compact manifold never has polynomial growth, and that was
extended in 1970 by André Avez [Ave70] to the case of a nonpositively curved
compact manifold as long as it is not flat.
The corollary implies, for example, that a plane with two or more punctures does
not admit a complete metric with nonnegative Gaussian curvature.
On the other hand, if we strengthen the hypothesis and assume positive sectional
curvature, then we can reach a stronger conclusion. The following theorem was
proved in 1936 by John L. Synge [Syn36].
Proof. Suppose for the sake of contradiction that the appropriate hypothesis holds
! M be the universal covering manifold of
but the conclusion is false. Let W M
M with the pullback metric gz D g. We can give M the pullback orientation in the
Manifolds of Positive Curvature 365
Pz-
Txz0 M Txz1 M
d xz0 d xz1
? P- ?
Tx M Tx M:
Because V is parallel along and sec.V; 0 / > 0 for all t , this second derivative is
strictly negative, which implies that Lg .s / < Lg . / for sufficiently small s. But
this contradicts the fact that minimizes length in its free homotopy class. t
u
The next corollary shows that there are only two possibilities for fundamental
groups of positively curved compact manifolds in the even-dimensional case. (Com-
pare this to Problem 12-2, which shows that the only even-dimensional manifolds
that admit complete metrics of constant positive curvature are spheres and real pro-
jective spaces.)
Further Results
We end the book with a brief look at some other local-to-global theorems about man-
ifolds with positive or nonnegative curvature, whose proofs are beyond our scope.
Some of the most powerful applications of comparison theory have been to prove
“pinching theorems.” Given a positive real number ı, a Riemannian manifold is said
to be ı-pinched if there exists a positive constant c such that all sectional curvatures
satisfy
Manifolds of Positive Curvature 367
ıc sec.˘ / c:
It is said to be strictly ı-pinched if at least one of the inequalities is strict. The
following celebrated theorem was originally proved by Marcel Berger and Wilhelm
Klingenberg in the early 1960s [Ber60, Kli61].
Theorem 12.34 (The Sphere Theorem). A complete, simply connected, Riemann-
ian n-manifold that is strictly 14 -pinched is homeomorphic to Sn .
This result is sharp, at least in even dimensions, because the Fubini–Study met-
rics on complex projective spaces are 14 -pinched (Problem 8-13).
For noncompact manifolds, there is the following remarkable theorem of Jeff
Cheeger and Detlef Gromoll [CG72].
Theorem 12.35 (The Soul Theorem). If .M; g/ is a complete, connected, noncom-
pact Riemannian manifold with nonnegative sectional curvature, then there exists a
compact totally geodesic submanifold S M (called a soul of M ) such that M is
diffeomorphic to the normal bundle of S in M . If M has strictly positive sectional
curvature, then S is a point and M is diffeomorphic to a Euclidean space.
Even if we assume only nonnegative Ricci curvature, something nontrivial can be
said in the noncompact case. The next theorem is also due to Cheeger and Gromoll
[CG71]. Recall that a line in a Riemannian manifold is the image of a nonconstant
geodesic defined on R whose restriction to each compact subinterval is minimizing.
Theorem 12.36 (The Splitting Theorem). If .M; g/ is a complete, connected, non-
compact Riemannian manifold with nonnegative Ricci curvature, and M contains a
line, then there is a Riemannian manifold .N; h/ with nonnegative Ricci curvature
such that M is isometric to the Riemannian product N R.
The proofs of all three of the above theorems, which can be found, for example,
in [Pet16], are elaborate applications of comparison theory.
All of our comparison theorems, and indeed most of the things we have proved
about Riemannian manifolds, are based on the analysis of ordinary differential
equations—the geodesic equation, the parallel transport equation, the Jacobi equa-
tion, and the Riccati equation. Using techniques of partial differential equations can
lead to much stronger conclusions in some cases. For instance, in 1982, Richard
Hamilton [Ham82] proved the following striking result about 3-manifolds.
Theorem 12.37 (Hamilton’s 3-Manifold Theorem). Suppose M is a simply con-
nected compact Riemannian 3-manifold with positive definite Ricci curvature. Then
M is diffeomorphic to S3 .
He followed it up four years later [Ham86] with an analogous result about 4-
manifolds, with Ricci curvature replaced by the curvature operator described in
Problem 8-33.
Theorem 12.38 (Hamilton’s 4-Manifold Theorem). Suppose M is a simply con-
nected compact Riemannian 4-manifold with positive curvature operator. Then M
is diffeomorphic to S4 .
368 12 Curvature and Topology
Much more recently, Christoph Böhm and Burkhard Wilking [BW08] extended
this result to all dimensions.
Theorem 12.39 (Böhm–Wilking). Every simply connected compact Riemannian
manifold with positive curvature operator is diffeomorphic to a sphere.
The method of proof of these three sphere theorems was to start with an initial
metric g0 , and then define a one-parameter family of metrics fgt W t 0g evolving
according to the following partial differential equation, called the Ricci flow:
@
gt D 2Rct ;
@t
where Rct is the Ricci curvature of gt . Under the appropriate curvature assumption
on g0 , the metric gt can be rescaled to converge to a metric with constant positive
sectional curvature as t ! 1, and then it follows from the Killing–Hopf theorem
that the original manifold must be diffeomorphic to a sphere. Since Hamilton first
introduced it, this technique has been vastly generalized by Hamilton and others,
culminating in 2003 in a proof by Grigory Perelman of the Thurston geometrization
conjecture described in Chapter 3.
The Ricci flow has also been used to improve the original 14 -pinched sphere theo-
rem significantly. The techniques used to prove the original theorem were not strong
enough to prove that M is diffeomorphic to Sn , leaving open the possibility that
there might exist strictly 14 -pinched metrics on exotic spheres (topological spheres
with nonstandard smooth structures). In 2009, Simon Brendle and Richard Schoen
[BS09] were able to use Ricci-flow techniques to close this gap. They also were
able to prove the theorem under a weaker hypothesis: a Riemannian metric g is said
to be pointwise ı-pinched if for every p 2 M , there exists a positive number c.p/
such that
ıc.p/ sec.˘ / c.p/
for every 2-plane ˘ Tp M , and strictly pointwise ı-pinched if at least one of the
inequalities is strict at each point.
Theorem 12.40 (Differentiable Sphere Theorem). Let .M; g/ be a compact, sim-
ply connected Riemannian manifold of dimension n 4. If M is strictly pointwise
1
4
-pinched, then it is diffeomorphic to Sn with its standard smooth structure.
For a nice exposition of the proof, see the recent book [Bre10].
Problems
12-1. Suppose .M; g/ is a simply connected (but not necessarily complete) Rie-
mannian n-manifold with constant sectional curvature. Prove that there ex-
ists an isometric immersion from M into one of the model spaces Rn ,
Sn .R/, Hn .R/.
Problems 369
12-2. Prove that if n is even, then every orientation-preserving orthogonal linear map
from RnC1 to itself fixes at least one point of the unit sphere, and use this
to prove that every even-dimensional spherical space form is isometric to
either a round sphere or a real projective space with a constant multiple of
the metric defined in Example 2.34. (Used on p. 349.)
12-3. Show that every compact 2-dimensional Euclidean space form is diffeomor-
phic to the torus or the Klein bottle, and every noncompact one is diffeo-
morphic to R2 , S1 R2 , or the open Möbius band (see Example 2.35).
12-4. A lattice in Rn is an additive subgroup
Rn of the form
D fm1 v1 C C mn vn W m1 ; : : : ; mn 2 Zg;
This book is designed for readers who already have a solid understanding of
basic topology and smooth manifold theory, at roughly the level of [LeeTM] and
[LeeSM]. In this appendix, we summarize the main ideas of these subjects that
are used throughout the book. It is included here as a review, and to establish our
notation and conventions.
Topological Preliminaries
An n-dimensional topological manifold (or simply an n-manifold) is a second-
countable Hausdorff topological space that is locally Euclidean of dimension n,
meaning that every point has a neighborhood homeomorphic to an open subset of
Rn . Given an n-manifold M , a coordinate chart for M is a pair .U; '/, where
U M is an open set and ' W U ! Uy is a homeomorphism from U to an open
subset Uy Rn . If p 2 M and .U; '/ is a chart such that p 2 U , we say that .U; '/
is a chart containing p.
We also occasionally need to consider manifolds with boundary. An n-dimen-
sional topological manifold with boundary is a second-countable Hausdorff space
in which every point has a neighborhood homeomorphic either to an open subset
of Rn or to a (relatively) open subset of the half-space RnC D f.x 1 ; : : : ; x n / 2 Rn W
x n 0g. A pair .U; '/, where U is an open subset of M and ' is a homeomorphism
from U to an open subset of Rn or RnC , is called an interior chart if '.U / is an
open subset of Rn or an open subset of RnC that does not meet @RnC D fx 2 RnC W
x n D 0g; and it is called a boundary chart if '.U / is an open subset of RnC with
'.U / \ @RnC ¤ ¿. A point p 2 M is called an interior point of M if it is in the
domain of an interior chart, and it is a boundary point of M if it is in the domain
of a boundary chart taking p to a point of @RnC .
Notice that our convention is that a manifold without further qualification is al-
ways assumed to be a manifold without boundary, and the term “manifold with
boundary” encompasses ordinary manifolds as a special case. But for clarity, we
© Springer International Publishing AG 2018 371
J. M. Lee, Introduction to Riemannian Manifolds, Graduate Texts
in Mathematics 176, https://doi.org/10.1007/978-3-319-91755-9
372 Review of Smooth Manifolds
sometimes use redundant phrases such as “manifold without boundary” for an ordi-
nary manifold, and “manifold with or without boundary” when we wish to empha-
size that the discussion applies equally well in either case.
Proposition A.1. [LeeTM, Props. 4.23 & 4.64] Every topological manifold with or
without boundary is locally path-connected and locally compact.
Proposition A.2. [LeeTM, Thm. 4.77] Every topological manifold with or without
boundary is paracompact.
If M and N are topological spaces, a map F W M ! N is said to be a closed
map if for each closed subset K M , the image set F .K/ is closed in N , and an
open map if for each open subset U M , the image set F .U / is open in N . It is a
quotient map if it is surjective and V N is open if and only if F 1 .V / is open.
Proposition A.3. [LeeTM, Prop. 3.69] Suppose M and N are topological spaces,
and F W M ! N is a continuous map that is either open or closed.
(a) If F is surjective, it is a quotient map.
(b) If F is injective, it is a topological embedding.
(c) If F is bijective, it is a homeomorphism.
The next lemma often gives a convenient shortcut for proving that a map is
closed.
Lemma A.4 (Closed Map Lemma). [LeeTM, Lemma 4.50] If F is a continuous
map from a compact space to a Hausdorff space, then F is a closed map.
Here is another condition that is frequently useful for showing that a map is
closed. A map F W M ! N between topological spaces is called a proper map if
for each compact subset K N , the preimage F 1 .K/ is compact.
Proposition A.5 (Proper Continuous Maps Are Closed). [LeeTM, Thm. 4.95]
Suppose M is a topological space, N is a locally compact Hausdorff space, and
F W M ! N is continuous and proper. Then F is a closed map.
Fundamental Groups
If there exists such a path homotopy, then ˛0 and ˛1 are said to be path-homotopic;
this is an equivalence relation on the set of all paths in M from p to q. The
equivalence class of a path ˛, called its path class, is denoted by Œ˛.
Paths ˛ and ˇ that satisfy ˛.1/ D ˇ.0/ are said to be composable; in this case,
their product is the path ˛ ˇ W Œ0; 1 ! M defined by
˛.2s/; 0 s 12 ;
˛ ˇ.s/ D
ˇ.2s 1/; 12 s 1:
Path products are well defined on path classes: if Œ˛0 D Œ˛1 and Œˇ0 D Œˇ1 , and
˛0 and ˇ0 are composable, then so are ˛1 and ˇ1 , and Œ˛0 ˇ0 D Œ˛1 ˇ1 . Thus we
obtain a well-defined product of path classes by Œ˛ Œˇ D Œ˛ ˇ.
A path from a point p to itself is called a loop based at p. If M is a topological
space and p 2 M , then any two loops based at p are composable, and the set of
path classes of loops based at p is a group under path class products, called the
fundamental group of M based at p and denoted by 1 .M; p/. The class of the
constant path cp .s/ p is the identity element, and the class of the reverse path
˛ 1 .s/ D ˛.1 s/ is the inverse of Œ˛. (Although multiplication of path classes is
associative, path products themselves are not, so a multiple product such as ˛ ˇ
is to be interpreted as .˛ ˇ/ .)
Proposition A.6 (Induced Homomorphisms). [LeeTM, Prop. 7.24 & Cor. 7.26]
Suppose M and N are topological spaces and p 2 M . If F W M ! N is a contin-
uous map, then the map F W 1 .M; p/ ! 1 .N; F .p// defined by F .Œ˛/ D ŒF ı˛
is a group homomorphism called the homomorphism induced by F . If F is a homeo-
morphism, then F is an isomorphism.
A topological space M is said to be simply connected if it is path-connected and
for some (hence every) point p 2 M , the fundamental group 1 .M; p/ is trivial.
I Exercise A.7. Show that if M is a simply connected topological space and p; q are
any two points in M , then all paths in M from p to q are path-homotopic.
The setting for the study of Riemannian geometry is provided by smooth manifolds,
which are topological manifolds endowed with an extra structure that allows us to
differentiate functions and maps. First, we note that when U is an open subset of Rn ,
a map F W U ! Rk is said to be smooth (or of class C 1 ) if all of its component
functions have continuous partial derivatives of all orders. More generally, if the
domain U is an arbitrary subset of Rn , not necessarily open (such as a relatively
open subset of RnC ), then F is said to be smooth if for each x 2 U , F has a smooth
extension to a neighborhood of x in Rn . A diffeomorphism is a bijective smooth
map whose inverse is also smooth.
If M is a topological n-manifold with or without boundary, then two coordinate
charts .U; '/, .V; / for M are said to be smoothly compatible if both of the transition
maps ı ' 1 and ' ı 1 are smooth where they are defined (on '.U \ V / and
.U \ V /, respectively). Since these maps are inverses of each other, it follows that
both transition maps are in fact diffeomorphisms. An atlas for M is a collection
of coordinate charts whose domains cover M . It is called a smooth atlas if any
two charts in the atlas are smoothly compatible. A smooth structure on M is a
smooth atlas that is maximal, meaning that it is not properly contained in any larger
smooth atlas. A smooth manifold is a topological manifold endowed with a specific
smooth structure; and similarly, a smooth manifold with boundary is a topological
manifold with boundary endowed with a smooth structure. (We usually just say “M is
a smooth manifold,” or “M is a smooth manifold with boundary,” with the smooth
structure understood from the context.) If M is a set, a smooth manifold structure
on M is a second countable, Hausdorff, locally Euclidean topology together with a
smooth structure making it into a smooth manifold.
I Exercise A.9. Let M be a topological manifold with or without boundary. Show that
every smooth atlas for M is contained in a unique maximal smooth atlas, and two smooth
atlases determine the same smooth structure if and only if their union is a smooth atlas.
The manifolds in this book are always assumed to be smooth. As in most parts
of differential geometry, the theory still works under weaker differentiability as-
sumptions (such as k times continuously differentiable), but such considerations are
usually relevant only when one is treating questions of hard analysis that are beyond
our scope.
If M is a smooth manifold with or without boundary, then every coordinate chart
in the given maximal atlas is called a smooth coordinate chart for M or just a smooth
chart. The set U is called a smooth coordinate domain; if its image is an open ball
in Rn , it is called a smooth coordinate ball. If in addition ' extends
to a smooth
coordinate map ' 0 W U 0 ! Rn on an open set U 0 Ux such that ' 0 Ux is the closure
of the open ball '.U / in Rn , then U is called a regular coordinate ball. In this
case, Ux is diffeomorphic to a closed ball and is thus compact.
Proposition A.10. [LeeSM, Prop. 1.19] Every smooth manifold has a countable ba-
sis of regular coordinate balls.
Smooth Manifolds and Smooth Maps 375
Given a smooth chart .U; '/ for M , the component of' are called local
functions
coordinates for M , and are typically written as x 1 ; : : : ; x n , x i , or x, depending
on context. Although, formally speaking, a coordinate chart is a map from an open
subset U M to Rn , it is common when one is working in the domain of a specific chart
innR , and to identify a point
n
to use the coordinate map to identify U with its image
in U with its coordinate representation x ; : : : ; x 2 R .
1 n
In this book, we always write coordinates with upper indices, as in x i , and
expressions with indices are interpreted using the Einstein summation convention:
if in some monomial term the same index name appears exactly twice, once as an
upper index and once as a lower index, then that term is understood to be summed
over all possible values of that index (usually P from 1 to the dimension of the space).
Thus the expression ai vi is to be read as i ai vi . As we will see below, index posi-
tions for other sorts of objects such as vectors and covectors are chosen whenever
possible in such a way that summations that make mathematical sense obey the rule
that each repeated index appears once up and once down in each term to be summed.
Because of the result of Exercise A.9, to define a smooth structure on a manifold,
it suffices to exhibit a single smooth atlas for it. For example, Rn is a topological
n-manifold, and it has a smooth atlas consisting of the single chart Rn ; IdRn . More
generally, if V is an n-dimensional vector space, then every basis .b1 ; : : : ; bn / for
V yields a linear basis isomorphism B W Rn ! V by B.x 1 ; : : : ; x n / D x i bi , whose
inverse is a global coordinate chart, and it is easy to check that all such charts are
smoothly compatible. Thus every finite-dimensional vector space has a natural smooth
structure, which we call its standard smooth structure. We always consider Rn or
any other finite-dimensional vector space to be a smooth manifold with this structure
unless otherwise specified.
If M is a smooth n-manifold with or without boundary and W M is an open
subset, then W has a natural smooth structure consisting of all smooth charts .U; '/
for M such that U W , so every open subset of a smooth n-manifold is a smooth n-
manifold in a natural way, and similarly for manifolds with boundary. If M1 ; : : : ; Mk
are smooth manifolds of dimensions n1 ; : : : ; nk , respectively, then their Cartesian
product M1 Mk has a natural smooth atlas consisting of charts of the form
.U1 Uk ; '1 'k /, where .Ui ; 'i / is a smooth chart for Mi ; thus the prod-
uct space is a smooth manifold of dimension n1 C C nk .
Suppose M and N are smooth manifolds with or without boundary. A map
F W M ! N is said to be smooth if for every p 2 M , there exist smooth charts
.U; '/ for M containing p and .V; / for N containing F .p/ such that F .U / V
and the composite map ı F ı ' 1 is smooth from '.U / to .V /. In particular, if
N is an open subset of Rk or RkC with its standard smooth structure, we can take
to be the identity map of N , and then smoothness of F just means that each point
of M is contained in the domain of a chart .U; '/ such that F ı ' 1 is smooth.
It is an easy consequence of the definition that identity maps, constant maps, and
compositions of smooth maps are all smooth. A map F W M ! N is said to be a
diffeomorphism if it is smooth and bijective and F 1 W N ! M is also smooth.
If F W M ! N is smooth, and .U; '/ and .V; / are any smooth charts for M
and N respectively, the map Fy D ı F ı ' 1 is a smooth map between appropriate
376 Review of Smooth Manifolds
Tangent Vectors
Let M be a smooth manifold with or without boundary. There are various equiv-
alent ways of defining tangent vectors on M . The following definition is the most
convenient to work with in practice. For every point p 2 M , a tangent vector at p
is a linear map v W C 1 .M / ! R that is a derivation at p, meaning that for all
f; g 2 C 1 .M / it satisfies the product rule
The set of all tangent vectors at p is denoted by Tp M and called the tangent space
at p.
Suppose M is n-dimensional and ' W U ! Uy Rn is a smooth coordinate
chart
on some open subset U M . Writing the coordinate functions of ' as x 1 ; : : : ; x n ,
Tangent Vectors 377
These vectors form a basis for Tp M , which therefore has dimension n. When there
can be no confusion about which coordinates are meant, we usually abbreviate
@=@x i jp by the notation @i jp . Thus once a smooth coordinate chart has been chosen,
every tangent vector v 2 Tp M can be written uniquely in the form
v D v i @i jp D v 1 @1 jp C C v n @n jp ; (A.3)
For every p 2 M , the dual vector space to Tp M is called the cotangent space
at p. This is the space Tp M D .Tp M / of real-valued linear functionals on Tp M ,
called (tangent) covectors at p. For every f 2 C 1 .M / and p 2 M , there is a
covector dfp 2 Tp M called the differential of f at p, defined by
Submanifolds
The theory of submanifolds is founded on the inverse function theorem and its corol-
laries.
Theorem A.14 (Inverse Function Theorem for Manifolds). [LeeSM, Thm. 4.5]
Suppose M and N are smooth manifolds and F W M ! N is a smooth map. If the
linear map dFp is invertible at some point p 2 M , then there exist connected neigh-
borhoods U0 of p and V0 of F .p/ such that F jU0 W U0 ! V0 is a diffeomorphism.
The most useful consequence of the inverse function theorem is the following. A
smooth map F W M ! N is said to have constant rank if the linear map dFp has
the same rank at every point p 2 M .
Theorem A.15 (Rank Theorem). [LeeSM, Thm. 4.12] Suppose M and N are
smooth manifolds of dimensions m and n, respectively, and F W M ! N is a smooth
map with constant rank r. For each p 2 M there exist smooth charts .U; '/ for M
centered at p and .V; / for N centered at F .p/ such that F .U / V; in which F
has a coordinate representation of the form
Fy x 1 ; : : : ; x r ; x rC1 ; : : : ; x m D x 1 ; : : : ; x r ; 0; : : : ; 0 : (A.5)
Here are the most important types of constant-rank maps. In all of these defini-
tions, M and N are smooth manifolds, and F W M ! N is a smooth map.
F is a submersion if its differential is surjective at each point, or equivalently
if it has constant rank equal to dim N .
F is an immersion if its differential is injective at each point, or equivalently if
it has constant rank equal to dim M .
F is a local diffeomorphism if every point p 2 M has a neighborhood U such
that F jU is a diffeomorphism onto an open subset of N , or equivalently if F is
both a submersion and an immersion.
F is a smooth embedding if it is an injective immersion that is also a topologi-
cal embedding (a homeomorphism onto its image, endowed with the subspace
topology).
Theorem A.16 (Local Embedding Theorem). [LeeSM, Thm. 4.25] Every smooth
immersion is locally an embedding: if F W M ! N is a smooth immersion, then for
every p 2 M , there is a neighborhood U of p in M such that F jU is an embedding.
Theorem A.17 (Local Section Theorem). [LeeSM, Thm. 4.26] Suppose M and
N are smooth manifolds and W M ! N is a smooth map. Then is a smooth
submersion if and only if every point of M is in the image of a smooth local section
of (a map W U ! M defined on some open subset U N , with ı D IdU ).
Theorem A.20 (Global Rank Theorem). [LeeSM, Thm. 4.14] Suppose M and N
are smooth manifolds, and F W M ! N is a smooth map of constant rank.
(a) If F is surjective, then it is a smooth submersion.
(b) If F is injective, then it is a smooth immersion.
(c) If F is bijective, then it is a diffeomorphism.
Proposition A.22 (Slice Coordinates). [LeeSM, Thms. 5.8 & 5.51] Let M be
a smooth m-manifold without boundary and let M M be an embedded n-
dimensional submanifold with or without boundary. Then for each p 2 M there
exist a neighborhood Uz of p in M on
and smooth coordinates x 1 ; : : : ; x m for M
z z
U such that M \ U is a set of one of the following forms:
z fx 2 Uz W x nC1 D D x m D 0g if p … @M;
M \U D
fx 2 Uz W x nC1 D D x m D 0 and x n 0g if p 2 @M:
In such a chart, x 1 ; : : : ; x n form smooth local coordinates for M .
380 Review of Smooth Manifolds
Theorem A.24 (Constant-Rank Level Set Theorem). [LeeSM, Thm. 5.12] Sup-
pose M and N are smooth manifolds, and ˚ W M ! N is a smooth map with con-
stant rank r. Every level set of ˚ is a properly embedded submanifold of codimen-
.
sion r in M
Corollary A.25 (Submersion Level Set Theorem). [LeeSM, Cor. 5.13] Suppose
M and N are smooth manifolds, and ˚ W M ! N is a smooth submersion. Every
, whose codimension is
level set of ˚ is a properly embedded submanifold of M
equal to dim N .
In fact, a map does not have to be a submersion, or even to have constant rank, for
its level sets to be embedded submanifolds. If ˚ W M ! N is a smooth map, a point
p2M is called a regular point of ˚ if the linear map d ˚p W Tp M ! T˚.p/ N is
surjective, and p is called a critical point of ˚ if it is not. A point c 2 N is called a
regular value of ˚ if every point of ˚ 1 .c/ is a regular point of ˚, and a critical
value otherwise. A level set ˚ 1 .c/ is called a regular level set of ˚ if c is a regular
value of ˚.
Corollary A.26 (Regular Level Set Theorem). [LeeSM, Cor. 5.14] Let M and N
be smooth manifolds, and let ˚ W M ! N be a smooth map. Every regular level
whose codimension is equal to
set of ˚ is a properly embedded submanifold of M
dim N .
Tp M D Ker d ˚p :
Vector Bundles
The smooth GL.k; R/-valued maps ˛ˇ of this lemma are called transition func-
tions for E .
If W E ! M is a smooth vector bundle over M , then a section of E is a continuous
map W M ! E such that ı D IdM , or equivalently, .p/ 2 Ep for all p. If
U M is an open set, then a local section of E over U is a continuous map
W U ! E satisfying the analogous equation ı D IdU . A smooth (local or global)
section of E is just a section that is smooth as a map between smooth manifolds. If we
need to speak of “sections” that are not necessarily continuous, we use the following
terminology: a rough (local) section of E is a map W U ! E defined on some open
set U M , not necessarily smooth or even continuous, such that ı D IdU .
A local frame for E is an ordered k-tuple .1 ; : : : ; k / of local sections over an
open set U whose values at each p 2 U constitute a basis for Ep . It is called a global
frame if U D M . If is a (local or global) section of E and .1 ; : : : ; k / is a local
frame for E over U M , then the value of at each p 2 U can be written
Lemma A.33 (Local Frame Criterion for Smoothness). [LeeSM, Prop. 10.22]
Let W E ! M be a smooth vector bundle, and let W M ! E be a rough section.
If .i / is a smooth local frame for E over an open subset U M , then is smooth
on U if and only if its component functions with respect to .i / are smooth.
Lemma A.34 (Local Frame Criterion for Subbundles). [LeeSM, Lemma 10.32]
Let W E ! M be a smooth vector bundle, and suppose that for each p 2 M
we are given a k-dimensional linear subspace Dp Ep . Suppose further that
each p 2 M has a neighborhood U on which there are smooth local sections
1 ; : : : ; k W S
U ! E such that 1 .q/; : : : ; k .q/ form a basis for Dq at each q 2 U .
Then D D p2M Dp E is a smooth subbundle of E.
384 Review of Smooth Manifolds
Lemma A.36. [LeeSM, Prop. 8.19 & Cor. 8.21] Let F W M ! N be a diffeomor-
phism between smooth manifolds with or without boundary. For every X 2 X.M /,
there is a unique vector field F X 2 X.N /, called the pushforward of X , that is
F -related to X . For every f 2 C 1 .N /, it satisfies
.F X /f ı F D X.f ı F /: (A.7)
Lemma A.37. [LeeSM, Prop. 8.15] Let M be a smooth manifold with or without
boundary. A map D W C 1 .M / ! C 1 .M / is a derivation if and only if it is of the
form Df D Xf for some X 2 X.M /.
Proposition A.39 (Naturality of Lie Brackets). [LeeSM, Prop. 8.30 & Cor. 8.31]
Let F W M ! N be a smooth map between manifolds with or without boundary, and
let X1 ; X2 2 X.M / and Y1 ; Y2 2 X.N / be vector fields such that Xi is F -related
to Yi for i D 1; 2. Then ŒX1 ; X2 is F -related to ŒY1 ; Y2 . In particular, if F is a
diffeomorphism, then F ŒX1 ; X2 D ŒF X1 ; F X2 .
I Exercise A.41. Let M be a smooth manifold with or without boundary and let M
M be an embedded submanifold with or without boundary. Show that a vector field X 2
X M is tangent to M if and only if Xf jM D 0 whenever f 2 C 1 M is a function
that vanishes on M .
varying smoothly as the point varies. These integral curves are all encoded into a
global object called a flow, which we now define.
Given a smooth manifold M (without boundary), a flow domain for M is an
open subset D R M with the property that for each p 2 M , the set D .p/ D
ft 2 R W .t; p/ 2 Dg is an open interval containing 0. Given a flow domain D and a
map W D ! M , for each t 2 R we let Mt D fp 2 M W .t; p/ 2 Dg, and we define
maps t W Mt ! M and .p/ W D .p/ ! M by t .p/ D .p/ .t / D .t; p/.
A flow on M is a continuous map W D ! M , where D R M is a flow
domain, that satisfies
0 D IdM ;
t ı s .p/ D tCs .p/ wherever both sides are defined.
Although the fundamental theorem guarantees only that each point lies on an
integral curve that exists for a short time, the next lemma can often be used to prove
that a particular integral curve exists for all time.
Proposition A.45 (Canonical Form for a Vector Field). [LeeSM, Thm. 9.22] Let
X be a smooth vector field on a smooth manifold M , and suppose p 2 M is a point
where Xp ¤ 0. There exist smooth coordinates x i on some neighborhood of p in
which X has the coordinate representation @=@x 1 .
This formula is useless for computation, however, because typically the flow of
a vector field is difficult or impossible to compute explicitly. Fortunately, there is
another expression for the Lie derivative that is much easier to compute.
Corollary A.57 (Lifting Maps from Simply Connected Spaces). [LeeTM, Cor.
11.19] Suppose W M ! M and F W B ! M satisfy the hypotheses of Theorem
A.56, and in addition B is simply connected. Then every continuous map F W B !
M has a lift to M . Given any b 2 B, the lift can be chosen to take b to any point in
the fiber over F .b/.
Proof. By Theorem A.56, the hypothesis implies that the identity map Id W M ! M
eW M ! M
has a lift Id , which in this case is a continuous inverse for . If is a
smooth covering map, then the lift is also smooth. t
u
Of all the constructions in smooth manifold theory, the ones that play the most
fundamental roles in Riemannian geometry are tensors and tensor fields. Most of
the technical machinery of Riemannian geometry is built up using tensors; indeed,
Riemannian metrics themselves are tensor fields. This appendix offers a brief review
of their definitions and properties. For a more detailed exposition of the material
summarized here, see [LeeSM].
Covectors
Let V be an n-dimensional vector space (all of our vector spaces are assumed to be
real). When we work with bases for V , it is usually important to consider ordered
bases, so will assume that each basis comes endowed with a specific ordering. We
use parentheses to denote ordered k-tuples, and braces to denote unordered ones, so
an ordered basis is designated by either .b1 ; : : : ; bn / or .bi /, and the corresponding
unordered basis by fb1 ; : : : ; bn g or fbi g.
The dual space of V , denoted by V , is the space of linear maps from V to R.
Elements of the dual space are called covectors or linear functionals on V . Under
the operations of pointwise addition and multiplication by constants, V is a vector
space.
Suppose .b1 ; : : : ; bn / is an ordered basis for V . For each i D 1; : : : ; n, define a
covector ˇ i 2 V by
© Springer International Publishing AG 2018 391
J. M. Lee, Introduction to Riemannian Manifolds, Graduate Texts
in Mathematics 176, https://doi.org/10.1007/978-3-319-91755-9
392 Review of Tensors
ˇ i .bj / D ıji ;
where ıji is the Kronecker delta symbol, defined by
1 if i D j;
ıji D (B.1)
0 if i ¤ j:
It is a standard exercise to prove that ˇ 1 ; : : : ; ˇ n is a basis for V , called the dual
basis to .bi /. Therefore, V is finite-dimensional, and its dimension is the same as
that of V . Every covector ! 2 V can thus be written in terms of the dual basis as
! D !j ˇ j ; (B.2)
!.v/ D !j v j : (B.3)
(Here and throughout the rest of this appendix we use the Einstein summation con-
vention; see p. 375.)
Every vector v 2 V uniquely determines a linear functional on V , by ! 7! !.v/.
Because we are assuming V to be finite-dimensional, it is straightforward to show that
this correspondence gives a canonical (basis-independent) isomorphism between V
and V (the dual space of V ). Given ! 2 V and v 2 V , we can denote the natural
action of ! on v either by the traditional functional notation !.v/, or by either of the
more symmetric notations h!; vi or hv; !i. The latter notations are meant to emphasize
that it does not matter whether we think of the resulting number as the effect of the linear
functional ! acting on the vector v, or as the effect of v 2 V acting on the covector
!. Note that when applied to a vector and a covector, this pairing makes sense without
any choice of an inner product on V .
š
F W V V ! R:
k copies
We often need to consider tensors of mixed types as well. A mixed tensor of type
(k, l), also called a k-contravariant, l-covariant tensor, is a multilinear map
š ™
F W V V V V ! R:
k copies l copies
T k .V / D fcovariant k-tensors on V gI
T k .V / D fcontravariant k-tensors on V gI
T .k;l/ .V / D fmixed .k; l/-tensors on V g:
The rank of a tensor is the number of arguments (vectors and/or covectors) it takes.
By convention, a 0-tensor is just a real number. (You should be aware that the no-
tation conventions for describing the spaces of covariant, contravariant, and mixed
tensors are not universally agreed upon. Be sure to look closely at each author’s
conventions.)
Tensor Products
There is a natural product, called the tensor product, linking the various ten-
sor spaces over V : if F 2 T .k;l/ .V / and G 2 T .p;q/ .V /, the tensor F ˝ G 2
T .kCp;lCq/ .V / is defined by
It follows that T .k;l/ .V / has dimension nkCl , where n D dim V . Every tensor F 2
T .k;l/ .V / can be written in terms of this basis (using the summation convention) as
i :::i
F D Fj11:::jkl bi1 ˝ ˝ bik ˝ ˇ j1 ˝ ˝ ˇ jl ; (B.5)
where
i :::i
Fj11:::jkl D F ˇ i1 ; : : : ; ˇ ik ; bj1 ; : : : ; bjl :
If the arguments of a mixed tensor F occur in a nonstandard order, then the
horizontal as well as vertical positions of the indices are significant and reflect which
arguments are vectors and which are covectors. For example, if A is a .1; 2/-tensor
whose first argument is a vector, second is a covector, and third is a vector, its basis
expression would be written
A D Ai j k ˇ i ˝ bj ˝ ˇ k ;
where
Ai j k D A bi ; ˇ j ; bk : (B.6)
There are obvious identifications among some of these tensor spaces:
T .0;0/ .V / D T 0 .V / D T 0 .V / D R;
T .1;0/ .V / D T 1 .V / D V;
T .0;1/ .V / D T 1 .V / D V ; (B.7)
T .k;0/ .V / D T k .V /;
T .0;k/ .V / D T k .V /:
T .1;1/ .V / Š End.V /;
where End.V / denotes the space of linear maps from V to itself (also called the
endomorphisms of V ). This is a special case of the following proposition.
I Exercise B.2. Prove Proposition B.1. [Hint: In the special case k D 0, l D 1, consider
the map ˚ W End.V / ! T .1;1/ .V / defined by letting ˚A be the .1; 1/-tensor defined by
˚A.!; v/ D !.Av/. The general case is similar.]
Tensors on a Vector Space 395
We can use the result of Proposition B.1 to define a natural operation called
trace or contraction, which lowers the rank of a tensor by 2. In one special case,
it is easy to describe: the operator tr W T .1;1/ .V / ! R is just the trace of F when
it is regarded as an endomorphism of V , or in other words the sum of the di-
agonal entries of any matrix representation of F . Since the trace of a linear en-
domorphism is basis-independent, this is well defined. More generally, we define
tr W T .kC1;lC1/ .V / ! T .k;l/ .V / by letting .tr F /.! 1 ; : : : ; ! k ; v1 ; : : : ; vl / be the trace
of the .1; 1/-tensor
F ! 1 ; : : : ; ! k ; ; v1 ; : : : ; vl ; 2 T .1;1/ .V /:
In other words, just set the last upper and lower indices equal and sum. Even more
generally, we can contract a given tensor on any pair of indices as long as one is
contravariant and one is covariant. There is no general notation for this operation,
so we just describe it in words each time it arises. For example, we can contract the
tensor A with components given by (B.6) on its first and second indices to obtain a
covariant 1-tensor B whose components are Bk D Ai i k .
I Exercise B.3. Show that the trace on any pair of indices (one upper and one lower) is
a well-defined linear map from T .kC1;lC1/ .V / to T .k;l/ .V /.
Symmetric Tensors
There are two classes of tensors that play particularly important roles in differential
geometry: the symmetric and alternating tensors. Here we discuss the symmetric
ones; we will take up alternating tensors later in this appendix when we discuss
differential forms.
If V is a finite-dimensional vector space, a covariant tensor F 2 T k .V / is said
to be symmetric if its value is unchanged by interchanging any pair of arguments:
F .v1 ; : : : ; vi ; : : : ; vj ; : : : ; vk / D F .v1 ; : : : ; vj ; : : : ; vi ; : : : ; vk /
fined by
1 X
.SymF /.v1 ; : : : ; vk / D F v.1/ ; : : : ; v.k/ ;
kŠ
2Sk
where Sk is the group of all permutations of f1; : : : ; kg, called the symmetric group
on k elements. It is easy to check that SymF is symmetric, and SymF D F if and
only if F is symmetric. Then if F and G are symmetric tensors on V , of ranks k
and l, respectively, their symmetric product is defined to be the .k C l/-tensor F G
(denoted by juxtaposition without an explicit product symbol) given by
F G D Sym.F ˝ G/:
The most important special case is the symmetric product of two 1-tensors, which
can be characterized as follows: if ! and are covectors, then
! D 12 .! ˝ C ˝ !/:
(To prove this, just evaluate both sides on an arbitrary pair of vectors .v; w/, and
use the definition of !.) If ! is any 1-tensor, the notation ! 2 means the symmetric
product !!, which in turn is equal to ! ˝ !.
There are the usual identifications among these bundles that follow from (B.7): for
example, T 1 TM D T .1;0/ TM D TM and T 1 T M D T .0;1/ TM D †1 T M D
T M .
I Exercise B.4. Show that each tensor bundle is a smooth vector bundle over M , with a
local trivialization over every open subset that admits a smooth local frame for TM .
A tensor field on M is a section of some tensor bundle over M (see p. 383 for the
definition of a section of a bundle). A section of T 1 T M D T .0;1/ TM (a covariant
1-tensor field) is also called a covector field. As we do with vector fields, we write
the value of a tensor field F at p 2 M as Fp or F jp . Because covariant tensor
fields are the most common and important tensor fields we work with, we use the
following shorthand notation for the space of all smooth covariant k-tensor fields:
T k .M / D T k T M :
i :::i
where each coefficient Fj11:::jkl is a smooth real-valued function on U .
I Exercise B.5. Suppose F W M ! T .k;l/ TM is a rough .k; l/-tensor field. Show that
F is smooth on an open set U M if and only if whenever ! 1 ; : : : ; ! k are smooth covec-
torfields and X1 ; : : : ; Xl are smooth vector fields defined on U , the real-valued function
F ! 1 ; : : : ; ! k ; X1 ; : : : ; Xl , defined on U by
F ! 1 ; : : : ; ! k ; X1 ; : : : ; Xl .p/ D Fp ! 1 jp ; : : : ; ! k jp ; X1 jp ; : : : ; Xl jp ;
is smooth.
œ
Fz W T 1 .M / T 1 .M / X.M / X.M / ! C 1 .M /:
k factors l factors
It is easy to check that this map is multilinear over C 1 .M /, that is, for all functions
u; v 2 C 1 .M / and smooth vector or covector fields ˛, ˇ,
Fz .: : : ; u˛ C vˇ; : : : / D uFz .: : : ; ˛; : : : / C v Fz .: : : ; ˇ; : : : /:
Even more important is the converse: as the next lemma shows, every such map that
is multilinear over C 1 .M / defines a tensor field. (This lemma is stated and proved
in [LeeSM] for covariant tensor fields, but the same argument works in the case of
mixed tensors.)
œ
F W T 1 .M / T 1 .M / X.M / X.M / ! C 1 .M /
k factors l factors
is induced by a smooth .k; l/-tensor field as above if and only if it is multilinear over
C 1 .M /. Similarly, a map
œ
F W T 1 .M / T 1 .M / X.M / X.M / ! X.M /
k factors l factors
Because of this result, it is common to use the same symbol for both a tensor
field and the multilinear map on sections that it defines, and to refer to either of
these objects as a tensor field.
(a) F .f B/ D .f ı F /F B.
(b) F .A ˝ B/ D F A ˝ F B.
(c) F .A C B/ D F A C F B.
(d) F B is a (continuous) tensor field, and it is smooth if B is smooth.
(e) .F ı G/ B D G .F B/.
(f) .IdN / B D B.
We can extend the notion of Lie derivatives to tensor fields. This can be done for
mixed tensor fields of any rank, but for simplicity we restrict attention to covariant
tensor fields. Suppose X is a smooth vector field on M and is its flow. If A is a
smooth covariant tensor field on M , the Lie derivative of A with respect to X is
the smooth covariant tensor field LX A defined by
ˇ
d ˇˇ
d.t /p At .p/ Ap
.LX A/p D . A/p D lim :
dt ˇtD0 t t!0 t
As we did for vector fields, we say that a covariant tensor field A is invariant
under a flow if .t / A D A wherever it is defined. The next proposition is a
tensor analogue of Proposition A.47.
In addition to symmetric tensors, the other class of tensors that play a special role
in differential geometry is that of alternating tensors, which we now define. Let
V be a finite-dimensional vector space. If F is a covariant k-tensor on V , we say
that F is alternating if its value changes sign whenever two different arguments are
interchanged:
F .v1 ; : : : ; vi ; : : : ; vj ; : : : ; vk / D F .v1 ; : : : ; vj ; : : : ; vi ; : : : ; vk /:
Again, it is easy to check that Alt F is alternating, and Alt F D F if and only if F is
alternating. Given ! 2 ƒk .V / and 2 ƒl .V /, we define their wedge product by
.k C l/Š
! ^ D Alt.! ˝ /:
kŠlŠ
It is immediate that ! ^ is an alternating .k C l/-tensor. The wedge product is
easily seen to be bilinear and anticommutative, which means that
It is also the case, although not so easy to see, that it is associative; see [LeeSM,
Prop. 14.11] for a proof. If .bi / is a basis for V and .ˇ i / is the dual basis, the wedge
products ˇ i1 ^ ^ ˇ ik , as .i1 ; : : : ; ik / range over strictly increasing
multi-indices,
form a basis for ƒk .V /, which therefore has dimension kn D nŠ= kŠ.n k/Š ,
where n D dim V . In terms of these basis elements, the wedge product satisfies
i
ˇ i1 ^ ^ ˇ ik .bj1 ; : : : ; bjk / D det ıjpq :
More generally, if ! 1 ; : : : ; ! k are any covectors and v1 ; : : : ; vk are any vectors, then
Differential Forms and Integration 401
! 1 ^ ^ ! k .v1 ; : : : ; vk / D det ! i .vj / : (B.10)
The wedge product can be defined analogously for contravariant alternating ten-
sors ˛ 2 ƒp .V /, ˇ 2 ƒq .V /, simply by regarding ˛, ˇ, and ˛ ^ ˇ as alternating
multilinear functionals on V .
The convention we use for the wedge product is referred to in [LeeSM] as the
determinant convention. There is another convention that is also in common use,
the Alt convention, which amounts to multiplying the right-hand side of (B.10) by a
factor of 1=kŠ. The choice of which definition to use is a matter of taste, though there
are various reasons to justify each choice depending on the context. The determinant
convention is most common in introductory differential geometry texts, and is used,
for example, in [Boo86, Cha06, dC92, LeeJeff09, LeeSM, Pet16, Spi79, Tu11]. The
Alt convention is used in [KN96] and is more common in complex differential geom-
etry.
Given an alternating k-tensor ! 2 ƒk .V / and a vector v 2 V , we define an
alternating .k 1/-tensor v ³ ! by
Exterior Derivatives
The most important operation on differential forms is the exterior derivative, defined
as follows. Suppose M is a smooth n-manifold with or without boundary, and x i
are any smooth local coordinates on M . A smooth k-form ! can be expressed in
these coordinates as
X
!D !j1 :::jk dx j1 ^ ^ dx jk ;
j1 <<jk
and then we define the exterior derivative of !, denoted by d!, to be the .k C 1/-
form defined in coordinates by
402 Review of Tensors
X X
n
@!j 1 :::jk
d! D dx i ^ dx j1 ^ ^ dx jk :
@x i
j1 <<jk iD1
X
n
@f
df D dx i ;
@x i
iD1
d.! ^ / D d! ^ C .1/k ! ^ d:
Orientations
If V is a finite-dimensional vector space, an orientation of V is an equivalence
class of ordered bases for V , where two ordered bases are considered equivalent
Differential Forms and Integration 403
if the transition matrix that expresses one basis in terms of the other has posi-
tive determinant. Every vector space has exactly two orientations. Once an orien-
tation is chosen, a basis is said to be positively oriented if it belongs to the chosen
orientation, and negatively oriented if not. The standard orientation of Rn is the one
determined by the standard basis .e1 ; : : : ; en /, where ei is the vector .0; : : : ; 1; : : : ; 0/
with 1 in the i th place and zeros elsewhere.
If M is a smooth manifold, an orientation for M is a choice of orientation for
each tangent space that is continuous in the sense that in a neighborhood of every
point there is a (continuous) local frame that determines the given orientation at
each point of the neighborhood. If there exists an orientation for M , we say that
M is orientable. An oriented manifold is a smooth orientable manifold together
with a choice of orientation.
i If M is an oriented n-manifold, then a smooth
coordinate
chart U;
x is said to be an oriented chart if the coordinate frame
@=@x 1 ; : : : ; @=@x n is positively oriented at each point. Exactly the same definitions
apply to manifolds with boundary.
Proof. Corollary A.59 shows that a simply connected manifold does not admit two-
sheeted covering maps. t
u
provided the integral is well defined. This will always be the case, for example, if f
is continuous and has compact support in U .
In general, if ! is a compactly supported n-form on a smooth n-manifold M
with or without boundary, we define the integral of ! over M by choosing finitely
many oriented smooth coordinate charts fUi gkiD1 whose domains cover the support
of !, together with a smooth partition of unity f 1 gkiD1 subordinate to this cover,
and defining
Z X k Z
1
!D 'i . i !/;
M y
iD1 Ui
where Uyi D 'i .Ui /, and the integrals on the right-hand side are defined as above.
Proposition 16.5 in [LeeSM] shows that this definition does not depend on the
choice of oriented charts or partition of unity.
Densities
On an oriented n-manifold with or without boundary, n-forms are the natural objects
to integrate. But in order to integrate on a nonorientable manifold, we need closely
related objects called densities.
If V is an n-dimensional real vector space, a density on V is a function
™
W V V ! R
n copies
Lie groups play many important roles in Riemannian geometry, both as examples of
Riemannian manifolds and as isometry groups of other manifolds. In this appendix,
we summarize the main facts about Lie groups that are used in this book. For details,
consult [LeeSM], especially Chapters 7, 8, 20, and 21.
Proposition C.1. [LeeSM, Thms. 7.5 & 21.27] Suppose F W G ! H is a Lie group
homomorphism.
(a) F has constant rank.
(b) The kernel of F is an embedded Lie subgroup of G.
(c) The image of F is an immersed Lie subgroup of H .
With the Lie bracket structure that it inherits from X.G/, the Lie algebra Lie.G/ is
called the Lie algebra of G .
Proposition C.3. [LeeSM, Thm. 8.37] Let G be a Lie group with identity e. The
evaluation map X 7! Xe is a vector space isomorphism from Lie.G/ to Te G, so
Lie.G/ has the same dimension as G itself.
The following proposition shows that every Lie group homomorphism induces a
Lie algebra homomorphism between the respective Lie algebras.
Proposition C.7 (Properties of the Lie Group Exponential Map). [LeeSM, Props.
20.5 & 20.8] Let G be a Lie group and let g be its Lie algebra.
(a) expG W g ! G is smooth.
(b) For every X 2 g, the map W R ! G defined by .t / D expG .tX / is the one-
parameter subgroup generated
by X .
(c) The differential d expG 0 W T0 g ! Te G is the identity map, under the canon-
ical identifications of T0 g and Te G with g.
The exponential map is the key ingredient in the proof of the following funda-
mental result.
Theorem C.8 (Closed Subgroup Theorem). [LeeSM, Thm. 20.12 & Cor. 20.13]
Suppose G is a Lie group and H G is a subgroup in the algebraic sense. Then H
is an embedded Lie subgroup of G if and only if it is closed in the topological sense.
Proposition C.9. [LeeSM, Prop. 21.28] Every discrete subgroup of a Lie group is a
closed Lie subgroup of dimension zero.
Group Actions on Manifolds 411
Adjoint Representations
Let G be a Lie group and g its Lie algebra. For every ' 2 G, conjugation by
' gives a Lie group automorphism C' W G ! G, called an inner automorphism,
by C' . / D ' ' 1 . Let Ad.'/ D .C' / W g ! g be the induced Lie algebra au-
tomorphism. It follows from the definition that C'1 ı C'2 D C'1 '2 , and therefore
Ad.'1 / ı Ad.'2 / D Ad.'1 '2 /; in other words, Ad W G ! GL.g/ is a representation,
called the adjoint representation of G . Proposition 20.24 in [LeeSM] shows that
Ad is a smooth map.
There is also an adjoint representation for Lie algebras. If g is any Lie algebra, a
representation of g is a Lie algebra homomorphism from g to gl.V /, the Lie algebra
of all linear endomorphisms of some vector space V . For each X 2 g, define a
map ad.X / W g ! g by ad.X /Y D ŒX; Y . This defines ad as a map from g to gl.g/,
and a straightforward computation shows that it is a representation of g, called the
adjoint representation of g.
The next proposition shows how the two adjoint representations are related.
Proposition C.11. [LeeSM, Thm. 20.27] Let G be a Lie group and g its Lie
algebra, and let Ad W G ! GL.g/ and ad W g ! gl.g/ be their respective adjoint
representations. The induced Lie algebra representation Ad W g ! gl.g/ is given
by Ad D ad.
write ' .p/ in place of ' p, with a similar convention for right actions. Since right
actions can be converted to left actions and vice versa by setting ' p D p ' 1 ,
for most purposes we lose no real generality by restricting attention to left actions.
(But there are also situations in which right actions arise naturally.)
Given an action of G on M , for each p 2 M the isotropy subgroup at p is
the subgroup Gp G consisting of all elements that fix p: that is, Gp D f' 2 G W
' p D pg. The group action is said to be free if ' p D p for some ' 2 G and
p 2 M implies ' D e, or in other words, if Gp D feg for every p. It is said to be
effective if '1 p D '2 p for all p if and only if '1 D '2 , or equivalently, if the only
element of G that fixes every element of M is the identity. For effective actions, each
element of G is uniquely determined by the map p 7! ' p. In such cases, we will
sometimes use the same notation ' to denote either the element ' 2 G or the map
p 7! ' p. The action is said to be transitive if for every pair of points p; q 2 M ,
there exists ' 2 G such that ' p D q.
If M is a smooth manifold, an action by a group G on M is said to be an action
by diffeomorphisms if for each ' 2 G, the map p 7! ' p is a diffeomorphism
of M . If G is a Lie group, an action of G on a smooth manifold M is said to be
a smooth action if the defining map G M ! M is smooth. In this case, it is
also an action by diffeomorphisms, because each map p 7! ' p is smooth and has
p 7! ' 1 p as an inverse. If G is any countable group, an action by G on M is an
action by diffeomorphisms if and only if it is a smooth action when G is regarded
as a 0-dimensional Lie group with the discrete topology.
Now suppose G is a Lie group, and M and N are smooth manifolds endowed
with G-actions (on the left, say). A map F W M ! N is said to be equivariant with
respect to the given G actions if F .' x/ D ' F .x/ for all ' 2 G and x 2 M .
Proposition C.15. [LeeSM, Prop. 21.5] Suppose G is a Lie group acting smoothly
on a smooth manifold M . The action is proper if and only if the following condition
is satisfied: if .pi / is a sequence in M and .'i / is a sequence in G such that both
.pi / and .'i pi / converge, then a subsequence of .'i / converges.
Corollary C.16. Every smooth action by a compact Lie group on a smooth manifold
is proper.
Proof. If G is a compact Lie group, then every sequence in G has a convergent
subsequence, so every smooth G-action is proper by the preceding proposition. u
t
The next theorem is the most important application of proper actions. If a group
G acts on a manifold M , then each p 2 M determines a subset G p D f' p W ' 2 Gg,
called the orbit of p. Because two orbits are either identical or disjoint, the orbits
form a partition of G. The set of orbits is denoted by M=G, and with the quotient
topology it is called the orbit space of the action.
Theorem C.17 (Quotient Manifold Theorem). [LeeSM, Thm. 21.10] Suppose G
is a Lie group acting smoothly, freely, and properly on a smooth manifold M . Then
the orbit space M=G is a topological manifold whose dimension is equal to the
difference dim M dim G; and it has a unique smooth structure with the property
that the quotient map W M ! M=G is a smooth submersion.
Example C.18 (Real Projective Spaces). For each nonnegative integer n, the n-
dimensional real projective space, denoted by RP n , is defined as the set of one-
dimensional linear subspaces of RnC1 . It can be identified with the orbit space of
RnC1 Xf0g under the R of nonzero real numbers given by scalar
action of the group
multiplication: x 1 ; : : : ; x nC1 D x 1 ; : : : ; x nC1 . It is easy to check that this
action is smooth and free. To see that it is proper, we use Proposition C.15. Suppose
.xi / is a sequence in RnC1 X f0g and .i / is a sequence in R such that xi ! x 2
RnC1 X f0g and i xi ! y 2 RnC1 X f0g. Then ji j D ji xi j=jxi j converges to the
nonzero real number jyj=jxj. Thus the numbers i all lie in a compact set of the
form f W 1=C jj C g for some positive number C , so a subsequence converges
to a nonzero real number. Therefore, by the quotient manifold theorem, RP n has
a unique structure as a smooth n-dimensional manifold such that the quotient map
RnC1 X f0g ! RP n is a smooth submersion. //
Example C.19 (Complex Projective Spaces). Similarly, the n-dimensional com-
plex projective space, denoted by CP n , is the set of 1-dimensional complex sub-
spaces of C nC1 , identified with the orbit space of C nC1 X f0g under the C -action
given by z D z. The same argument as in the preceding example shows that this
action is smooth, free, and proper, so CP n is a smooth 2n-dimensional manifold
and the quotient map W C nC1 X f0g ! CP n is a smooth submersion. //
There is a close connection between smooth covering maps and smooth group
and M are smooth manifolds and W M
actions. To begin, suppose M !M
414 Review of Lie Groups
Let W M ! M be a smooth covering
Proposition C.20. [LeeSM, Prop. 21.12]
is a discrete Lie group acting smoothly,
map. With the discrete topology, Aut M
.
freely, and properly on M
[AP94] Marco Abate and Giorgio Patrizio, Finsler Metrics, a Global Approach: with Appli-
cations to Geometric Function Theory, Springer, New York, 1994.
[Amb56] Warren Ambrose, Parallel translation of Riemannian curvature, Ann. of Math. (2) 64
(1956), 337–363.
[Ave70] André Avez, Variétés riemanniennes sans points focaux, C. R. Acad. Sci. Paris Sér.
A-B 270 (1970), A188–A191 (French).
[BCS00] David Bao, Shiing-Shen Chern, and Zhongmin Shen, An Introduction to Riemann–
Finsler Geometry, Graduate Texts in Mathematics, vol. 200, Springer, New York,
2000.
[Ber60] M. Berger, Les variétés Riemanniennes (1/4)-pincées, Ann. Scuola Norm. Sup. Pisa
(3) 14 (1960), 161–170 (French).
[Ber03] Marcel Berger, A Panoramic View of Riemannian Geometry, Springer-Verlag, Berlin,
2003.
[Bes87] Arthur L. Besse, Einstein Manifolds, Springer, Berlin, 1987.
[BBBMP] Laurent Bessières, Gérard Besson, Michel Boileau, Sylvain Maillot, and Joan Porti,
Geometrisation of 3-Manifolds, EMS Tracts in Mathematics, vol. 13, European Math-
ematical Society (EMS), Züurich, 2010.
[Bis63] Richard L. Bishop, A relation between volume, mean curvature, and diameter,
Notices Amer. Math. Soc. 10 (1963), 364.
[BC64] Richard L. Bishop and Richard J. Crittenden, Geometry of Manifolds, Pure and
Applied Mathematics, Vol. XV, Academic Press, New York-London, 1964.
[BW08] Christoph Böhm and Burkhard Wilking, Manifolds with positive curvature operators
are space forms, Ann. of Math. (2) 167 (2008), no. 3, 1079–1097.
[Boo86] William M. Boothby, An Introduction to Differentiable Manifolds and Riemannian
Geometry, 2nd ed., Academic Press, Orlando, FL, 1986.
[Bre10] Simon Brendle, Ricci Flow and the Sphere Theorem, Graduate Studies in Mathemat-
ics, vol. 111, American Mathematical Society, Providence, RI, 2010.
[BS09] Simon Brendle and Richard Schoen, Manifolds with 1/4-pinched curvature are space
forms, J. Amer. Math. Soc. 22 (2009), 287–307.
[Car26] Élie Cartan, Sur une classe remarquable d’espaces de Riemann, Bull. Soc. Math.
France 54 (1926), 214–264 (French).
[Cha06] Isaac Chavel, Riemannian Geometry, 2nd ed., Cambridge Studies in Advanced Math-
ematics, vol. 98, Cambridge University Press, Cambridge, 2006. A modern introduc-
tion.
[CE08] Jeff Cheeger and David G. Ebin, Comparison Theorems in Riemannian Geometry,
AMS Chelsea Publishing, Providence, 2008. Revised reprint of the 1975 original.
[CG71] Jeff Cheeger and Detlef Gromoll, The splitting theorem for manifolds of nonnegative
Ricci curvature, J. Differential Geometry 6 (1971), 119–128.
[CG72] Jeff Cheeger and Detlef Gromoll. On the structure of complete manifolds of nonneg-
ative curvature, Ann. Math. (2) 96 (1972), 413–443.
[Che75] Shiu-Yuen Cheng, Eigenvalue comparison theorems and its geometric applications,
Math. Z. 143 (1975), no. 3, 289–297.
[Che55] Shiing-Shen Chern, An elementary proof of the existence of isothermal parameters on
a surface, Proc. Amer. Math. Soc. 6 (1955), 771–782.
[CB09] Yvonne Choquet-Bruhat, General Relativity and the Einstein Equations, Oxford
Mathematical Monographs, Oxford University Press, Oxford, 2009.
[Cla70] Chris J. S. Clarke, On the Global Isometric Embedding of Pseudo-Riemannian Mani-
folds, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 314 (1970), no. 1518, 417–428.
[CM11] Tobias H. Colding and William P. Minicozzi II, A Course in Minimal Surfaces, Grad-
uate Studies in Mathematics, vol. 121, American Mathematical Society, Providence,
RI, 2011.
[dC92] Manfredo Perdigão do Carmo, Riemannian Geometry, Mathematics: Theory & Appli-
cations, Birkhäuser Boston, Inc., Boston, MA, 1992. Translated from the second Por-
tuguese edition by Francis Flaherty.
[DF04] David S. Dummit and Richard M. Foote, Abstract Algebra, 3rd ed., John Wiley &
Sons, Inc., Hoboken, NJ, 2004.
[Fra61] Theodore Frankel, Manifolds with positive curvature, Pacific J. Math. 11 (1961), 165–
174.
[GHL04] Sylvestre Gallot, Dominique Hulin, and Jacques Lafontaine, Riemannian Geometry,
3rd ed., Universitext, Springer-Verlag, Berlin, 2004.
[Gan73] David Gans, An Introduction to Non-Euclidean Geometry, Academic Press, New
York, 1973.
[Gau65] Carl F. Gauss, General Investigations of Curved Surfaces, Raven Press, New York,
1965.
[Gra82] Alfred Gray, Comparison theorems for the volumes of tubes as generalizations of the
Weyl tube formula, Topology 21 (1982), no. 2, 201–228.
[Gra04] Alfred Gray, Tubes, 2nd ed., Progress in Mathematics, vol. 221, Birkhäuser Verlag,
Basel, 2004. With a preface by Vicente Miquel.
[Gre93] Marvin J. Greenberg, Euclidean and Non-Euclidean Geometries: Development and
History, W. H. Freeman, New York, 1993.
[Gre70] Robert E. Greene, Isometric Embeddings of Riemannian and Pseudo-Riemannian
Manifolds, Memoirs of the Amer. Math. Soc., No. 97, Amer. Math. Soc., Providence,
1970.
[Gro07] Misha Gromov, Metric Structures for Riemannian and Non-Riemannian Spaces,
Reprint of the 2001 English edition, Modern Birkhäuser Classics, Birkhäuser Boston,
Inc., Boston, MA, 2007. Based on the 1981 French original; With appendices by M.
Katz, P. Pansu and S. Semmes; Translated from the French by Sean Michael Bates.
[Gün60] Paul Günther, Einige Sätze über das Volumenelement eines Riemannschen Raumes,
Publ. Math. Debrecen 7 (1960), 78–93 (German).
[Ham82] Richard S. Hamilton, Three-manifolds with positive Ricci curvature, J. Differential
Geom. 17 (1982), 255–306.
[Ham86] Richard S. Hamilton, Four-manifolds with positive curvature operator, J. Differential
Geom. 24 (1986), no. 2, 153–179.
[Hat02] Allen Hatcher, Algebraic Topology, Cambridge University Press, Cambridge, 2002.
[HE73] Stephen W. Hawking and George F. R. Ellis, The Large-Scale Structure of Space-
Time, Cambridge University Press, Cambridge, 1973.
[Hel01] Sigurdur Helgason, Differential Geometry, Lie Groups, and Symmetric Spaces, Grad-
uate Studies in Mathematics, vol. 34, American Mathematical Society, Providence,
RI, 2001. Corrected reprint of the 1978 original.
[Her63] Robert Hermann, Homogeneous Riemannian manifolds of non-positive sectional cur-
vature, Nederl. Akad. Wetensch. Proc. Ser. A 66 D Indag. Math. 25 (1963), 47–56.
[Hic59] Noel J. Hicks, A theorem on affine connexions, Illinois J. Math. 3 (1959), 242–254.
REFERENCES 417
[Hil71] David Hilbert, The Foundations of Geometry, Open Court Publishing Co., Chicago,
1971. Translated from the tenth German edition by Leo Unger.
[Hop89] Heinz Hopf, Differential Geometry in the Large, 2nd ed., Lecture Notes in Mathe-
matics, vol. 1000, Springer, 1989. Notes taken by Peter Lax and John W. Gray; with
a preface by S. S. Chern; with a preface by K. Voss.
[Ive92] Birger Iversen, Hyperbolic Geometry, London Mathematical Society Student Texts,
vol. 25, Cambridge University Press, Cambridge, 1992.
[Jos17] Jürgen Jost, Riemannian Geometry and Geometric Analysis, 7th ed., Universitext,
Springer, Cham, 2017.
[JP13] Marek Jarnicki and Peter Pflug, Invariant Distances and Metrics in Complex Analysis,
Second extended edition, De Gruyter Expositions in Mathematics, vol. 9, Walter de
Gruyter GmbH & Co. KG, Berlin, 2013.
[Kaz85] Jerry Kazdan, Prescribing the Curvature of a Riemannian Manifold, CBMS Regional
Conf. Ser. in Math., Amer. Math. Soc., Providence, 1985.
[KL08] Bruce Kleiner and John Lott, Notes on Perelman’s papers, Geom. Topol. 12 (2008),
no. 5, 2587–2855.
[Kli61] Wilhelm P. A. Klingenberg, Über Riemannsche Mannigfaltigkeiten mit positiver
Krümmung, Comment. Math. Helv. 35 (1961), 47–54 (German).
[Kli95] Wilhelm P. A. Klingenberg, Riemannian Geometry, 2nd ed., De Gruyter Studies in
Mathematics, vol. 1, Walter de Gruyter & Co., Berlin, 1995.
[Kob72] Shoshichi Kobayashi, Transformation Groups in Differential Geometry, Springer,
Berlin, 1972.
[KN96] Shoshichi Kobayashi and Katsumi Nomizu, Foundations of Differential Geometry,
Wiley Classics Library, vol. I–II, John Wiley & Sons, Inc., New York, 1996. Reprint
of the 1963 and 1969 originals; A Wiley-Interscience Publication.
[LeeJeff09] Jeffrey M. Lee, Manifolds and Differential Geometry, Graduate Studies in Mathemat-
ics, vol. 107, American Mathematical Society, Providence, RI, 2009.
[LeeAG] John M. Lee, Axiomatic Geometry, Pure and Applied Undergraduate Texts, vol. 21,
American Mathematical Society, Providence, RI, 2013.
[LeeSM] John M. Lee, Introduction to Smooth Manifolds, 2nd ed., Graduate Texts in Mathe-
matics, vol. 218, Springer, New York, 2013.
[LeeTM] John M. Lee, Introduction to Topological Manifolds, 2nd ed., Graduate Texts in Math-
ematics, vol. 202, Springer, New York, 2011.
[LP87] John M. Lee and Thomas H. Parker, The Yamabe problem, Bull. Amer. Math. Soc.
(N.S.) 17 (1987), 37–91.
[MP11] William H. Meeks III and Joaqun Pérez, The classical theory of minimal surfaces,
Bull. Amer. Math. Soc. (N.S.) 48 (2011), 325–407.
[Mil63] John W. Milnor, Morse Theory, based on lecture notes by M. Spivak and R. Wells.
Annals of Mathematics Studies, No. 51, Princeton University Press, Princeton, N.J.,
1963.
[Mil68] John W. Milnor, A note on curvature and fundamental group, J. Differential Geometry
2 (1968), 1–7.
[Mil76] John W. Milnor, Curvatures of left invariant metrics on Lie groups, Advances in Math.
21 (1976), no. 3, 293–329.
[Mor98] Frank Morgan, Riemannian Geometry: A Beginner’s Guide, 2nd ed., A K Peters Ltd.,
Wellesley, MA, 1998.
[MF10] John W. Morgan and Frederick Tsz-Ho Fong, Ricci Flow and Geometrization of 3-
Manifolds, University Lecture Series, vol. 53, American Mathematical Society, Prov-
idence, RI, 2010.
[MT14] John W. Morgan and Gang Tian, The Geometrization Conjecture, Clay Mathematics
Monographs, vol. 5, American Mathematical Society, Providence, RI; Clay Mathe-
matics Institute, Cambridge, MA, 2014.
[Mun56] James Raymond Munkres, Some Applications of Triangulation Theorems, ProQuest
LLC, Ann Arbor, MI, 1956. Thesis (Ph.D.)-University of Michigan.
418 REFERENCES
[Mye41] Sumner B. Myers, Riemannian manifolds with positive mean curvature, Duke Math.
J. 8 (1941), 401–404.
[MS39] Sumner B. Myers and Norman E. Steenrod, The group of isometries of a Riemannian
manifold, Ann. of Math. (2) 40 (1939), no. 2, 400–416.
[Nas56] John Nash, The imbedding problem for Riemannian manifolds, Ann. Math. 63 (1956),
20–63.
[O’N83] Barrett O’Neill, Semi-Riemannian Geometry with Applications to General Relativity,
Academic Press, New York, 1983.
[Pet16] Peter Petersen, Riemannian Geometry, 3rd ed., Graduate Texts in Mathematics, vol.
171, Springer, Cham, 2016.
[Poo81] Walter A. Poor, Differential Geometric Structures, McGraw-Hill, 1981.
[Pre43] Alexandre Preissman, Quelques propriétés globales des espaces de Riemann, Com-
ment. Math. Helv. 15 (1943), 175–216 (French).
[Rad25] Tibor Radó, Über den Begriff der Riemannschen Fläche, Acta Litt. Sci. Szeged. 2
(1925), 101–121 (German).
[Rat06] John G. Ratcliffe, Foundations of Hyperbolic Manifolds, 2nd ed., Graduate Texts in
Mathematics, vol. 149, Springer, New York, 2006.
[Sco83] Peter Scott, The Geometries of 3-Manifolds, Bull. London Math. Soc. 15 (1983), 401–
487.
[Spi79] Michael Spivak, A Comprehensive Introduction to Differential Geometry, Vol. I–V,
Publish or Perish, Berkeley, 1979.
[Str86] Robert S. Strichartz, Sub-Riemannian geometry, J. Differential Geom. 24 (1986), no.
2, 221–263.
[Syn36] John L. Synge, On the connectivity of spaces of positive curvature, Q. J. Math. os-7
(1936), 316–320.
[Thu97] William P. Thurston, Three-Dimensional Geometry and Topology. Vol. 1, Princeton
Mathematical Series, vol. 35, Princeton University Press, Princeton, NJ, 1997. Edited
by Silvio Levy.
[Tu11] Loring W. Tu, An Introduction to Manifolds, 2nd ed., Universitext, Springer, New
York, 2011.
[Wol11] Joseph A. Wolf, Spaces of Constant Curvature, 6th ed., AMS Chelsea Publishing,
Providence, RI, 2011.
Notation Index
Symbols C
^ (wedge product), 400 C (Cotton tensor), 219
(Kulkarni–Nomizu product), 213 C 1 (infinitely differentiable), 374
q (disjoint union), 382 C 1 .M / (smooth real-valued functions),
³ (interior multiplication), 31, 401 376
] (sharp), 26–28 C 1 .M; N / (smooth maps), 376
(Hodge star operator), 50 CP n (complex projective space), 413
j j (length of a vector), 12 conv.x/ (convexity radius at x), 186
j jg (length of a vector), 12 Cut.p/ (cut locus of p), 308
Œ (path class), 373
Œ ; (commutator bracket), 409 D
Œ ; (Lie bracket), 385 ıji (Kronecker delta), 392
. ; / (global inner product on forms), 53 (Laplacian), 32
h ; i (inner product), 9, 12 @=@x i (coordinate vector field), 384
h ; i (pairing between vector and covector), @=@x i jp (coordinate vector), 377
392 @i (coordinate vector field), 384
h ; i (scalar product), 40 @i jp (coordinate vector), 377
h ; ig (inner product), 12 @r (radial vector field), 158, 179
@s (velocity of a transverse curve), 152,
197
A
@t (velocity of a main curve), 152, 197
A (adjoint of a matrix), 408
r (covariant derivative), 89
AT (transpose of a matrix), 408
r ? (normal connection), 231
A.TM / (set of connections on TM ), 94
r > (tangential connection), 93, 228
ad (adjoint representation of a Lie algebra),
r 2 u (covariant Hessian), 100
411 2
rX;Y (second total covariant derivative),
Ad (adjoint representation of a Lie group),
99
411
rF (total covariant derivative), 97
Alt (alternating
projection), 400
(covering automorphism group), rX Y (covariant derivative), 89
Aut M d (exterior derivative), 401
414 D (exterior covariant derivative), 209, 236
d (adjoint of d ), 52
B dFp (differential of a map), 377
[ (flat), 26–28 dg .p; q/ (Riemannian distance), 36
Bc .p/ (geodesic ball), 163 dg .p; S/ (distance to a subset), 174
Bx c .p/ (closed geodesic ball), 163 DM (density bundle), 406
Bn .R/ (Poincaré ball), 62 D .p/ (domain of .p/ ), 387
© Springer International Publishing AG 2018 419
J. M. Lee, Introduction to Riemannian Manifolds, Graduate Texts
in Mathematics 176, https://doi.org/10.1007/978-3-319-91755-9
420 Notation Index
Ds (covariant derivative along transverse gl.V / (Lie algebra of linear maps), 409
curves), 153 ŒG1 ; G2 (group generated by commuta-
Dt (covariant derivative along a curve), 101 tors), 72
d Vg (Riemannian density), 31 g1 ˚ f 2 g2 (warped product metric), 20
d Vg (Riemannian volume form), 30 g1 ˚ g2 (product metric), 20
div X (divergence of X ), 32 GL.V / (group of invertible linear maps),
408
E grad f (gradient of f ), 27
e (identity of a Lie group), 407
E (normal exponential map), 133 H
E (domain of the exponential map), 128 h (scalar second fundamental form), 235
E jM (restriction of a bundle), 384 H (mean curvature), 238
E.n/ (Euclidean group), 57 Hn (hyperbolic space), 66
EP (domain of the normal exponential Hn .R/ (hyperbolic space of radius R), 62,
map), 133 66
Ep (domain of the restricted exponential Hn (Heisenberg group), 72
map), 128 HV r (traceless Hessian operator), 337
End.V / (space of endomorphisms), 394 Hr (Hessian operator), 320
exp (exponential map), 128 Hr;s .R/ (pseudohyperbolic space), 79
expG (exponential map of a Lie group), 410 Hol.p/ (holonomy group), 150
expp (restricted exponential map), 128 Hol0 .p/ (restricted holonomy group), 150
F I
'rz (pullback of a connection), 110 iv (interior multiplication), 401
˚ 1 .fyg/ (level set), 380 In (identity matrix), 408
˚ 1 .y/ (level set), 380 I.V; W / (index form), 301
F (induced fundamental group homomor- ID.p/ (injectivity domain), 310
phism), 373 II (second fundamental form), 227
F (induced Lie algebra homomorphism), IIN (second fundamental form in direction
409 N ), 229
F (pushforward of vector fields), 385 inj.M / (injectivity radius of M ), 166
F (pullback of a density), 406 inj.p/ (injectivity radius at p), 165
F (pullback of a tensor field), 398 Iso.M; g/ (isometry group), 13
F G (symmetric product of F and G), 396 Isop .M; g/ (isotropy group of a point), 56
G J
0 (velocity vector), 386 J./ (space of Jacobi fields), 285
J ? ./ (normal Jacobi fields), 287
0 ai˙ (one-sided velocity vectors), 34
J > ./ (tangential Jacobi fields), 287
v (geodesic with initial velocity v), 105
.E / (smooth sections of E ), 89, 383 K
k
ij (connection coefficients), 91, 123
(Cayley transform), 66
.s; t/ (family of curves), 152
(geodesic curvature), 232
gy (pullback of the round metric), 293
i (principal curvatures), 4
x (Euclidean metric), 12
g
N (signed curvature), 273
gV (round metric), 16, 58 K (Gaussian curvature), 238
gV R (round metric of radius R), 58 Kn .R/ (Beltrami–Klein model), 62
gM (hyperbolic metric), 66
gM R (hyperbolic metric of radius R), 62, 66 L
gij (metric coefficients), 13 ƒk T M (bundle of alternating tensors),
g ij (inverse of gij ), 26 401
GL.n; C/ (complex general linear group), ƒk .V / (space of alternating tensors), 400
408 Lg ./ (length of curve), 34
GL.n; R/ (general linear group), 408 Lie.G/ (Lie algebra of left-invariant vector
gl.n; R/ (matrix Lie algebra), 409 fields on G), 408
Notation Index 421
M SC (surface of revolution), 20
jj (density associated with an n-form), Sc .p/ (geodesic sphere), 163
406 Sn (unit n-sphere), 16
M (M with opposite orientation), 405 Sn .R/ (n-sphere of radius R), 58
M=G (orbit space), 413 S ? (orthogonal complement), 10, 40
M.n k; R/ (space of matrices), 48 Sr;s .R/ (pseudosphere), 79
M.n; R/ (space of n n matrices), 411 SL.n; C/ (complex special linear group),
408
N SL.n; R/ (special linear group), 408
Np M (normal space), 16 SO.n/ (special orthogonal group), 408
NM (normal bundle), 16 SU.n/ (special unitary group), 408
sec.˘ / (sectional curvature), 8, 250
O sec.v; w/ (sectional curvature), 250
!i j (connection 1-forms), 113 sgn (sign of a permutation), 400
k .M / (space of k-forms), 401 Sol (3-dimensional solvable Lie group), 72
O.M / (orthonormal bases on M ), 56 supp f (support of f ), 376
O.n/ (orthogonal group), 408 Sym (symmetrization), 395
O.n; 1/ (Lorentz group), 67
OC .n; 1/ (orthochronous Lorentz group), T
67 (torsion tensor), 112, 121
O.n C 1/ (orthogonal group), 58 T n (n-torus), 389
TM (tangent bundle), 382
P TM jM (ambient tangent bundle), 386
1 .M; p/ (fundamental group), 373 Tp M (tangent space), 376
? (normal projection), 226 >
T.t / M (space of vectors tangent to a
> (tangential projection), 226 curve), 287
P (Schouten tensor), 215 ?
T.t / M (space of vectors normal to a
Pt0 t1 (parallel transport operator), 108 curve), 287
T k .M / (space of tensor fields), 397
Q T k TM (bundle of contravariant k-
x.r;s/ (pseudo-Euclidean metric), 43, 79
q tensors), 396
.r;s/
qM R (pseudohyperbolic metric), 79 T k T M (bundle of covariant k-tensors),
.r;s/
qV R (pseudospherical metric), 79 396
T k .V / (space of contravariant k-tensors),
R 393
./ (rotation index), 264 T k .V / (space of covariant k-tensors),
r (radial distance function), 158, 179 393
R (curvature endomorphism), 196 T .k;l/ TM (bundle of mixed tensors), 396
Rn (Euclidean space), 12, 57 T .k;l/ .V / (space of mixed tensors), 393
Rr;s (pseudo-Euclidean space), 43, 79 TCL.p/ (tangent cut locus), 310
RP n (real projective space), 413 tr (trace of a tensor), 395
R.V / (space of algebraic curvature ten- trg (trace with respect to g), 28
sors), 212
R.X; Y / (curvature endomorphism), 196 U
R.X; Y / (curvature endomorphism), 205 U.n/ (unitary group), 408
Rc (Ricci tensor), 207 U n (Poincaré half-space), 62
Rm (curvature tensor), 198
V
S V (dual space of V ), 391
†k T M (bundle of symmetric tensors), jv ^ wj (norm of an alternating 2-tensor),
396 250
†k .V / (space of symmetric tensors), 395 Vk .Rn / (Stiefel manifold), 48
s (shape operator), 235 Vol.M / (volume of a Riemannian mani-
S (scalar curvature), 208 fold), 30
422 Notation Index
complex projective space, 24, 82, 258, 367, contracted Bianchi identity, 209
413 contraction, 395
component functions, 383 contravariant tensor, 392
composable paths, 373 control theory, 46
conformal diffeomorphism, 59 convex hypersurface, 259
conformal invariance convex subset, 281
of the Cotton tensor, 219 convex, geodesically, 166, 281
of the Weyl tensor, 217 convexity radius, 186
conformal Laplacian, 223 coordinate representation, 376
conformal metrics, 59 coordinate ball
conformal transformation, 59 regular, 374
of the curvature, 217 smooth, 374
of the Levi-Civita connection, 217 coordinate chart, 371
conformally equivalent metrics, 59 smooth, 374
conformally flat, locally, 59, 218 coordinate domain, 374
hyperbolic space, 66 coordinate frame, 385
spheres, 61 coordinate vector, 377
conformally related metrics, 59 coordinate vector field, 384
congruent, 2, 142 coordinates, 375
conjugate point, 298 boundary slice, 380
critical point of expp , 299 Fermi, 136
geodesic not minimizing past, 303 graph, 18
conjugate point comparison theorem, 333, have upper indices, 375
339 local, 375
connection, 89 natural, on tangent bundle, 384
Euclidean, 92 normal, 132
existence of, 93 polar Fermi, 184
flat, 224 polar normal, 184
in a vector bundle, 89 semigeodesic, 181, 182
in components, 91 slice, 379, 380
in the tangent bundle, 91 standard, on Rn , 12
Koszul, 89 cosmological constant, 211
naturality of, 125 cotangent bundle, 396
normal, 231 cotangent space, 377
on a manifold, 91 Cotton tensor, 219
on tensor bundles, 95–97 conformal invariance of, 219
Riemannian, 122, 123 covariant derivative, 89
tangential, 92–93 along a curve, 101–103
connection 1-forms, 113, 274, 275 exterior, 209, 236
connection coefficients, 91 of tensor field, 95–97
transformation law for, 92 second, 99
constant Gaussian curvature, 6 total, 98
constant mean curvature, 242 covariant Hessian, 100, 112, 320
constant rank, 378 covariant tensor, 392
constant sectional curvature, 254 covector, 377, 391
characterization of, 8 covector field, 397
classification, 349, 350 covering automorphism, 414
formula for curvature tensor, 255 covering automorphism group, 414
formula for metric, 293 covering map, 171, 388
local uniqueness, 294 homeomorphism criterion, 390
model spaces, 8 normal, 414
uniqueness, 348 Riemannian, 24, 25
constant-speed curve, 35 smooth, 388
constraint equations, Einstein, 260 universal, 390
426 Subject Index
critical point, 27, 157, 210, 380 delta, Kronecker, 10, 392
of the exponential map, 299 density, 405–406
critical value, 380 on a manifold, 406
crystallographic groups, 350 on a vector space, 405
curl, 52 pullback of, 406
curvature, 2–8, 196 Riemannian, 31
conformal transformation of, 217 density bundle, 406
constant sectional, 8, 254, 255, 294, 348, derivation, 376
349 of C 1 .M /, 385
Gaussian, 5–6, 238 determinant convention for wedge product,
geodesic, 232 401
mean, 238 diameter, 39
of a curve in a manifold, 232, 233 diameter comparison theorem, 339
of a plane curve, 2 diffeomorphism, 375
principal, 4, 237 in Rn , 374
Ricci, 207 local, 378
Riemann, 196, 198 difference tensor, 94
scalar, 208 differentiable sphere theorem, 368
sectional, 8, 250, 251 differential
signed, 3, 273 global, 385
curvature 2-forms, 222 of a function, 377, 402
curvature endomorphism, 196, 224 of a map, 377
curvature operator, 262 differential Bianchi identity, 204
curvature tensor, 196, 198 differential form, 401
determined by sectional curvatures, 252 closed, 402
is isometry invariant, 199 exact, 402
symmetries of, 202 dihedral groups, 350
curve, 33, 386 direct product of Lie groups, 408
admissible, 34 directional derivative of a vector field on
in a manifold, 33 Rn , 86
piecewise regular, 33 Dirichlet eigenvalue, 51
plane, 2 Dirichlet’s principle, 52
smooth, 33 discrete Lie group, 408
curve segment, 33 discrete subgroup, 410
curved polygon, 265, 271 disjoint union, 382
curved triangle, 276 distance function, 174
cusp vertex, 265 has geodesic integral curves, 176
cut locus, 308 has unit gradient, 175
and injectivity radius, 312 local, 176
tangent, 310 distance, Riemannian, 36
cut point, 308 on a disconnected manifold, 36, 53
cut time, 307, 308 diverge to infinity, 187
is continuous, 309 divergence, 32
cylinder, principal curvatures, 4 in coordinates, 32
in terms of covariant derivatives, 148
D of a tensor field, 149
ı-pinched, 366 divergence theorem, 32, 50
pointwise, 368 domain of the exponential map, 128
strictly, 367 dot product, 9
Darboux theorem, 193 double elliptic geometry, 144
de Sitter space, 80 double of a manifold, 381
defining function, 380 dual basis, 392
local, 248, 380 dual coframe, 397
del, 89 dual space, 391
Subject Index 427
E extension
"-tubular neighborhood, 133 of functions, 380
edge of a curved polygon, 265 of sections of vector bundles, 384
effective group action, 412 exterior angle, 265, 271
eigenfunction of the Laplacian, 51 exterior covariant derivative, 209, 236
eigenvalue of the Laplacian, 51 exterior derivative, 401–402
Dirichlet, 51 naturality of, 402
Neumann, 51 extrinsic curvature of a curve, 233
Einstein constraint equations, 260
Einstein equation, 210 F
Einstein field equation, 211, 259 F -related, 385
Einstein metric, 210, 361 faithful representation, 408
2-dimensional, 216 family of curves, 152
Einstein summation convention, 375 admissible, 152
Einstein, Albert, 43, 211 Fermi coordinates, 136, 178
Einstein–Hilbert action, 211 polar, 184
embedded submanifold, 379 Fermi, Enrico, 136
embedding, 378 fiber
isometric, 15 of a map, 380
endomorphism of a vector bundle, 382
curvature, 196 fiber metric, 29
equivalent to a .1; 1/ tensor, 394 on a tensor bundle, 29
of a vector space, 394 on differential forms, 49
energy functional, 189 figure eight, 100, 101, 112
equivariant map, 412 Finsler metric, 47
equivariant rank theorem, 412 first Bianchi identity, 203
escape lemma, 387 first fundamental form, 227
Euclidean connection, 92 first structure equation, 113
Euclidean geodesics, 137 first variation, 154
Euclidean group, 57 flat ([), 26–28
Euclidean metric, 12, 57 flat connection, 224
Euclidean space, 57 flat metric, 12, 195, 200
sectional curvature of, 254 flat vertex, 265
Euclidean space form, 349 flatness criterion, 195
Euclidean triangle, 2 flow, 387
Euclidean, locally, 371 fundamental theorem on, 387
Euler characteristic, 276, 280 flow domain, 387
Euler–Lagrange equation, 157 focal point, 317
evenly covered neighborhood, 388 form
exact form, 402 bilinear, 40
existence and uniqueness closed, 402
for linear ODEs, 106 exact, 402
of geodesics, 103 forward reparametrization, 34
of Jacobi fields, 285 frame
exponential map coordinate, 385
differential of, 128 global, 383
domain of, 128 local, 383
naturality of, 130 frame-homogeneous, 56, 79
normal, 133 locally, 74
of a bi-invariant metric, 147 Frankel, Theodore, 315
of a Lie group, 147, 410 free group action, 412
of a Riemannian manifold, 128 free homotopy class, 174
extendible Riemannian manifold, 189 trivial, 174
extendible vector fields, 100 freely homotopic, 174
428 Subject Index
representation S
of a Lie algebra, 411 SSS theorem, 2
of a Lie group, 408 scalar curvature, 208
rescaling lemma, 127 geometric interpretation of, 253
restricted exponential map, 128 scalar product, 40
restricted holonomy, 150 scalar product space, 40
Riccati comparison theorem, 324–327 scalar second fundamental form, 235, 259
matrix, 325 Schoen, Richard, 211, 368
Riccati equation, 323, 343 Schouten tensor, 215
Riccati, Jacopo, 324 Schur’s lemma, 210
Ricci curvature, 207–209 secant angle function, 267
second Bianchi identity, 204
geometric interpretation of, 253
second covariant derivative, 99
Ricci decomposition of the curvature tensor,
second fundamental form, 227
216
geometric interpretation of, 233
Ricci flow, 368
scalar, 235, 259
Ricci identities, 205
second structure equation, 222
Ricci tensor, 207–209 second variation formula, 300
geometric interpretation of, 253 section
Riemann curvature endomorphism, 196 of a submersion, 378
Riemann curvature tensor, 198 of a vector bundle, 383
Riemann, Bernhard, v, 47, 255 smooth, 383
Riemannian connection, 122–125 sectional curvature, 8, 250, 251
Riemannian covering, 24, 25 constant, 254
Riemannian density, 31 determines the curvature tensor, 252
Riemannian distance, 36 of Euclidean space, 254
isometry invariance of, 37 of hyperbolic spaces, 254
on a disconnected manifold, 36, 53 of spheres, 254
Riemannian geodesics, 124 segment, curve, 33
Riemannian isometry, 186 self-dual, 50
Riemannian manifold, 1, 11 semi-Riemannian metric, 42
with boundary, 11 semicolon between indices, 98
Riemannian metric, 1, 11 semidirect product, 57, 412
existence of, 11 semigeodesic coordinates, 181, 182
Riemannian normal coordinates, 132 shape operator, 235, 259
Riemannian submanifold, 15, 225, 226 sharp (]), 26–28
of a pseudo-Riemannian manifold, 44 sheets of a covering, 388
Riemannian submersion, 22, 146, 224, 258 side of a curved polygon, 265
side-side-side theorem, 2
Riemannian symmetric space, 77, 189
sign conventions for curvature tensor, 198
locally, 78, 222, 295
sign of a permutation, 400
Riemannian volume form, 30
signature of symmetric bilinear form, 42
is parallel, 126
signed curvature, 3
right action, 411 of curved polygon, 273
right-invariant metric, 67 simple closed curve, 264
right-invariant vector field, 408 simply connected space, 373
right translation, 407 covering of, 390
rigid motion, 2 single elliptic geometry, 145
rotation index, 264 singular Riemannian metric, 46
of a curved polygon, 266, 272 slice coordinates, 379, 380
rotation index theorem, 266 boundary, 380
for a curved polygon, 272 smooth atlas, 374
rough section, 383 smooth chart, 374
round metric, 16 smooth coordinate ball, 374
Subject Index 435