Differential Geometry
Differential Geometry
DIFFERENTIAL GEOMETRY
28 March 2021
ii
Preface
These are notes for the lecture course “Differential Geometry I” given by the
second author at ETH Zürich in the fall semester 2017. They are based on
a lecture course1 given by the first author at the University of Wisconsin–
Madison in the fall semester 1983.
One can distinguish extrinsic differential geometry and intrinsic differ-
ential geometry. The former restricts attention to submanifolds of Euclidean
space while the latter studies manifolds equipped with a Riemannian metric.
The extrinsic theory is more accessible because we can visualize curves and
surfaces in R3 , but some topics can best be handled with the intrinsic theory.
The definitions in Chapter 2 have been worded in such a way that it is easy
to read them either extrinsically or intrinsically and the subsequent chapters
are mostly (but not entirely) extrinsic. One can teach a self contained one
semester course in extrinsic differential geometry by starting with Chapter 2
and skipping the sections marked with an asterisk like §2.8.
This document is designed to be read either as a .pdf file or as a printed
book.
We thank everyone who pointed out errors or typos in earlier versions
of this book. In particular, we thank Charel Antony and Samuel Trautwein
for many helpful comments.
1
Extrinsic Differential Geometry
iii
iv
Contents
2 Foundations 15
2.1 Submanifolds of Euclidean Space . . . . . . . . . . . . . . . . 15
2.2 Tangent Spaces and Derivatives . . . . . . . . . . . . . . . . . 24
2.2.1 Tangent Space . . . . . . . . . . . . . . . . . . . . . . 24
2.2.2 Derivative . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2.3 The Inverse Function Theorem . . . . . . . . . . . . . 31
2.3 Submanifolds and Embeddings . . . . . . . . . . . . . . . . . 34
2.4 Vector Fields and Flows . . . . . . . . . . . . . . . . . . . . . 38
2.4.1 Vector Fields . . . . . . . . . . . . . . . . . . . . . . . 38
2.4.2 The Flow of a Vector Field . . . . . . . . . . . . . . . 41
2.4.3 The Lie Bracket . . . . . . . . . . . . . . . . . . . . . 46
2.5 Lie Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.5.1 Definition and Examples . . . . . . . . . . . . . . . . . 52
2.5.2 The Lie Algebra of a Lie Group . . . . . . . . . . . . . 55
2.5.3 Lie Group Homomorphisms . . . . . . . . . . . . . . . 58
2.5.4 Closed Subgroups . . . . . . . . . . . . . . . . . . . . 62
2.5.5 Lie Groups and Diffeomorphisms . . . . . . . . . . . . 67
2.5.6 Smooth Maps and Algebra Homomorphisms . . . . . . 69
2.5.7 Vector Fields and Derivations . . . . . . . . . . . . . . 71
2.6 Vector Bundles and Submersions . . . . . . . . . . . . . . . . 72
2.6.1 Submersions . . . . . . . . . . . . . . . . . . . . . . . 72
2.6.2 Vector Bundles . . . . . . . . . . . . . . . . . . . . . . 74
v
vi CONTENTS
4 Geodesics 175
4.1 Length and Energy . . . . . . . . . . . . . . . . . . . . . . . . 175
4.1.1 The Length and Energy Functionals . . . . . . . . . . 175
4.1.2 The Space of Paths . . . . . . . . . . . . . . . . . . . . 178
4.1.3 Characterization of Geodesics . . . . . . . . . . . . . . 180
CONTENTS vii
5 Curvature 223
5.1 Isometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
5.2 The Riemann Curvature Tensor . . . . . . . . . . . . . . . . . 232
5.2.1 Definition and Gauß–Codazzi . . . . . . . . . . . . . . 232
5.2.2 Covariant Derivative of a Global Vector Field . . . . . 234
5.2.3 A Global Formula . . . . . . . . . . . . . . . . . . . . 237
5.2.4 Symmetries . . . . . . . . . . . . . . . . . . . . . . . . 239
5.2.5 Riemannian Metrics on Lie Groups . . . . . . . . . . . 241
5.3 Generalized Theorema Egregium . . . . . . . . . . . . . . . . 244
5.3.1 Pushforward . . . . . . . . . . . . . . . . . . . . . . . 244
5.3.2 Theorema Egregium . . . . . . . . . . . . . . . . . . . 245
5.3.3 Gaußian Curvature . . . . . . . . . . . . . . . . . . . . 250
5.4 The Group of Isometries* . . . . . . . . . . . . . . . . . . . . 255
5.4.1 The Myers–Steenrod Theorem . . . . . . . . . . . . . 255
5.4.2 The Topology on the Space of Isometries . . . . . . . 257
5.4.3 Killing Vector Fields . . . . . . . . . . . . . . . . . . . 260
5.4.4 Proof of the Myers–Steenrod Theorem . . . . . . . . . 264
5.4.5 Isometries of Compact Lie Groups . . . . . . . . . . . 273
5.4.6 Examples and Exercises . . . . . . . . . . . . . . . . . 278
5.5 Curvature in Local Coordinates* . . . . . . . . . . . . . . . . 281
viii CONTENTS
A Notes 351
A.1 Maps and Functions . . . . . . . . . . . . . . . . . . . . . . . 351
A.2 Normal Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . 352
A.3 Euclidean Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 354
References 357
Index 362
Chapter 1
What is Differential
Geometry?
1
2 CHAPTER 1. WHAT IS DIFFERENTIAL GEOMETRY?
U φ
Perhaps the user of such a map will be content to use the map to plot
the shortest path between two points p and q in U . This path is called a
geodesic and is denoted by γpq . It satisfies L(γpq ) = dU (p, q), where
where dE (φ(p), φ(q)) is the length of the shortest path in the plane. It is
also a consequence of the Theorema Egregium that there is no such chart.
Now suppose our user is content to have a map which makes it easy to
navigate close to the shortest path connecting two points. Ideally the user
would use a straight edge, magnetic compass, and protractor to do this.
S/he would draw a straight line on the map connecting p and q and steer a
course which maintains a constant angle (on the map) between the course
and meridians. This can be done by the method of stereographic projection.
This chart is conformal (which means that it preserves angles). According
to Wikipedia stereographic projection was known to the ancient Greeks
and a map using stereographic projection was constructed in the early 16th
century. Exercises 3.7.5, 3.7.12, and 6.4.22 use stereographic projection; the
latter exercise deals with the Poincaré model of the hyperbolic plane. The
hyperbolic plane provides a counterexample to Euclid’s Parallel Postulate.
Exercise 1.1.1. It is more or less obvious that for any surface M ⊂ R3
there is a unique shortest path in M connecting two points if they are
sufficiently close. (This will be proved in Theorem 4.5.3.) This shortest
path is called the minimal geodesic connecting p and q. Use this fact to
prove that the minimal geodesic joining two points p and q in S 2 is an arc
of the great circle through p and q. (This is the intersection of the sphere
with the plane through p, q, and the center of the sphere.) Also prove that
the minimal geodesic connecting two points in a plane is the straight line
1.1. CARTOGRAPHY AND DIFFERENTIAL GEOMETRY 3
n
p
φ(p)
segment connecting them. Hint: Both a great circle in a sphere and a line
in a plane are preserved by a reflection. (See also Exercise 4.2.5 below.)
Exercise 1.1.3. It may seem fairly obvious that you can’t draw an accurate
map of a portion of the earth because the sphere is curved. However the
cylinder C = {(x, y, z) ∈ R3 | x2 + y 2 = 1} is also curved, but the map
ψ : R2 → C defined by ψ(s, t) = (cos(t), sin(t), s) preserves lengths of curves,
i.e. L(ψ ◦ γ) = L(γ) for any curve γ : [a, b] → R2 . Prove this.
1.2 Coordinates
The rest of this chapter defines the category of smooth manifolds and smooth
maps between them. Before giving the precise definitions we will introduce
some terminology and give some examples.
A = {(φα , Uα )}α∈A
The idea is that if φ(p) = (x1 (p), . . . , xm (p)) for p ∈ U , then the func-
tions xi form a system of local coordinates defined on the subset U of M .
The dimension of M should be m since it takes m numbers to uniquely spec-
ify a point of U . We will soon impose conditions on charts (φ, U ), however
for the moment we are assuming nothing about the maps φ (other than that
they are bijective).
h : Ω → Rn
A = AT ∈ R(m+1)×(m+1)
(Here r + 1 is the rank of the matrix A.) The set M = F −1 (1) has an atlas
of 2m + 2 charts by the same construction as in Example 1.2.4, in fact S m
is the special case where A = 1lm+1 , the (m + 1) × (m + 1) identity matrix.
(See Example 2.1.12 below for another way to construct charts.)
Figure 1.3 enumerates the familiar quadric surfaces in R3 . The para-
boloids are examples of graphs as in Example 1.2.3 with Ω = R2 and n = 1,
and the ellipsoid and the two hyperboloids are instances of the quadric hyper-
surfaces defined in Example 1.2.5. The sphere is an instance of the ellipsoid
(with a = b = c = 1) and the cylinder is a limit of the ellipsoid as well as of
the elliptic hyperboloid of one sheet (as a = b = 1 and c → ∞). The pictures
were generated by computer using the parameterizations
for the elliptic hyperboloid of two sheets. These quadric surfaces will be
often used in the sequel to illustrate important concepts.
6 CHAPTER 1. WHAT IS DIFFERENTIAL GEOMETRY?
Unit Sphere x2 + y 2 + z 2 = 1
x2 y 2 z 2
Ellipsoid + 2 + 2 =1
a2 b c
Cylinder x2 + y 2 = 1
Elliptic
x2 y 2 z 2
Hyperboloid + 2 − 2 =1
a2 b c
(of one sheet)
Elliptic
x2 y 2 z 2
Hyperboloid + 2 − 2 = −1
a2 b c
(of two sheets)
Hyperbolic
x2 y 2
Paraboloid z= − 2
a2 b
Elliptic
x2 y 2
Paraboloid z= + 2
a2 b
In the following two examples K denotes either the field R of real numbers
or the field C of complex numbers, K∗ := {λ ∈ K | λ 6= 0} denotes the
corresponding multiplicative group, and V denotes a vector space over K.
Example 1.2.6. The projective space of V is the set of lines (through
the origin) in V . In other words,
P (V ) = {` ⊂ V | ` is a 1-dimensional K-linear subspace}
When K = R and V = Rm+1 this is denoted by RPm and when K = C
and V = Cm+1 this is denoted by CPm . For our purposes we can identify the
spaces Cm+1 and R2m+2 but the projective spaces CPm and RP2m are very
different. The various lines ` ∈ P (V ) intersect in the origin, however, after
the identification P (V ) = {[v] | v ∈ V \ {0}} with [v] := K∗ v = Kv \ {0} the
elements of P (V ) become disjoint, i.e. P (V ) is the set of equivalence classes
of an equivalence relation on the open set V \ {0}. Assume that V = Km+1
and define an atlas on P (V ) as follows. For each integer i = 0, 1, . . . , m
define Ui := {[v] | v = (x0 , . . . , xm ), xi 6= 0} and define φi : Ui → Km by
x0 xi−1 xi+1 xm
φi ([v]) = ,..., , ..., .
xi xi xi xi
This atlas consists of m + 1 charts.
Example 1.2.7. For each positive integer k the set
Gk (V ) := {` ⊂ V | ` is a k-dimensional K-linear subspace}
is called the Grassmann manifold of k-planes in V . Thus G1 (V ) = P (V ).
Assume that V = Kn and define an atlas on Gk (V ) as follows. Let e1 , . . . , en
be the standard basis for Kn , i.e. ei is the ith column of the n × n identity
matrix 1ln . Each partition {1, 2, . . . , n} = I ∪ J, I = {i1 < · · · < ik },
J = {j1 < · · · < jn−k } of the first n natural numbers determines a direct
sum decomposition Kn = V = VI ⊕ VJ via the formulas
VI = Kei1 + · · · + Keik , VJ = Kej1 + · · · + Kejn−k .
Let UI denote the set of all k-planes ` ∈ Gk (V ) which are transverse to VJ ,
i.e. such that ` ∩ VJ = {0}. The elements of UI are precisely those k-planes
of the form ` = graph(A), where A : VI → VJ is a linear map. Define the
map φI : UI → Kk×(n−k) by the formula
n−k
X
φI (`) = (ars ), Aeir = ars ejs , r = 1, . . . , k.
s=1
Exercise: Prove that the set of all pairs (φI , UI ) as I ranges over the subsets
of {1, . . . , n} of cardinality k form an atlas.
8 CHAPTER 1. WHAT IS DIFFERENTIAL GEOMETRY?
φ21 = φ2 ◦ φ−1
1 : φ1 (U1 ∩ U2 ) → φ2 (U1 ∩ U2 ) (1.4.2)
φ ◦ φ−1
1 : φ1 (U ∩ U1 ∩ U2 ) → φ(U ∩ U1 ∩ U2 ),
φ2 ◦ φ−1 : φ(U ∩ U1 ∩ U2 ) → φ2 (U ∩ U1 ∩ U2 )
Definitions 1.4.2 and 1.3.2 are mutatis mutandis the same, so every
smooth atlas on a set M is a fortiori a topological atlas, i.e. every smooth
manifold is a topological manifold. (See Lemma 1.3.3.) Moreover the defi-
nitions are worded in such a way that it is obvious that every smooth map
is continuous.
Exercise 1.4.4. Show that each of the atlases from the examples in §1.2 is a
smooth atlas. (You must show that the transition maps from Exercise 1.3.4
are smooth.)
12 CHAPTER 1. WHAT IS DIFFERENTIAL GEOMETRY?
A collection of sets and maps between them is called a category iff the
collection of maps contains the identity map of every set in the collection and
the composition of any two maps in the collection is also in the collection.
The sets are called the objects of the category and the maps are called the
morphisms of the category. An invertible morphism whose inverse is also in
the category is called an isomorphism. Some examples are the category of
all sets and maps, the category of topological spaces and continuous maps
(the isomorphisms are the homeomorphisms), the category of topological
manifolds and continuous maps between them, and the category of smooth
manifolds and smooth maps (the isomorphisms are the diffeomorphisms).
Each of the last three categories is a subcategory of the preceding one.
Often categories are enlarged by a kind of “gluing process”. For example,
the “global” category of smooth manifolds and smooth maps was constructed
from the “local” category of open sets in Euclidean space and smooth maps
between them via the device of charts and atlases. (The chain rule shows
1.5. THE MASTER PLAN 13
that this local category is in fact a category.) The point of Definition 1.3.2
is to show (via Lemma 1.3.3) that topological manifolds can be defined
in a manner analogous to the definition we gave for smooth manifolds in
Definition 1.4.2.
Other kinds of manifolds (and hence other kinds of geometry) are de-
fined by choosing other local categories, i.e. by imposing conditions on the
transition maps in Equation (1.4.2). For example, a real analytic manifold
is one where the transition maps are real analytic, a complex manifold is one
whose coordinate charts take values in Cn and whose transition maps are
holomorphic diffeomorphisms, and a symplectic manifold is one whose coor-
dinate charts take values in R2n and whose transition maps are canonical
transformations in the sense of classical mechanics. Thus CPn is a complex
manifold and RPn is a real analytic manifold.
Foundations
15
16 CHAPTER 2. FOUNDATIONS
U M M φ Ω
In Definition 2.1.3 we have used the fact that the domain of a smooth
map can be an arbitrary subset of Euclidean space and need not be open
(see page 15). The term m-manifold in Rk is short for m-dimensional sub-
manifold of Rk . In keeping with the Master Plan §1.5 we will sometimes say
manifold rather than submanifold of Rk to indicate that the context holds
in both the intrinsic and extrinsic settings.
Lemma 2.1.4. If M ⊂ Rk is a nonempty smooth m-manifold, then m ≤ k.
Proof. Fix an element p0 ∈ M , choose a coordinate chart φ : U ∩ M → Ω
with p0 ∈ U and values in an open subset Ω ⊂ Rm , and denote its in-
verse by ψ := φ−1 : Ω → U ∩ M . Shrinking U , if necessary, we may as-
sume that φ extends to a smooth map Φ : U → Rm . This extension satis-
fies Φ(ψ(x)) = φ(ψ(x)) = x and hence dΦ(ψ(x))dψ(x) = id : Rm → Rm for
all x ∈ Ω, by the chain rule. Hence the derivative dψ(x) : Rm → Rk is in-
jective for all x ∈ Ω, and hence m ≤ k because Ω is nonempty. This proves
Lemma 2.1.4.
2.1. SUBMANIFOLDS OF EUCLIDEAN SPACE 17
M := S 2 = (x, y, z) ∈ R3 | x2 + y 2 + z 2 = 1
φ(x, y, z) := (x, y)
U ∩ M = f −1 (0) = {p ∈ U | f (p) = 0} .
φ = (φ1 , φ2 , . . . , φk ) : U → Ω ⊂ Rk .
φ0 := (φ1 , . . . , φm )|U0 : U0 → Ω0 ,
and part (iii) holds with f := (φm+1 , . . . , φk ) : U → Rk−m . This shows that
part (ii) implies both (i) and (iii).
2.1. SUBMANIFOLDS OF EUCLIDEAN SPACE 19
x0 := φ0 (p0 ) ∈ Ω0 .
e ∩ W,
U := U e ∩ ψ −1 (W ).
Ω := Ω
ψ(x, y) ∈ M ⇐⇒ y = 0. (2.1.1)
Since the pairs (x, y) and (φ0 (p), 0) both belong to the set Ω
e and the re-
striction of ψ to Ω is injective we obtain x = φ0 (p) and y = 0. This
e
proves (2.1.1). It follows from (2.1.1) that the map φ := (ψ|Ω )−1 : U → Ω
satisfies φ(U ∩ M ) = {(x, y) ∈ Ω | ψ(x, y) ∈ M } = Ω ∩ (Rm × {0}). Thus we
have proved that (i) implies (ii).
20 CHAPTER 2. FOUNDATIONS
φ0 := φ|U 0 : U 0 → Ω0
and hence
φ0 (U 0 ∩ M ) = Ω0 ∩ Rm × {0} .
M := f −1 (c) = p ∈ U f (p) = c
f (x) := xT Ax.
Then df (x)ξ = 2xT Aξ for x, ξ ∈ Rk and hence the linear map df (x) : Rk → R
is surjective if and only if Ax 6= 0. Thus c = 0 is the only singular value of f
and, for c ∈ R \ {0}, the set
n o
M := f −1 (c) = x ∈ Rk xT Ax = c
is a smooth 5-manifold.
RP2 := S 2 /{±1}.
(See Example 1.2.6.) It is equipped with the quotient topology, i.e. a sub-
set U ⊂ RP2 is open, by definition, iff its preimage under the obvious projec-
tion S 2 → RP2 is an open subset of S 2 . Now the map f : S 2 → R6 defined
by
f (x, y, z) := (x2 , y 2 , z 2 , yz, zx, xy)
descends to a homeomorphism from RP2 onto M . The submanifold M is
covered by the local smooth parameterizations
p
Ω → M : (x, y) 7→ f (x, y, 1 − x2 − y 2 ),
p
Ω → M : (x, z) 7→ f (x, 1 − x2 − z 2 , z),
p
Ω → M : (y, z) 7→ f ( 1 − y 2 − z 2 , y, z),
onto the last three coordinates is called the Roman surface and was dis-
covered by Jakob Steiner. The Roman surface can also be represented as
the set of solutions (ξ, η, ζ) ∈ R3 of the equation η 2 ζ 2 + ζ 2 ξ 2 + ξ 2 η 2 = ξηζ.
It is not a submanifold of R3 .
Exercise: Prove this. Show that M is diffeomorphic to a submanifold of R4 .
Show that M is diffeomorphic to RP2 as defined in Example 1.2.6.
p + Tp M
Tp M v p
0
Figure 2.3: The tangent space Tp M and the translated tangent space p+Tp M .
Tp M = ker df (p).
To prove the first inclusion in (2.2.1), choose a constant r > 0 such that
Br (x0 ) := {x ∈ Rm | |x − x0 | < r} ⊂ Ω0 .
ε |ξ| ≤ r.
In particular, the tangent space of O(n) at the identity matrix is the space
of skew-symmetric matrices
n o
o(n) := T1l O(n) = ξ ∈ Rn×n | ξ T + ξ = 0
To see this, choose a smooth curve R → O(n) : t 7→ g(t). Then g(t)T g(t) = 1l
for all t ∈ R and, differentiating this identity with respect to t, we ob-
tain g(t)T ġ(t) + ġ(t)T g(t) = 0 for every t. Hence every matrix v ∈ Tg O(n)
satisfies the equation g T v + v T g = 0. With this understood, the claim fol-
lows from the fact that g T v + v T g = 0 if and only if the matrix ξ := g −1 v
is skew-symmetric and that the space of skew-symmetric matrices in Rn×n
has dimension n(n − 1)/2.
Exercise 2.2.11. Let Ω ⊂ Rm be an open set and h : Ω → Rk−m be a
smooth map. Prove that the tangent space of the graph of h at a point
(x, h(x)) is the graph of the derivative dh(x) : Rm → Rk−m :
M = {(x, h(x)) | x ∈ Ω} , T(x,h(x)) M = {(ξ, dh(x)ξ) | ξ ∈ Rm } .
28 CHAPTER 2. FOUNDATIONS
2.2.2 Derivative
A key purpose behind the concept of a smooth manifold is to carry over
the notion of a smooth map and its derivatives from the realm of first year
analysis to the present geometric setting. Here is the basic definition. It ap-
peals to the notion of a smooth map between arbitrary subsets of Euclidean
spaces as introduced on page 15.
f : M → R`
df (p) : Tp M → R`
γ:R→M
satisfying
γ(0) = p, γ̇(0) = v.
Now define the vector df (p)v ∈ R` by
d f (γ(h)) − f (p)
df (p)v := f (γ(t)) = lim . (2.2.2)
dt t=0 h→0 h
That the limit on the right in equation (2.2.2) exists follows from our
assumptions. We must prove, however, that the derivative is well defined,
i.e. that the right hand side of (2.2.2) depends only on the tangent vector v
and not on the choice of the curve γ used in the definition. This is the
content of the first assertion in the next theorem.
2.2. TANGENT SPACES AND DERIVATIVES 29
df (p)Tp M ⊂ Tf (p) N.
Let dF (p) ∈ R`×k denote the Jacobian matrix (i.e. the matrix of all first
partial derivatives) of F at p. Then, since γ(t) ∈ U ∩ M for t sufficiently
small, we have
dF (p)v = dF (γ(0))γ̇(0)
d
= F (γ(t))
dt t=0
d
= f (γ(t)).
dt t=0
The right hand side of this identity is independent of the choice of F while
the left hand side is independent of the choice of γ. Hence the right hand
side is also independent of the choice of γ and this proves (i). Assertion (ii)
follows immediately from the identity
df (p)v = dF (p)v
just established.
30 CHAPTER 2. FOUNDATIONS
M N
p0 f q0
U0 V0
φ0 ψ0
~
n ~ f ~ n
R U0 x0 V0 R
fe := ψ0 ◦ f ◦ φ−1
0 : U0 → V0
e e
Regular Values
Definition 2.2.18 (Regular value). Let M ⊂ Rk be a smooth m-manifold,
let N ⊂ R` be a smooth n-manifold, and let f : M → N be a smooth map.
An element q ∈ N is called a regular value of f iff, for every p ∈ M
with f (p) = q, the derivative df (p) : Tp M → Tf (p) N is surjective.
Theorem 2.2.19 (Regular Values). Let f : M → N be as in Defini-
tion 2.2.18 and let q ∈ N be a regular value of f . Then the set
P := f −1 (q) = {p ∈ M | f (p) = q}
and so v ∈ ker df (p). Hence Tp P ⊂ ker df (p) and equality holds because
both Tp P and ker df (p) are (m − n)-dimensional linear subspaces of Rk .
This proves Theorem 2.2.19.
2.2. TANGENT SPACES AND DERIVATIVES 33
are all surjective, hence so is their composition, and by the chain rule this
composition is the derivative df0 (x) : Rm → Rn . With this understood, it
follows from Theorem 2.1.10 that the set
f0−1 (c0 ) = x ∈ φ0 (U0 ) | f (φ−1
0 (x)) = q = φ0 (U0 ∩ P )
U0 φ 0(U0)
M φ0
p
0
P
P := f (N ), p0 := f (q0 ) ∈ P.
F (q, z) ∈ P ⇐⇒ z = 0. (2.3.2)
Proof. Choose any coordinate chart φ0 : U0 → Rm on an M -open neighbor-
hood U0 ⊂ M of p0 . Then d(φ0 ◦ f )(q0 ) = dφ0 (f (q0 )) ◦ df (q0 ) : Tq0 N → Rm
is injective. Hence there is a linear map B : Rm−n → Rm such that the map
F (q, z) := φ−1
0 (φ0 (f (q)) + Bz) .
This map is smooth, it satisfies F (q, 0) = f (q) for all q ∈ f −1 (U0 ), and
the derivative dF (q0 , 0) : Tq0 N × Rm−n → Tp0 M is the composition of the
map (2.3.3) with dφ0 (p0 )−1 : Rm → Tp0 M and so is a vector space isomor-
phism. Thus the Inverse Function Theorem 2.2.17 asserts that there is an
N -open neighborhood V0 ⊂ N of q0 and an open neighborhood W0 ⊂ Rm−n
of the origin such that V0 × W0 ⊂ Ω, the set U0 := F (V0 × W0 ) is M -open,
and the restriction of F to V0 × W0 is a diffeomorphism onto U0 . Thus we
have constructed a diffeomorphism F : V0 × W0 → U0 that satisfies (2.3.1).
We claim that the restriction of F to the product V × W of sufficiently
small open neighborhoods V ⊂ N of q0 and W ⊂ Rm−n of the origin also
satisfies (2.3.2). Otherwise, there exist sequences qi ∈ V0 converging to q0
and zi ∈ W0 \ {0} converging to zero such that F (qi , zi ) ∈ P . Hence there
exists a sequence qi0 ∈ N such that F (qi , zi ) = f (qi0 ). This sequence converges
to f (q0 ). Since f is proper we may assume, passing to a suitable subsequence
if necessary, that qi0 converges to a point q00 ∈ N . Then
Example 2.3.6. Let S 1 ⊂ R2 ∼ = C be the unit circle and consider the map
1 2
f : S → R given by f (x, y) := (x, xy). This map is a proper immersion but
is not injective (the points (0, 1) and (0, −1) have the same image under f ).
The image f (S 1 ) is a figure 8 in R2 and is not a submanifold (Figure 2.6).
X(p) ∈ Tp M
Exercise 2.4.2. Prove that the set of smooth vector fields on M is a real
vector space.
Example 2.4.3. Denote the standard cross product on R3 by
x 2 y3 − x 3 y2
x × y := x3 y1 − x1 y3
x 1 y2 − x 2 y1
X(p) := ξ × p, Y (p) := (ξ × p) × p.
These are vector fields with zeros ±ξ. Their integral curves (see Defini-
tion 2.4.6 below) are illustrated in Figure 2.8.
This vector field has a single zero at the origin and its integral curves are
illustrated in Figure 2.9.
2.4. VECTOR FIELDS AND FLOWS 39
for every t ∈ I.
for every t ∈ I.
(ii) If γ1 : I1 → M and γ2 : I2 → M are two solutions of (2.4.1) on open
intervals I1 and I2 containing 0, then γ1 (t) = γ2 (t) for every t ∈ I1 ∩ I2 .
40 CHAPTER 2. FOUNDATIONS
Ω := φ0 (U0 ) ⊂ Rm
This map is smooth and hence, by the basic existence and uniqueness the-
orem for ordinary differential equations in Rm (see [46]), the equation
I := I1 ∩ I2
D := {(t, p0 ) | p0 ∈ M, t ∈ I(p0 )}
K ⊂ U, (−ε, ε) × U ⊂ D,
Here the last implication follows from (2.4.3). Moreover, for p ∈ Uk and
t0 /N − ε < t < kt0 /N + ε, we have, by (2.4.4), that
Since φ(t0 /N, p) ∈ Uk−1 for p ∈ Uk the right hand side is a smooth map
on the open set (t0 /N − ε, kt0 /N + ε) × Uk . Since Uk ⊂ U , φ is also a
smooth map on (−ε, ε) × Uk and hence on (−ε, kt0 /N + ε) × Uk . This
completes the induction. With k = N we have found an open neighborhood
of (t0 , p0 ) contained in D, namely the set (−ε, t0 + ε) × UN , on which φ is
smooth. The case t0 ≤ 0 is treated similarly. This proves (i) and (ii) and
Theorem 2.4.9.
for 0 < s < t < T . Thus the limit p0 := limt&0 γ(t) exists in Rk and, since K
is a closed subset of Rk , we have p0 ∈ K ⊂ M . Define γ0 : [0, T ) → M by
p0 , for t = 0,
γ0 (t) :=
γ(t), for 0 < t < T.
We prove that γ0 is differentiable at t = 0 and γ̇0 (0) = X(p0 ). To see this, fix
a constant ε > 0. Since the curve [0, T ) → Rk : t 7→ X(γ(t)) is continuous,
there exists a constant δ > 0 such that
0<t≤δ =⇒ |X(γ(t)) − X(p0 )| ≤ ε.
Hence, for 0 < s < t ≤ δ, we have
Z t
|γ(t) − γ(s) − (t − s)X(p0 )| = (γ̇(r) − X(p0 )) dr
Zs t
= (X(γ(r)) − X(p0 )) dr
s
Z t
≤ |X(γ(r)) − X(p0 )| dr
s
≤ (t − s)ε.
Take the limit s → 0 to obtain
γ(t) − p0 |γ(t) − γ(s) − (t − s)X(p0 )|
− X(p 0 = lim
) ≤ε
t s→0 t−s
for 0 < t ≤ δ. Thus γ0 is differentiable at t = 0 with γ̇0 (0) = X(p0 ), as
claimed. Hence γ extends to an integral curve γ e : (−ρ, T ) → M of X for
some ρ > 0 via γe(t) := φ(t, p0 ) for −ρ < t ≤ 0 and γ
e(t) := γ(t) for 0 < t < T .
Here φ is the flow of X. That γ also extends beyond t = T , follows by re-
placing γ(t) with γ(T −t) and X with −X. This proves Corollary 2.4.15.
2.4. VECTOR FIELDS AND FLOWS 45
Diff(M ) := {φ : M → M | φ is a diffeomorphism} .
This is a group. The group operation is composition and the neutral element
is the identity. Now equation (2.4.5) asserts that the flow of a complete
vector field X ∈ Vect(M ) is a group homomorphism
R → Diff(M ) : t 7→ φt .
d t
φ (p) = X(φt (p)), φ0 (p) = p
dt
for all p ∈ M and t ∈ R. We will often abbreviate this equation in the form
d t
φ = X ◦ φt , φ0 = id. (2.4.6)
dt
I × M → Rk : (t, p) 7→ Xt (p)
be a smooth map such that Xt ∈ Vect(M ) for every t. Prove that there is
a smooth family of diffeomorphisms I × M → M : (t, p) 7→ φt (p) satisfying
d
φt = Xt ◦ φt , φ0 = id (2.4.7)
dt
I → Diff(M ) : t 7→ φt
for q ∈ N .
φ∗ X = (φ−1 )∗ X (2.4.10)
and
(ψ ◦ φ)∗ X = ψ∗ φ∗ X, (ψ ◦ φ)∗ Z = φ∗ ψ ∗ Z. (2.4.11)
for all q ∈ N (Corollary 2.2.15) and the equations in (2.4.11) follow directly
from the chain rule (Theorem 2.2.14). This proves Lemma 2.4.17.
X : M → Rk
dX(p) : Tp M → Rk .
In general, this derivative will no longer take values in the tangent space
Tp M . However, if we have two vector fields X and Y on M , then the next
lemma shows that the difference of the derivative of X in the direction Y
and of Y in the direction X does take values in the tangent spaces of M .
2.4. VECTOR FIELDS AND FLOWS 47
R → Diff(M ) : t 7→ φt , R → Diff(M ) : t 7→ ψ t
∂2β
γ̈(0) = 2 (0, 0). (2.4.16)
∂s∂t
48 CHAPTER 2. FOUNDATIONS
In both cases the right hand side is the derivative of a smooth curve in the
tangent space Tp M and so is itself an element of Tp M . Moreover, we have
1 ∂
γ̈(0) = dφs (φ−s (p))Y (φ−s (p))
2 ∂s s=0
∂ ∂
= φs ◦ ψ t ◦ φ−s (p)
∂s s=0 ∂t t=0
∂ ∂
= φs ◦ ψ t ◦ φ−s (p)
∂t t=0 ∂s s=0
∂
X(ψ t (p)) − dψ t (p)X(p)
=
∂t t=0
∂ ∂
= dX(p)Y (p) − ψ t ◦ φs (p)
∂t t=0 ∂s s=0
∂ ∂
= dX(p)Y (p) − ψ t ◦ φs (p)
∂s s=0 ∂t t=0
∂
= dX(p)Y (p) − Y (φs (p))
∂s s=0
= dX(p)Y (p) − dY (p)X(p).
[ξ, η] := ξη − ηξ.
in Lemma 2.4.18, where φt and ψ t are the flows of the vector fields X and Y ,
respectively. Since γ̇(0) = 0, Exercise 2.4.19 asserts that
√ √ √ √
1 d
[X, Y ](p) = γ̈(0) = φ t ◦ ψ t ◦ φ− t ◦ ψ − t (p). (2.4.21)
2 dt t=0
Exercise 2.5.4. Prove that GL(n, C) can be identified with the group
0 −1l
G := {Φ ∈ GL(2n, R) | ΦJ0 = J0 Φ} , J0 := .
1l 0
Hint: Use the isomorphism Rn × Rn → Cn : (x, y) 7→ x + iy. Show that a
matrix Φ ∈ R2n×2n commutes with J0 if and only if it has the form
X −Y
Φ= , X, Y ∈ Rn×n .
Y X
What is the relation between the real determinant of Φ and the complex
determinant of X + iY ?
Exercise 2.5.5. Let J0 be as in Exercise 2.5.4 and define
n o
Sp(2n) := Ψ ∈ GL(2n, R) | ΨT J0 Ψ = J0 .
This is the symplectic linear group. Prove that Sp(2n) is a Lie group.
Hint: See [36, Lemma 1.1.12].
Example 2.5.6 (Unit Quaternions). The Quaternions form a four-
dimensional associative unital algebra H, equipped with a basis 1, i, j, k.
The elements of H are vectors of the form
x = x0 + ix1 + jx2 + kx3 x0 , x1 , x2 , x3 ∈ R. (2.5.1)
The product structure is the bilinear map H × H → H : (x, y) 7→ xy, deter-
mined by the relations
i2 = j2 = k2 = −1, ij = −ji = k, jk = −kj = i, ki = −ik = j.
This product structure is associative but not commutative. The quaternions
are equipped with an involution H → H : x 7→ x̄, which assigns to a quater-
nion x of the form (2.5.1) its conjugate x̄ := x0 − ix1 − jx2 − kx3 . This
involution satisfies the conditions
x + y = x̄ + ȳ, xy = ȳx̄, xx̄ = |x|2 , |xy| = |x| |y|
p
for x, y ∈ H, where |x| := x20 + x22 + x22 + x23 denotes the Euclidean norm
of the quaternion (2.5.1). Thus the unit quaternions form a group
Sp(1) := {x ∈ H | |x| = 1}
with the inverse map x 7→ x̄. Note that the group Sp(1) is diffeomorphic
to the 3-sphere S 3 ⊂ R4 under the isomorphism H ∼ = R4 . Warning: The
unit quaternions (a compact Lie group) are not to be confused with the
symplectic linear group in Exercise 2.5.5 (a noncompact Lie group) despite
the similarity in notation.
54 CHAPTER 2. FOUNDATIONS
G × G → G : (g, h) 7→ gh, G → G : g 7→ g −1
are smooth (see [47]). Fixing an element h ∈ G we find that the derivative
of the map G → G : g 7→ gh at g ∈ G is given by the linear map
Here gb and h are both matrices in Rn×n and gbh denotes the matrix prod-
uct. In fact, if gb ∈ Tg G, then, since G is a manifold, there exists a smooth
curve γ : R → G with γ(0) = g and γ̇(0) = gb. Since G is a group we obtain
a smooth curve β : R → G given by β(t) := γ(t)h. It satisfies β(0) = gh and
so gbh = β̇(0) ∈ Tgh G.
The linear map (2.5.2) is obviously a vector space isomorphism whose
inverse is given by right multiplication with h−1 . It is sometimes convenient
to define the map Rh : G → G by
Rh (g) := gh
dRh (g)b
g = gbh for gb ∈ Tg G.
dLg (h)b
h = gb
h h ∈ Th G.
for b (2.5.3)
Tg G → Tg−1 G : v 7→ −g −1 vg −1 .
g = Lie(G) := T1l G.
This terminology is justified by the fact that g is in fact a Lie algebra, i.e.
it is invariant under the standard Lie bracket operation
[ξ, η] := ξη − ηξ
on the space Rn×n of square matrices (see Lemma 2.5.9 below). The proof
requires the notion of the exponential matrix. For ξ ∈ Rn×n and t ∈ R
we define
∞ k k
X t ξ
exp(tξ) := . (2.5.4)
k!
k=0
A standard result in first year analysis asserts that this series converges
absolutely (and uniformly on compact t-intervals), that the map
R → Rn×n : t 7→ exp(tξ)
d
exp(tξ) = ξ exp(tξ) = exp(tξ)ξ, (2.5.5)
dt
and that
exp((s + t)ξ) = exp(sξ) exp(tξ), exp(0ξ) = 1l (2.5.6)
for all s, t ∈ R. This shows that the matrix exp(tξ) is invertible for each t
and that the map R → GL(n, R) : t 7→ exp(tξ) is a group homomorphism.
In other words, the infinitesimal Lie group commutator is the matrix com-
mutator. (Compare Equations (2.5.7) and (2.4.21).)
56 CHAPTER 2. FOUNDATIONS
Xξ (g) := ξg ∈ Tg G, g ∈ G. (2.5.8)
By (2.5.5), the curve (−ε, ε) → Rn×n : t 7→ exp(tξ) satisfies the same initial
value problem and hence, by uniqueness, we have exp(tξ) = γ(t) ∈ G for
all t ∈ R with |t| < ε. Now let t ∈ R and choose N ∈ N such that Nt < ε.
Then exp( Nt ξ) ∈ G and hence it follows from (2.5.6) that
N
t
exp(tξ) = exp ξ ∈ G.
N
This proves (i).
We prove (ii). Consider the smooth curve γ : R → Rn×n defined by
γ(t) := g exp(tη)g −1 .
gηg −1 = γ̇(0) ∈ g.
Hence γ is the integral curve of the vector field Xξ in (2.5.8) with γ(0) = 1l.
This implies γ(t) = exp(tξ) for every t ∈ R, as claimed.
Example 2.5.11. Since the general linear group GL(n, R) is an open subset
of Rn×n its Lie algebra is the space of all real n × n-matrices
(see Exercise 2.2.9) and the Lie algebra of the special orthogonal group is
n o
so(n) := Lie(SO(n)) = ξ ∈ gl(n, R) ξ T + ξ = 0 = o(n)
These are vector spaces over the reals. Determine their real dimensions.
Which of these are also complex vector spaces?
Exercise 2.5.13. Let G ⊂ GL(n, R) be a subgroup. Prove that G is a Lie
group if and only if it is a closed subset of GL(n, R) in the relative topology.
58 CHAPTER 2. FOUNDATIONS
Lemma 2.5.16. Let G and H be Lie groups and denote their Lie algebras
by g := Lie(G) and h := Lie(H). Let ρ : G → H be a Lie group homomor-
phism and denote its derivative at 1l ∈ G by
ρ̇ := dρ(1l) : g → h.
Show that ρ is a Lie group homomorphism. Find a formula for the map
ρ̇ := dρ(1l) : sp(1) → so(3)
and show that it is a Lie algebra isomorphism. For x, y ∈ Sp(1) prove
that ρ(x) = ρ(y) if and only if y = ±x.
2.5. LIE GROUPS 61
Now suppose that g = Lie(G) is the Lie algebra of a Lie group G. Then
there is a map Ad : G → Aut(g) defined by
for g ∈ G and η ∈ g. Part (ii) of Lemma 2.5.9 asserts that Ad(g) maps g
to itself for every g ∈ G. It follows directly from the definitions that the
map Ad(g) : g → g is a Lie algebra automorphism for every g ∈ G and that
the map Ad : G → Aut(g) is a Lie group homomorphism. The associated
Lie algebra homomorphism is the linear map ad : g → Der(g) defined by
˙ we compute
for ξ, η ∈ g. To verify the equation ad = Ad
˙ d d
Ad(ξ)η = Ad(exp(tξ))η = exp(tξ)η exp(−tξ) = [ξ, η].
dt t=0 dt t=0
Exercise 2.5.23. Let g be any Lie algebra. Define the map ad : g → End(g)
by (2.5.16) and prove that the endomorphism ad(ξ) : g → g is a derivation
for every ξ ∈ g. Prove that ad : g → Der(g) is a Lie algebra homomorphism.
Proof of Theorem 2.5.26 (i) =⇒ (ii) and (2.5.17). Assume that H is a Lie
subgroup of G and let h ⊂ g be defined by (2.5.17). We prove that h is
the Lie algebra of H. Assume first that η ∈ h. Then the curve γ : R → G
defined by γ(t) := exp(tη) for t ∈ R takes values in H and satisfies γ(0) = 1l
and γ̇(0) = η, and this implies η ∈ T1l H = Lie(H). Conversely, if η ∈ Lie(H),
then Lemma 2.5.9 asserts that exp(tη) ∈ H for all t ∈ R and hence η ∈ h.
This shows that h = Lie(H).
Next we prove in three steps that H is a closed subset of G. Choose any
inner product on g, denote by |·| the associated norm, and denote by h⊥ ⊂ g
the orthogonal complement of h with respect to this inner product.
Step 1. There exist open neighborhoods V ⊂ H of 1l and W ⊂ h⊥ of the
origin such that the map φ : V × W → G, defined by
φ(h, ξ) := h exp(ξ)
for k ∈ N. Then
γ(t/k) − γ(0)
lim ξk = t lim = tγ̇(0) = tξ
k→∞ k→∞ t/k
and hence
ξk k
exp(tξ) = lim 1l + = lim γ(t/k)k .
k→∞ k k→∞
(See [47, Satz 1.5.2].) This proves Lemma 2.5.27.
64 CHAPTER 2. FOUNDATIONS
exp(ξi ) ∈ H, ξi 6= 0
Then ξ ∈ h.
Proof. Fix a real number t. Then, for each i ∈ N, there exists a unique
integer mi ∈ Z such that mi τi ≤ t < (mi + 1)τi . The sequence mi satisfies
ξi
lim mi τi = t, lim mi ξi = lim mi τi = tξ
i→∞ i→∞ i→∞ τi
and hence
Proof of Theorem 2.5.26 (ii) =⇒ (i). Choose any inner product on g and
define a Riemannian metric on G by |v| := |g −1 v| for g ∈ G and v ∈ Tg G.
Let H ⊂ G be a closed subgroup of G and define the set h ⊂ g by (2.5.17).
Then h is a Lie subalgebra of g by Lemma 2.5.28. Define
k := dim(h), ` := dim(g) ≥ k,
and choose a basis η1 , . . . , η` of g such that the vectors η1 . . . . , ηk form a basis
of h and ην ∈ h⊥ for ν > k. Let h0 ∈ H and define the map Θ : R` → G by
Θ(t1 , . . . , t` ) := h0 exp(t1 η1 + · · · + tk ηk ) exp(tk+1 ηk+1 + · · · + t` η` ).
Then Θ(0) = h0 , Θ(Rk × {0}) ⊂ H, and the derivative dΘ(0) : R` → Th0 G
is bijective. Hence the inverse function theorem asserts that Θ restricts to a
diffeomorphism from an open neighborhood Ω ⊂ R` of the origin to an open
neighborhood U := Θ(Ω) ⊂ G of h0 that satisfies
Θ Ω ∩ (Rk × {0}) ⊂ U ∩ H.
Θ(0) = h0 ,
We prove the following.
Claim. There exists an open set Ω0 ⊂ R` such that
Θ Ω0 ∩ (Rk × {0}) = U0 ∩ H,
0 ∈ Ω0 ⊂ Ω, U0 := Θ(Ω0 ). (2.5.19)
Assume, by contradiction, that such an open set Ω0 does not exist. Then
there exists a sequence ti = (t1i , . . . , t`i ) ∈ R` such that
lim ti = 0, ti ∈ Ω \ (Rk × {0}), Θ(ti ) ∈ H.
i→∞
Define
k `
!
X X
hi := h0 exp tνi ην ∈ H, ξi := tνi ην ∈ h⊥ \ {0}.
ν=1 ν=k+1
G → Diff(M ) : g 7→ φg . (2.5.27)
where φtdenotes the flow of X. Since the map (2.5.29) is the derivative
of the “Lie group” anti-homomorphism (2.5.28) we expect it to be a Lie
algebra anti-homomorphism. Indeed, one can show that
L[X,Y ] = LY LX − LX LY = −[LX , LY ] (2.5.30)
for X, Y ∈ Vect(M ). This confirms that our sign in the definition of the
Lie bracket is consistent with the standard conventions in the theory of
Lie groups. In the literature the difference between a vector field and the
associated derivation LX is sometimes neglected in the notation and many
authors write Xf := df · X = LX f , thus thinking of a vector field on a
manifold M as an operator on the space of functions. With this notation one
obtains the equation [X, Y ]f = Y (Xf ) − X(Y f ) and here lies the origin for
the use of the opposite sign for the Lie bracket in many books on differential
geometry.
Exercise 2.5.46. Prove that the map (2.5.29) is bijective. Hint: Fix a
derivation δ ∈ Der(F (M )) and prove the following. Fact 1: If U ⊂ M is
an open set and f ∈ F (M ) vanishes on U , then δ(f ) vanishes on U . Fact 2:
If p ∈ M and the derivative df (p) : Tp M → R is zero, then (δ(f ))(p) = 0.
(By Fact 1, the proof of Fact 2 can be reduced to an argument in local
coordinates.)
Exercise 2.5.47. Verify the formula (2.5.30).
72 CHAPTER 2. FOUNDATIONS
df (q) : Tq N → Tf (q) M
N q0
g f
M U p0
sends w ∈ Tq0 N to (df (q0 )w, Aw) and is therefore bijective. Hence it follows
from the inverse function theorem for manifolds (Theorem 2.2.17) that there
exists an N -open neighborhood V ⊂ N of q0 such that the set
W := ψ(N ) ⊂ M × Rn−m
g(p0 ) = ψ −1 (p0 , 0) = q0 .
This shows that (i) implies (ii). The converse is an easy consequence of the
chain rule and is left to the reader. This proves Lemma 2.6.1
Π = Π2 = ΠT , im Π = V. (2.6.1)
Prove that there is a unique matrix Π ∈ R`×` satisfying (2.6.1). Prove that,
for every symmetric matrix S = S T ∈ R`×` , the kernel of S is the orthogonal
complement of the image of S. If D ∈ R`×n is any injective matrix whose
image is V , prove that det(DT D) 6= 0 and
π −1 (U )
Φ / U × Rn .
NNN pp
NNN
N ppppp
π NNNN p
ppp pr1
NN' pw pp
U
Corollary 2.6.12. Let M ⊂ Rk be a smooth m-manifold. Then T M is a
vector bundle over M and hence is a smooth 2m-manifold in Rk × Rk .
Proof. Let φ : U → Ω be a coordinate chart on an M -open set U ⊂ M with
values in an open subset Ω ⊂ Rm . Denote its inverse by ψ := φ−1 : Ω → M .
By Theorem 2.2.3 the linear map dψ(x) : Rm → Rk is injective and its image
is Tψ(x) M for every x ∈ Ω. Hence the map D : U → Rk×m defined by
D(p) := dψ(φ(p)) ∈ Rk×m
is smooth and, for every p ∈ U , the linear map D(p) : Rm → Rk is injec-
tive and its image is Tp M . Thus the function ΠT M : M → Rk×k defined
by (2.6.4) with Ep = Tp M is given by
−1
ΠT M (p) = D(p) D(p)T D(p) D(p)T for p ∈ U.
Corollary 2.6.13. The pullback and normal bundles are vector bundles.
Proof. Let Π = ΠE : M → Rd×d be the map defined by (2.6.4). This map
is smooth by Theorem 2.6.10. Moreover, the corresponding maps for f ∗ E
and E ⊥ are given by
∗E ⊥
Πf = ΠE ◦ f : N → Rd×d , ΠE = 1l − ΠE : M → Rd×d .
These maps are smooth and hence it follows from Theorem 2.6.10 that f ∗ E
and E ⊥ are vector bundles.
2.6. VECTOR BUNDLES AND SUBMERSIONS 77
π = π ◦ fλ
and hence
Since dfλ (p, v) is bijective and dπ(p, λv) is surjective for λ < ε/ |v| it follows
that dπ(p, v) is surjective for every p ∈ M and every v ∈ Ep . Thus the
projection π : E → M is a submersion for every vector bundle E over M .
We prove that (i) implies (ii). Let p0 ∈ M and v0 ∈ Ep0 . We have
already proved that π is a submersion. Hence it follows from Lemma 2.6.1
that there exists an M -open neighborhood U ⊂ M of p0 and a smooth map
σ0 : U → E
such that
π ◦ σ0 = id : U → U, σ0 (p0 ) = (p0 , v0 ).
Define the map s0 : U → R` by
Then s0 (p0 ) = v0 and s0 (p) ∈ Ep for all p ∈ U . Now choose ε > 0 such that
{p ∈ M | |p − p0 | < ε} ⊂ U
We prove that (ii) implies (iii). Thus we assume that E satisfies (ii).
Choose p0 ∈ M and a basis v1 , . . . , vn of Ep0 . By (ii) there exists smooth
sections s1 , . . . , sn : M → R` of E such that si (p0 ) = vi for i = 1, . . . , n. Now
there exists an M -open neighborhood U ⊂ M of p0 such that the vec-
tors s1 (p), . . . , sn (p) are linearly independent, and hence form a basis of Ep
for every p ∈ U . Hence, for every p ∈ U , we have
Ep = imD(p), D(p) := [s1 (p) · · · sn (p)] ∈ R`×n .
By Exercise 2.6.9, this implies Π(p) = D(p)(D(p)T D(p))−1 D(p)T for ev-
ery p ∈ U . Thus every p0 ∈ M has a neighborhood U such that the re-
striction of Π to U is smooth. This shows that (ii) implies (iii).
We prove that (iii) implies (iv). Fix a point p0 ∈ M and choose a ba-
sis v1 , . . . , vn of Ep0 . For p ∈ M define
D(p) := [Π(p)v1 · · · Π(p)vn ] ∈ R`×n
Then D : M → R`×n is a smooth map and D(p0 ) has rank n. Hence the set
U := {p ∈ M | rankD(p) = n} ⊂ M
is an open neighborhood of p0 and Ep = imD(p) for all p ∈ U . Thus
π −1 (U ) = {(p, v) ∈ E | p ∈ U } ⊂ E
is an open set containing π −1 (p0 ). Define the map Φ : π −1 (U ) → U × Rn by
−1/2
Φp (v) := D(p)T D(p) D(p)T v
Φ(p, v) := p, Φp (v) ,
for p ∈ U and v ∈ Ep . This map is bijective and its inverse is given by
−1/2
Φ−1 (p, ξ) = p, Φ−1 −1 T
p (ξ) , Φ p (ξ) = D(p) D(p) D(p) ξ
v ∈ Ep ⇐⇒ dφ(p)v ∈ Rn × {0}.
U φ Ω
It is easy to show that (iii) =⇒ (ii) =⇒ (i) (see below). The hard part of
the theorem is to prove that (i) =⇒ (iii). This requires the following lemma.
Lemma 2.7.3. Let E ⊂ T M be an involutive subbundle and X ∈ Vect(M )
be a complete vector field such that X(p) ∈ Ep for every p ∈ M . Denote by
R → Diff(M ) : t 7→ φt
Lemma 2.7.3 implies Theorem 2.7.2. We prove first that (iii) implies (ii).
Let p0 ∈ M , choose a foliation box φ : U → Ω for E with p0 ∈ U , and define
N := (p ∈ U | φ(p) ∈ Rn × {y0 }}
∂ψ
(x, y) ∈ Eψ(x,y) (2.7.3)
∂xi
∂ψ ∂ψ
(0, 0) = Xi (p0 ), (0, 0) = Yj (p0 ),
∂xi ∂yj
U := ψ(Ω) ⊂ M
φ := (ψ|Ω )−1 : U → Ω
is a foliation box and this proves Theorem 2.7.2, assuming Lemma 2.7.3.
∂β ∂β
(s, 0) ∈ Eβ(s,0) , (s, t) ∈ Eβ(s,t) , (2.7.4)
∂s ∂t
for all s, t ∈ R, then
∂β
(s, t) ∈ Eβ(s,t) , (2.7.5)
∂s
for all s, t ∈ R.
E|U := {(p, v) | p ∈ U, v ∈ Ep }
e is involutive the Lie bracket [ζ, ζ 0 ] must take values in the graph
Since E
of A. Hence the right hand side vanishes and this proves Claim 1.
∂y ∂x
(s0 , t0 ) = A x(s0 , t0 ), y(s0 , t0 ) (s0 , t0 ), (2.7.7)
∂s ∂s
∂y ∂x
(s, t) = A x(s, t), y(s, t) (s, t) (2.7.8)
∂t ∂t
∂y ∂x
(s0 , t) = A x(s0 , t), y(s0 , t) (s0 , t) (2.7.9)
∂s ∂s
for all t ∈ J.
2.7. THE THEOREM OF FROBENIUS 85
over p = (x, y, z) ∈ R3 is not integrable and that any two points in R3 can
be joined by a path tangent to E.
Exercise 2.7.9. Consider the manifold M = S 3 ⊂ R4 = C2 and define
E := (z, ζ) ∈ C2 × C2 | |z| = 1, ζ ⊥ z, iζ ⊥ z .
Uα Uβ
φα φβ
φβα
φ21 = φ2 ◦ φ−1
1 : φ1 (U1 ∩ U2 ) → φ2 (U1 ∩ U2 ) (2.8.1)
W ⊂M
of real lines in Rn+1 . It can again be identified with the quotient space
RPn = Rn+1 \ {0} /R∗
Tn := Rn /Zn
equipped with the quotient topology. Thus two vectors x, y ∈ Rn are equiv-
alent iff their difference x − y ∈ Zn is an integer vector and we denote by
π : Rn → Tn the obvious projection which assigns to each vector x ∈ Rn its
equivalence class
π(x) := [x] := x + Zn .
Then a set U ⊂ Tn is open if and only if the set π −1 (U ) is an open subset
of Rn . An atlas on Tn is given by the open cover
denotes the set of unitary k-frames in Cn and the group U(k) acts on Fk (Cn )
contravariantly by D 7→ Dg for g ∈ U(k). The projection
π : Fk (Cn ) → Gk (Cn )
Example 2.8.9 (The real line with two zeros). A topological space M
is called Hausdorff iff any two points in M can be separated by disjoint
open neighborhoods. This example shows that a manifold need not be a
Hausdorff space. Consider the quotient space
M := R × {0, 1}/ ≡
where [x, 0] ≡ [x, 1] for x 6= 0. An atlas on M consists of two coordinate
charts φ0 : U0 → R and φ1 : U1 → R where
Ui := {[x, i] | x ∈ R} , φi ([x, i]) := x
for i = 0, 1. Thus M is a 1-manifold. But the topology on M is not
Hausdorff, because the points [0, 0] and [0, 1] cannot be separated by disjoint
open neighborhoods.
Example 2.8.10 (A 2-manifold without a countable atlas). Consider
the vector space X = R × R2 with the equivalence relation
either y1 = y2 6= 0, t1 + x1 y1 = t2 + x2 y2
[t1 , x1 , y1 ] ≡ [t2 , x2 , y2 ] ⇐⇒
or y1 = y2 = 0, t1 = t2 , x1 = x2 .
For y 6= 0 we have [0, x, y] ≡ [t, x − t/y, y], however, each point (x, 0) on
the x-axis gets replaced by the uncountable set R × {(x, 0)}. Our manifold
is the quotient space M := X/ ≡. This time we do not use the quotient
topology but the topology induced by our atlas (see Definition 2.8.2). The
coordinate charts are parametrized by the reals: for t ∈ R the set Ut ⊂ M
and the coordinate chart φt : Ut → R2 are given by
Ut := {[t, x, y] | x, y ∈ R} , φt ([t, x, y]) := (x, y).
A subset U ⊂ M is open, by definition, iff φt (U ∩ Ut ) is an open subset of R2
for every t ∈ R. With this topology each φt is a homeomorphism from Ut
onto R2 and M admits a countable dense subset S := {[0, x, y] | x, y ∈ Q}.
However, there is no atlas on M consisting of countably many charts. (Each
coordinate chart can contain at most countably many of the points [t, 0, 0].)
The function f : M → R given by f ([t, x, y]) := t + xy is smooth and each
point [t, 0, 0] is a critical point of f with value t. Thus f has no regular
value. Exercise: Show that M is a path-connected Hausdorff space.
In Theorem 2.9.12 we will show that smooth manifolds whose topology is
Hausdorff and second countable are precisely those that can be embedded in
Euclidean space. Most authors tacitly assume that manifolds are Hausdorff
and second countable and so will we after the end of the present chapter.
However before §2.9.1 there is no need to impose these hypotheses.
92 CHAPTER 2. FOUNDATIONS
fβα := ψβ ◦ f ◦ φ−1
α : φα (Uα ∩ f
−1
(Vβ )) → ψβ (Vβ ) (2.8.3)
The reader may check that the notion of a smooth map is independent
of the atlas used in the definition, that compositions of smooth maps are
smooth, and that sums and products of smooth maps from M to R are
smooth.
A = {(φα , Uα )}α∈A .
where
def −1
(α, x) ∼ (β, y) ⇐⇒ φ−1
α (x) = φβ (y).
eα := {[α, x] | x ∈ φα (Uα )} ,
U φeα ([α, x]) := x.
Prove that M
f is a smooth m-manifold and that it is diffeomorphic to M .
Exercise 2.8.16. Verify the phrase “and hence every” in Definition 2.8.14
and deduce that the map Tp M → TpA M in (2.8.6) is well defined. Show
that it is bijective.
From now on we will use either Definition 2.8.14 or Definition 2.8.15 or
both, whichever way is most convenient, and drop the superscript A .
Definition 2.8.17. For each smooth curve γ : R → M with γ(0) = p we
define the derivative γ̇(0) ∈ Tp M as the equivalence class
∼ d
γ̇(0) := [γ]p = α, φα (γ(t)) ∈ Tp M.
dt t=0 p
φ:U →Ω
M ⊂ Rk
All the theorems we have proved for embedded manifolds and their proofs
carry over almost word for word to the present setting. For example we have
the inverse function theorem, the notion of a regular value, the implicit
function theorem, the notion of an immersion, the notion of an embedding
as a proper injective immersion, and the fact from Theorem 2.3.4 that a
subset P ⊂ M is a submanifold if and only if it is the image of an embedding.
96 CHAPTER 2. FOUNDATIONS
is given by
φeβα (x, ξ) = (φβα (x), dφβα (x)ξ)
for x ∈ φα (Uα ∩ Uβ ) and ξ ∈ Rm where
φβα := φβ ◦ φ−1
α .
Thus the transition maps are all diffeomorphisms and so the coordinate
charts φeα define an atlas on T M . The topology on T M is determined by
this atlas via Definition 2.8.2. If M has a countable atlas, so does T M . The
remaining assertions are easy exercises.
X : M → TM
such that
π ◦ X = id : M → M.
Such a map is also called a section of the tangent bundle.
Now suppose A = {(φα , Uα )}α∈A is an atlas on M and X : M → T M
is a vector field on M . Then X determines a collection of smooth maps
Xα : φα (Uα ) → Rm
given by
Xα (x) := dφα (p)X(p), p := φ−1
α (x), (2.8.9)
for x ∈ φα (Uα ). We can think of each Xα as a vector field on the open set
φα (Uα ) ⊂ Rm , representing the vector field X on the coordinate patch Uα .
These local vector fields Xα satisfy the condition
[X, Y ]α (x) := [Xα , Yα ](x) = dXα (x)Yα (x) − dYα (x)Xα (x) (2.8.12)
[X, Y ]α : φα (Uα ) → Rm
satisfy (2.8.11) and hence determine a unique vector field [X, Y ] on M via
Thus the Lie bracket of X and Y is defined on Uα as the pullback of the Lie
bracket of the vector fields Xα and Yα under the coordinate chart φα . With
this understood all the results in §2.4 about vector fields and flows along with
their proofs carry over word for word to the intrinsic setting whenever M is a
Hausdorff space. This includes the existence and uniquess result for integral
curves in Theorem 2.4.7, the concept of the flow of a vector field in Defini-
tion 2.4.8 and its properties in Theorem 2.4.9, the notion of completeness
of a vector field (that the integral curves exist for all time), and the various
properties of the Lie bracket such as the Jacobi identity (2.4.20), the formu-
las in Lemma 2.4.18, and the fact that the Lie bracket of two vector fields
vanishes if and only if the corresponding flows commute (see Lemma 2.4.26).
One can also carry over the notion of a subbundle E ⊂ T M of rank n to
the intrinsic setting by the condition that E is a smooth submanifold of T M
and intersects each fiber Tp M in an n-dimensional linear subspace
Ep := {v ∈ Tp M | (p, v) ∈ E} .
x1 , . . . , xm : Uα → R.
Thus
∂ 1, if i = j,
dxi (p) (p) = δji :=
∂xj 0, if i 6= j,
for i, j = 1, . . . , m and ∂/∂xi is a vector field on the coordinate patch Uα .
For each p ∈ Uα it is the canonical basis of Tp M determined by φα . In the
notation of (2.8.5) and Remark 2.8.19 we have
∂
(p) = [α, ei ]p = dφα (p)−1 ei ,
∂xi
where ei = (0, . . . , 0, 1, 0, . . . , 0) (with 1 in the ith place) denotes the stan-
dard basis vector of Rm . In other words, for all ξ = (ξ 1 , . . . , ξ m ) ∈ Rm and
all p ∈ Uα , the tangent vector
v := dφα (p)−1 ξ ∈ Tp M
is given by
m
X ∂
v = [α, ξ]p = ξi (p). (2.8.16)
∂xi
i=1
Thus the restriction of a vector field X ∈ Vect(M ) to Uα has the form
m
X ∂
X|Uα = ξi ,
∂xi
i=1
Xα : φα (Uα ) → Rm
Xα ◦ φ−1 1 m
α = (ξ , . . . , ξ ).
2.9.1 Paracompactness
The existence of a countable atlas is of fundamental importance for almost
everything we will prove about manifolds. The next two remarks describe
several equivalent conditions.
U ⊂ 2M
the topology induced by the atlas as in Definition 2.8.2. Then the following
are equivalent.
(a) M admits a countable atlas.
(b) M is σ-compact, i.e. there is a sequence of compact
S∞ subsets Ki ⊂ M
such that Ki ⊂ int(Ki+1 ) for every i ∈ N and M = i=1 Ki .
(c) Every open cover of M has a countable subcover.
(d) M is second countable, i.e. there is a countable collection of open sets
B ⊂ U such that every open set U ∈ U is a union of open sets from the
collection B. (B is then called a countable base for the topology of M .)
That (a) =⇒ (b) =⇒ (c) =⇒ (a) and (a) =⇒ (d) follows directly from the
definitions. The proof that (d) implies (a) requires the construction of a
countable refinement and the axiom of choice. (A refinement of an open
cover {Ui }i∈I is an open cover {Vj }j∈J such that each set Vj is contained in
one of the sets Ui .)
That (a) implies (e) follows from the Urysohn Metrization Theorem
which asserts (in its original form) that every normal second countable topo-
logical space is metrizable [38, Theorem 34.1]. A topological space M is
called normal iff points are closed and, for any two disjoint closed sets
A, B ⊂ M , there are disjoint open sets U, V ⊂ M such that A ⊂ U and
B ⊂ V . It is called regular iff points are closed and, for every closed set
A ⊂ M and every b ∈ M \ A, there are disjoint open sets U, V ⊂ M such
that A ⊂ U and b ∈ V . It is called locally compact if, for every open
set U ⊂ M and every p ∈ U , there is a compact neighborhood of p con-
tained in U . It is easy to show that every manifold is locally compact and
every locally compact Hausdorff space is regular. Tychonoff ’s Lemma
asserts that a regular topological space with a countable base is normal [38,
Theorem 32.1]. Hence it follows from the Urysohn Metrization Theorem
that every Hausdorff manifold with a countable base is metrizable. That (e)
implies (f) follows from a more general theorem which asserts that every
metric space is paracompact (see [38, Theorem 41.4] and [45]). Conversely,
the Smirnov Metrization Theorem asserts that a paracompact Haus-
dorff space is metrizable if and only if it is locally metrizable, i.e. every
point has a metrizable neighborhood (see [38, Theorem 42.1]). Since ev-
ery manifold is locally metrizable this shows that (f) implies (e). Thus we
have (a) =⇒ (e) ⇐⇒ (f) for every Hausdorff manifold.
The proof that (f) implies (a) does not require the Hausdorff property
but we do need the assumption that M is connected. (A manifold with
uncountably many connected components, each of which is paracompact, is
itself paracompact but does not admit a countable atlas.) Here is a sketch.
If M is a paracompact manifold, then there is a locally finite open cover
{Uα }α∈A by coordinate charts. Since each set Uα has a countable dense
subset, the set {α ∈ A | Uα ∩ Uα0 6= ∅} is at most countable for each α0 ∈ A.
Since M is connected we can reach each point from Uα0 through a finite
sequence of sets Uα1 , . . . , Uα` with Uαi−1 ∩ Uαi 6= ∅. This implies that the
index set A is countable and hence M admits a countable atlas.
Exercise 2.9.4. Prove that every manifold is locally compact. Find an ex-
ample of a manifold M and a point p0 ∈ M such that every closed neighbor-
hood of p0 is non-compact. Hint: The example is necessarly non-Hausdorff.
θα : M → [0, 1], α ∈ A,
# {α ∈ A | θα |V 6≡ 0} < ∞, (2.9.1)
If {Uα }α∈A is an open cover of M , then a partition of unity {θα }α∈A (indexed
by the same set A) is called subordinate to the cover iff each θα is
supported in Uα , i.e.
supp(θα ) := {p ∈ M | θα (p) 6= 0} ⊂ Uα .
is an open neighborhood
S of p0 and we have U0 ∩ Vi = ∅ for every i ∈ I0 .
Hence p0 ∈
/ i∈I0 Vi . This proves Lemma 2.9.11.
2.9. CONSEQUENCES OF PARACOMPACTNESS* 105
J → I : j 7→ ij
such that
W j ⊂ V ij ∀ j ∈ J.
Let {χi }i∈I be the partition of unity constructed in Step 3. By the axiom
of choice there is a map I → A : i 7→ αi such that Vi ⊂ Uαi for every i ∈ I.
For α ∈ A define θα : M → [0, 1] by
X
θα := χi .
αi =α
Here the sum runs over all indices i ∈ I with αi = α. This sum is locally
finite and hence is a smooth function on M . Moreover, each point in M has
an open neighborhood in which only finitely many of the θα do not vanish.
Hence the sum of the θα is a well defined function on M and
X X X X
θα = χi = χi ≡ 1.
α∈A α∈A αi =α i∈I
This shows that the functions θα form a partition of unity. To prove the
inclusion supp(θα ) ⊂ Uα we consider the open sets
Wi := {p ∈ M | χi (p) > 0}
K ⊂ Uα1 ∪ · · · ∪ Uα` =: V ⊂ U.
φi := φαi , θi := θαi .
f (p0 ) = f (p1 ),
then
I := i θi (p0 ) > 0 = i θi (p1 ) > 0 6= ∅
and, for i ∈ I, we have θi (p0 ) = θi (p1 ), hence φi (p0 ) = φi (p1 ), and so p0 = p1 .
Moreover, for every p ∈ K the derivative df (p) : Tp M → Rk is injective, and
this proves Step 1.
108 CHAPTER 2. FOUNDATIONS
F0 : A × W0 → R2m+1 , F1 : A × W1 → R2m+1 ,
defined by
this implies
A ∈ A \ (π0 (M0 ) ∪ π1 (M1 )) .
Hence Af : M → R2m+1 is an injective immersion and this proves Step 2.
2.9. CONSEQUENCES OF PARACOMPACTNESS* 109
supp(ρi ) ⊂ Ui , Ki = ρ−1
i (1) ⊂ Ui , Ui ∩ Uj = ∅
S∞
for all i, j ∈ N with |i − j| ≥ 2 and M = i=1 Ki .
Bi := Ci \ Ci−1 for i ∈ N.
S
Then M = i∈N Bi and, for all i, j ∈ N with j ≥ i + 2, we have
Bi ⊂ Ci ⊂ int(Cj−1 ), Bj ⊂ Cj \ int(Cj−1 )
f : M → R4m+4
for p ∈ M .
2.9. CONSEQUENCES OF PARACOMPACTNESS* 111
Now let p0 , p1 ∈ M such that f (p0 ) = f (p1 ). Assume first that p0 ∈ K2i−1 .
odd odd
S
Then ρ (p1 ) = ρ (p0 ) = 1 and hence p1 ∈ j∈N K2j−1 . By (2.9.5), we
also have 2i − 2 < |f odd (p1 )| = |f odd (p0 )| < 2i and hence p1 ∈ K2i−1 . This
implies f2i−1 (p1 ) = f odd (p1 ) = f odd (p0 ) = f2i−1 (p0 ) and so p0 =S p1 . Now
assume p0 ∈ K2i . Then ρev (p1 ) = ρev (p0 ) = 1 and hence p1 ∈ j∈N K2j .
By (2.9.5), we also have 2i − 1 < |f ev (p1 )| = |f ev (p0 )| < 2i + 1, so p1 ∈ K2i ,
which implies f2i (p1 ) = f ev (p1 ) = f ev (p0 ) = f2i (p0 ), and so again p0 = p1 .
This shows that f is injective. That f is an immersion follows from the fact
that the derivative dfi (p) is injective for all p ∈ Ki and all i ∈ N.
We prove that f is proper and has a closed image. Let (pν )ν∈N be a
sequence in M such that the sequence (f (pν ))ν∈N in R4m+4 is bounded.
Choose i ∈S N such that |f odd (pν )| < 2i and |f ev (pν )| < 2i + 1 for all ν ∈ N.
Then pν ∈ 2i j=1 Kj for all ν ∈ N by (2.9.5). Hence (pν )ν∈N has a convergent
subsequence. Thus f : M → R4m+4 is an embedding with a closed image.
Next consider the pair of orthonormal vectors
x := (0, ξ, 0, 0), y := (0, 0, 0, ξ)
in R4m+4 = R × R2m+1 × R × R2m+1 . Let (pν )ν∈N be a sequence in M
that does not have a convergent subsequence and choose a sequence iν ∈ N
such that pν ∈ K2iν −1 ∪ K2iν for all ν ∈ N. Then iν tends to infinity.
If pν ∈ K2iν −1 for all ν, then we have lim supν→∞ |f odd (pν )|−1 |f ev (pν )| ≤ 1
by (2.9.5). Passing to a subsequence, still denoted by (pν )ν∈N , we may
assume that the limit λ := limν→∞ |f odd (fν )|−1 |f ev (pν )| exists. Then
odd
f (pν ) 1 |f ev (pν )| λ
0 ≤ λ ≤ 1, lim =√ , lim =√ ,
ν→∞ |f (pν )| 1 + λ2 ν→∞ |f (pν )| 1 + λ2
and it follows from (2.9.4) that
f odd (pν ) f ev (pν )
lim = ξ, lim = λξ.
ν→∞ |f odd (pν )| ν→∞ |f odd (pν )|
This implies
f (pν ) ξ λξ 1 λ
lim = 0, √ , 0, √ =√ x+ √ y.
ν→∞ |f (pν )| 1+λ 2 1 + λ2 1+λ2 1 + λ2
112 CHAPTER 2. FOUNDATIONS
Similarly, if pν ∈ K2iν for all ν, there exists a subsequence such that the
limit λ := limν→∞ |f ev (pν )|−1 |f odd (pν )| exists and, by (2.9.4), this implies
f (pν ) λξ ξ λ 1
lim = 0, √ , 0, √ =√ x+ √ y.
ν→∞ |f (pν )| 1 + λ2 1 + λ2 1 + λ2 1 + λ2
This shows that the vectors x and y satisfy the requirements of Step 4.
Step 5. There exists an embedding f : M → R2m+1 with a closed image.
For compact manifolds the result was proved in Steps 1 and 2 and for m = 0
the assertion is obvious, because then M is a finite or countable set with
the discrete topology. Thus assume that M is not compact and m ≥ 1.
Choose f : M → R4m+4 and x, y ∈ R4m+4 as in Step 4 and define
( )
the vectors Ax and Ay
A := A ∈ R(2m+1)×(4m+4) .
are linearly independent
This implies
Af (pν )
lim = sAx + tAy 6= 0
ν→∞ |f (pν )|
and hence limν→∞ |Af (pν )| = ∞. Thus the preimage of every compact sub-
set of R2m+1 under the map Af : M → R2m+1 is a compact subset of M ,
and hence Af is proper and has a closed image (Remark 2.3.3).
Now it follows from Step 2 that there exists a matrix A ∈ A such that the
map Af : M → R2m+1 is an injective immersion. Hence it is an embedding
with a closed image. This proves Step 5 and Theorem 2.9.12.
ψ0 := ψ|Ω0 ×G : Ω0 × G → P
Then
p0 g = lim ι(xi )gi = lim ι(x0i )gi0 = p0 g 0
i→∞ i→∞
and so g = because the group action is free. Thus the sequence (gi0 gi−1 )i∈N
g0
in G converges to 1l and hence belongs to the set Ω1 for i sufficiently large.
Since
ψ1 (xi , 1l) = ι(xi ) = ι(x0i )gi0 gi−1 = ψ1 (x0i , gi0 gi−1 )
for all i, this contradicts the injectivity of ψ1 . Thus we have proved that
the map ψ0 : Ω0 × G → P is injective for a suitable neighborhood Ω0 ⊂ Ω
of the origin. That it is an immersion is a direct consequence of the formula
We prove that the intrinsic topology on B agrees with the quotient topol-
ogy. To see this, fix a subset U ⊂ B. Then the following are equivalent.
(a) U is open with respect to the intrinsic topology on B.
(b) φα (U ∩ Uα ) is open in Rn for all α ∈ A.
(c) π −1 (U ∩ Uα ) is open in P for all α ∈ A.
(d) π −1 (U ) is open in P .
(e) U is open with respect to the quotient topology on B.
The equivalence of (a) and (b) follows from the definition of the intrinsic
topology. That (b) implies (c) follows from the three observations that the
set π −1 (Uα ) is open in P , the map φα ◦ π : π −1 (Uα ) → Ωα is continuous,
and (φα ◦ π)−1 (φα (U ∩ Uα )) = π −1 (U ∩ Uα ). That (c) implies (b) follows
from the fact that the map φα ◦ π : π −1 (Uα ) → Ωα is a submersion and hence
maps the open set π −1 (U ∩ Uα ) onto an open subset of Ωα (Corollary 2.6.2).
The equivalence of (c) and (d) follows from the fact that the map π : P → B
is continuous and Uα ⊂ B is open (both with respect to the intrinsic topology
on B) and so π −1 (Uα ) is open in P for all α ∈ A. The equivalence of (d)
and (e) follows from the definition of the quotient topology on B.
Now let ι : Ω → P be a local slice and define the set U := π(ι(Ω)) ⊂ B
and the map φ := (π ◦ ι)−1 : U → Ω. Then the composition
φα ◦ φ−1 = φα ◦ π ◦ ι : φ(U ∩ Uα ) → φα (U ∩ Uα )
Example 2.9.15. There are many important examples of free group actions
and principal bundles. A class of examples arises from orthonormal frame
bundles (§3.4). The complex projective space B = CPn arises from the
action of the circle G = S 1 on the unit sphere P = S 2n+1 ⊂ Cn+1 (Exam-
ple 2.8.5). The real projective space B = RPn arises from the action of the
finite group G = Z/2Z on the unit sphere P = S n ⊂ Rn+1 (Example 2.8.6).
The complex Grassmannian B = Gk (Cn ) arises from the action of G = U(k)
on the space P = Fk (Cn ) of unitary k-frames in Cn (Example 3.7.6). If G is
a Lie group and K ⊂ G is a compact subgroup, then by Theorem 2.9.14 the
homogeneous space G/K admits a unique smooth structure such that the
projection π : G → G/K is a submersion. The example SL(2, C)/SU(2) can
be identified with hyperbolic 3-space (§6.4.4), the example U(n)/O(n) can
be identified with the space of Lagrangian subspaces of a symplectic vector
space [36, Lemma 2.3.2], the example Sp(2n)/U(n) can be identified with
Siegel upper half space or the space of compatible linear complex structures
on a symplectic vector space [36, Lemma 2.5.12], and the example G2 /SO(4)
can be identified with the associative Grassmannian [51, Remark 8.4]. The
last three examples go beyond the scope of the present book.
Standing Assumption
We have seen that all the results in the present chapter carry over to the
intrinsic setting, assuming that the topology of M is Hausdorff and paracom-
pact. In fact, in many cases it is enough to assume the Hausdorff property.
However, these results mainly deal with introducing the basic concepts like
smooth maps, embeddings, submersions, vector fields, flows, and verifying
their elementary properties, i.e. with setting up the language for differen-
tial geometry and topology. When it comes to the substance of the subject
we shall deal with Riemannian metrics and they only exist on paracompact
Hausdorff manifolds. Another central ingredient in differential topology is
the theorem of Sard and that requires second countability. To quote Moe
Hirsch [22]: “Manifolds that are not paracompact are amusing, but they
never occur naturally and it is difficult to prove anything about them.”
Thus we will set the following convention for the remaining chapters.
We assume from now on that each intrinsic manifold M
is Hausdorff and second countable and hence is also paracompact.
For most of this text we will in fact continue to develop the theory for
submanifolds of Euclidean space and indicate, wherever necessary, how to
extend the definitions, theorems, and proofs to the intrinsic setting.
120 CHAPTER 2. FOUNDATIONS
Chapter 3
For a submanifold of Euclidean space the inner product on the ambient space
determines an inner product on each tangent space, the first fundamental
form. The second fundamental form is obtained by differentiating the map
which assigns to each point in M ⊂ Rn the orthogonal projection onto the
tangent space (§3.1). The covariant derivative of a vector field along a curve
is the orthogonal projection of the derivative in the ambient space onto the
tangent space (§3.2). We will show how the covariant derivative gives rise
to parallel transport (§3.3), examine the frame bundle (§3.4), discuss mo-
tions without “sliding, twisting, and wobbling”, and prove the development
theorem (§3.5).
In §3.6 we will see that the covariant derivative is determined by the
Christoffel symbols in local coordinates and thus carries over to the in-
trinsic setting. The intrinsic setting of Riemannian manifolds is explained
in §3.7. The covariant derivative takes the form of a family of linear op-
erators ∇ : Vect(γ) → Vect(γ), one for every smooth curve γ : I → M , and
these operators are uniquely characterized by the axioms of Theorem 3.7.8.
This family of linear operators is the Levi-Civita connection.
121
122 CHAPTER 3. THE LEVI-CIVITA CONNECTION
hv, wi = v1 w1 + v2 w2 + · · · + vn wn
and
Π(p)v = v ⇐⇒ v ∈ Tp M (3.1.3)
for p ∈ M and v ∈ Rn (see Exercise 2.6.9).
Lemma 3.1.2. The map Π : M → Rn×n defined by (3.1.2) and (3.1.3) is
smooth.
Proof. This follows directly from Theorem 2.6.10 and Corollary 2.6.12. More
explicitly, if U ⊂ M is an open set and φ : U → Ω is a coordinate chart onto
an open subset Ω ⊂ Rm with the inverse ψ := φ−1 : Ω → U , then
−1
Π(p) = dψ(φ(p)) dψ(φ(p))T dψ(φ(p)) dψ(φ(p))T
ν( p)
TpM
ν : U → Rm+1
satisfying
ν(p) ⊥ Tp M, |ν(p)| = 1 (3.1.4)
for all p ∈ U (see Figure 3.1). Such a map ν is called a Gauß map. The
function Π : M → Rn×n is in this case given by
for p ∈ U .
1 − x2
−xy −xz
Π(p) = 1l − ppT = −yx 1 − y 2 −yz
−zx −zy 1 − z 2
for p = (x, y, z) ∈ S 2 .
for ε > 0 sufficiently small. Show that there does not exist a global smooth
function ν : M → R3 satisfying (3.1.4).
124 CHAPTER 3. THE LEVI-CIVITA CONNECTION
R2 → M : (s, t) 7→ γ(s, t)
satisfying
∂γ ∂γ
γ(0, 0) = p, (0, 0) = v, (0, 0) = w,
∂s ∂t
(for example by doing this in local coordinates) and denote
∂γ ∂γ
X(s, t) := (s, t) ∈ Tγ(s,t) M, Y (s, t) := (s, t) ∈ Tγ(s,t) M.
∂s ∂t
Then
∂Y ∂2γ ∂X
= =
∂s ∂s∂t ∂t
and hence, using (3.1.7), we obtain
∂γ ∂γ ∂γ
dΠ(γ) = dΠ(γ) X
∂t ∂s ∂t
∂X
= 1l − Π(γ)
∂t
∂Y
= 1l − Π(γ)
∂s
∂γ
= dΠ(γ) Y
∂s
∂γ ∂γ
= dΠ(γ) .
∂s ∂t
With s = t = 0 we obtain
dΠ(p)w v = dΠ(p)v w ∈ Tp M ⊥
hp : Tp M × Tp M → Tp M ⊥ ,
defined by
hp (v, w) := (dΠ(p)v)w = (dΠ(p)w)v (3.1.8)
for p ∈ M and v, w ∈ Tp M is called the second fundamental form on M .
126 CHAPTER 3. THE LEVI-CIVITA CONNECTION
hp (v, w) = −ν(p)hdν(p)v, wi
for p ∈ M and v, w ∈ Tp M .
Thus f (0) = 0 and df (0) = 0. Prove that the second fundamental form
hp : Tp M × Tp M → Tp M ⊥ is given by the second derivatives of f , i.e.
m
X ∂2f
hp (v, w) = 0, (0)vi wj
∂xi ∂xj
i,j=1
for v, w ∈ Tp M = Rm × {0}.
L := {p + tv + w | t ∈ R, w ⊥ Tp M } .
X(t)
γ(t)
M
It follows directly from the definition that the covariant derivative along
a curve γ : I → M is a linear operator ∇ : Vect(γ) → Vect(γ). The following
lemma summarizes the basic properties of this operator.
Lemma 3.2.4 (Covariant Derivative). The covariant derivative satisfies
the following axioms for any two open intervals I, J ⊂ R.
(i) Let γ : I → M be a smooth curve, let λ : I → R be a smooth function,
and let X ∈ Vect(γ). Then
∇(λX) = λ̇X + λ∇X. (3.2.3)
∇X = 0 ⇐⇒ Ẋ = hγ (γ̇, X).
The next theorem shows that any given tangent vector v0 ∈ Tγ(t0 ) M extends
uniquely to a parallel vector field along γ.
Theorem 3.3.4 (Existence and Uniqueness). Let I ⊂ R be an interval
and γ : I → M be a smooth curve. Let t0 ∈ I and v0 ∈ Tγ(t0 ) M be given.
Then there is a unique parallel vector field X ∈ Vect(γ) such that X(t0 ) = v0 .
Proof. Choose a basis e1 , . . . , em of the tangent space Tγ(t0 ) M and let
X1 , . . . , Xm ∈ Vect(γ)
Xi (t0 ) = ei , i = 1, . . . , m.
(For example choose Xi (t) := Π(γ(t))ei .) Then the vectors Xi (t0 ) are lin-
early independent. Since linear independence is an open condition there is a
130 CHAPTER 3. THE LEVI-CIVITA CONNECTION
constant ε > 0 such that the vectors X1 (t), . . . , Xm (t) ∈ Tγ(t) M are linearly
independent for every t ∈ I0 := (t0 − ε, t0 + ε) ∩ I. Since Tγ(t) M is an
m-dimensional real vector space this implies that the vectors Xi (t) form a
basis of Tγ(t) M for every t ∈ I0 . We express the vector ∇Xi (t) ∈ Tγ(t) M in
this basis and denote the coefficients by aki (t) so that
m
X
∇Xi (t) = aki (t)Xk (t).
k=1
˙ + A(t)ξ(t) = 0,
ξ(t) A(t) := .. ..
.
. .
am
1 (t) · · ·
m
am (t)
Thus we have translated the equation ∇X = 0 over the interval I0 into a time
dependent linear ordinary differential equation. By a theorem in Analysis II
(see [47, Lemma 4.4.3]), this equation has a unique solution for any initial
condition at any point in I0 . Thus we have proved that every t0 ∈ I is
3.3. PARALLEL TRANSPORT 131
for the pullback tangent bundle. This set is a smooth submanifold of I ×Rn .
(See Theorem 2.6.10 and Corollary 2.6.13.) The next theorem summarizes
the properties of parallel transport. In particular, the last assertion shows
that the covariant derivative can be recovered from the parallel transport
maps.
132 CHAPTER 3. THE LEVI-CIVITA CONNECTION
for all s, t ∈ J.
(v) The map
is smooth.
(vi) For all X ∈ Vect(γ) and t, t0 ∈ I we have
d
Φγ (t0 , t)X(t) = Φγ (t0 , t)∇X(t).
dt
Proof. Assertion (i) holds because the sum of two parallel vector fields
along γ is again parallel and the product of a parallel vector field with a
constant real number is again parallel. Assertion (ii) follows directly from
the uniqueness statement in Theorem 3.3.4.
We prove (iii). Fix a number s ∈ I and two tangent vectors
v, w ∈ Tγ(s) M.
These vector fields are parallel. Thus, by equation (3.2.5) in Lemma 3.2.4,
we have
d
hX, Y i = h∇X, Y i + hX, ∇Y i = 0.
dt
Hence the function I → R : t 7→ hX(t), Y (t)i is constant and this proves (iii).
3.3. PARALLEL TRANSPORT 133
for t ∈ I. Thus X is the unique parallel vector field along γ that satisfies
X(α(s)) = v.
Denote
e := γ ◦ σ : J → M,
γ e := X ◦ σ : I → Rn
X
Then X
e is a vector field along γ
e and, by the chain rule, we have
d e d
X(t) = X(σ(t)) = σ̇(t)Ẋ(σ(t)).
dt dt
Projecting orthogonally onto the tangent space Tγ(σ(t)) M we obtain
∇X(t)
e = σ̇(t)∇X(σ(t)) = 0
Since each vector field Xi is parallel it satisfies Xi (t) = Φγ (t, s)Xi (s). Hence
m
X
Φγ (t, s)v = hXi (s), viXi (t) (3.3.2)
i=1
Evidently, the derivative of the first sum with respect to t is equal to the
second sum. This proves (vi) and the theorem.
as a real n × n matrix. The formula (3.3.2) in the proof of (v) shows that
this matrix can be expressed in the form
m
X
Φγ (t, s)Π(γ(s)) = Xi (t)Xi (s)T ∈ Rn×n .
i=1
The right hand side defines a smooth matrix valued function on I × I and
this is equivalent to the assertion in (v).
Remark 3.3.8. It follows from assertions (ii) and (iii) in Theorem 3.3.6
that
Φγ (t, s)−1 = Φγ (s, t) = Φγ (t, s)∗
for all s, t ∈ I. Here the linear map Φγ (t, s)∗ : Tγ(t) M → Tγ(s) M is under-
stood as the adjoint operator of Φγ (t, s) : Tγ(s) M → Tγ(t) M with respect to
the inner products on the two subspaces of Rn inherited from the Euclidean
inner product on the ambient space.
3.3. PARALLEL TRANSPORT 135
The two theorems in this section carry over verbatim to any smooth
vector bundle E ⊂ M × Rn over a manifold. As in the case of the tangent
bundle one can define the covariant derivative of a section of E along γ as
the orthogonal projection of the ordinary derivative in the ambient space Rn
onto the fiber Eγ(t) . Instead of parallel vector fields one then speaks about
horizontal sections and one proves as in Theorem 3.3.4 that there is a unique
horizontal section along γ through any point in any of the fibers Eγ(t0 ) . This
gives parallel transport maps from Eγ(s) to Eγ(t) for any pair s, t ∈ I and
Theorem 3.3.6 carries over verbatim to all vector bundles E ⊂ M × Rn . We
spell this out in more detail in the case where E = T M ⊥ ⊂ M × Rn is the
normal bundle of M .
Let γ : I → M be a smooth curve. A normal vector field along γ is
a smooth map Y : I → Rn such that Y (t) ⊥ Tγ(t) M for every t ∈ I. The set
of normal vector fields along γ will be denoted by
hp (u)∗ w = dΠ(p)u w, w ∈ Tp M ⊥ .
(3.3.5)
Π(γ(t))Y (t) = 0.
and hence
Ẏ (t) = Ẏ (t) − Π(γ(t))Ẏ (t) − dΠ(γ(t))γ̇(t) Y (t)
= ∇⊥ Y (t) − hγ(t) (γ̇(t))∗ Y (t)
for t ∈ I. Here the last equation follows from (i) and the definition of ∇⊥ .
This proves Lemma 3.3.9.
Theorem 3.3.4 and its proof carry over to the normal bundle T M ⊥ .
Thus, if γ : I → M is a smooth curve, then for all s ∈ I and w ∈ Tγ(s) M ⊥
there is a unique normal vector field Y ∈ Vect⊥ (γ) such that
∇⊥ Y ≡ 0, Y (s) = w.
Φ⊥ ⊥
γ (t, s) : Tγ(s) M → Tγ(t) M
⊥
defined by
Φ⊥
γ (t, s)w := Y (t)
for s, t ∈ I and w ∈ Tγ(s) M ⊥ , where Y is the unique normal vector field along
γ satisfying ∇⊥ Y ≡ 0 and Y (s) = w. These parallel transport maps satisfy
exactly the same conditions that have been spelled out in Theorem 3.3.6
for the tangent bundle and the proof carries over verbatim to the present
setting.
3.4. THE FRAME BUNDLE 137
a∗ b∗ e = (ba)∗ e, 1l∗ e = e
for a, b ∈ GL(m) and e ∈ Liso (Rm , V ). Moreover, the action is free, i.e. for
all a ∈ GL(m) and e ∈ Liso (Rm , V ), we have
a∗ e = e ⇐⇒ a = 1l.
F(M ) ⊂ Rn × Rn×m
π : F(M ) → M
for (p, e) ∈ F(M ), and the group GL(m) acts freely and transitively on each
of these fibers.
3.4. THE FRAME BUNDLE 139
G × P → P : (g, p) 7→ pg
by a Lie group G such that π(pg) = π(p) for all p ∈ P and g ∈ G and such
that the group G acts freely and transitively on the fiber Pb := π −1 (b) for
each b ∈ B. In this book we shall mostly be concerned with the frame bundle
of a manifold M and the orthonormal frame bundle.
e1 , . . . , em is an
eT e = 1l ⇐⇒ hei , ej i = δij ⇐⇒
orthonormal basis.
Exercise 3.4.4. Prove that O(M ) is a submanifold of F(M ) and that the
obvious projection π : O(M ) → M is a submersion. Prove that the action
of GL(m) on F(M ) restricts to an action of the orthogonal group O(m)
on O(M ) whose orbits are the fibers
= e ∈ Liso (Rm , Tp M ) eT e = 1l .
The next lemma gives an explicit formula for this tangent space in terms of
the second fundamental form hp : Tp M × Tp M → Tp M ⊥ in Definition 3.1.9.
Compare this formula with Lemma 4.3.1 in the next chapter.
Lemma 3.4.5. Let M ⊂ Rn be a smooth m-dimensional submanifold. Then
the tangent space of F(M ) at (p, e) is given by
( )
pb ∈ Tp M, eb ∈ Rn×m , and
T(p,e) F(M ) = (b
p, eb) . (3.4.3)
1l − Π(p) eb = hp (bp)e
Proof. We prove the inclusion “⊂” in (3.4.3). Let (b p, eb) ∈ T(p,e) F(M ) and
choose a smooth curve R → F(M ) : t 7→ (γ(t), e(t)) such that
Fix a vector ξ ∈ M and define the vector field X ∈ Vect(γ) by X(t) := e(t)ξ
for t ∈ R. Then the Gauß–Weingarten formula (3.2.2) asserts that
ė(t)ξ = Ẋ(t)
= ∇X(t) + hγ(t) (γ̇(t), X(t))
= Π(γ(t))ė(t)ξ + hγ(t) (γ̇(t), e(t)ξ)
(1l − Π(p))b
eξ = hp (b
p, eξ) = hp (b
p)eξ
for all ξ ∈ Rm . This proves the inclusion “⊂” in (3.4.3). Equality holds
because both sides of the equation are (m+m2 )-dimensional linear subspaces
of Rn × Rn×m . This proves Lemma 3.4.5.
π ◦ β = γ.
Any such lift has the form β(t) = (γ(t), e(t)) with e(t) ∈ Liso (Rm , Tγ(t) M ).
The associated curve of frames e(t) of the tangent spaces Tγ(t) M is called a
moving frame along γ. A curve
X(t) := e(t)ξ
along γ is parallel for every ξ ∈ Rm . Thus a horizontal lift of γ has the form
Lemma 3.4.7. (i) The tangent space of F(M ) at (p, e) ∈ F(M ) is the direct
sum
T(p,e) F(M ) = H(p,e) ⊕ V(p,e)
(ii) The vertical space V(p,e) at (p, e) ∈ F(M ) is the kernel of the linear map
and so H(p,e) ⊂ T(p,e) F(M ). Moreover, β̇(0) = (v, hp (v)e) ∈ H(p,e) = Hβ(0)
and this proves Step 2.
144 CHAPTER 3. THE LEVI-CIVITA CONNECTION
−1
F(M)p = π (p)
F(M)
(p,e)
M p
The reason for the terminology introduced in Definition 3.4.6 is that one
draws the extremely crude picture of the frame bundle displayed in Fig-
ure 3.3. One thinks of F(M ) as “lying over” M . One would then represent
the equation γ = π ◦ β by the following commutative diagram:
F(M ) ;
8
β
ppppp
p
ppp
π
ppp γ /M
R
hence the word “lift”. The vertical space is tangent to the vertical line in
Figure 3.3 while the horizontal space is transverse to the vertical space. This
crude imagery can be extremely helpful.
3.4. THE FRAME BUNDLE 145
3.5.1 Motion
Definition 3.5.1. A motion of M along M 0 (on an interval I ⊂ R) is
a triple (Ψ, γ, γ 0 ) of smooth maps
Ψ : I → O(n), γ : I → M, γ0 : I → M 0
such that
Ψ(t)Tγ(t) M = Tγ 0 (t) M 0 ∀ t ∈ I.
Note that a motion also matches normal vectors, i.e.
⊥
Ψ(t)Tγ(t) M ⊥ = Tγ 0 (t) M 0 ∀ t ∈ I.
Remark 3.5.2. Associated to a motion (Ψ, γ, γ 0 ) of M along M 0 is a family
of (affine) isometries ψt : Rn → Rn defined by
ψt (p) := γ 0 (t) + Ψ(t) p − γ(t)
(3.5.1)
for t ∈ I and p ∈ Rn . These isometries satisfy
ψt (γ(t)) = γ 0 (t), dψt (γ(t))Tγ(t) M = Tγ 0 (t) M 0 ∀ t ∈ I.
3.5. MOTIONS AND DEVELOPMENTS 147
(Ψ−1 , γ 0 , γ)
is a motion of M 0 along M .
Composition. If (Ψ, γ, γ 0 ) is a motion of M along M 0 on an interval I
and (Ψ0 , γ 0 , γ 00 ) is a motion of M 0 along M 00 on the same interval, then
(Ψ0 Ψ, γ, γ 00 )
is a motion of M along M 00 .
We now give the three simplest examples of “bad” motions; i.e. motions
which do not satisfy the concepts we are about to define. In all three of
these examples, p is a point of M and M 0 is the affine tangent space to M
at p:
M 0 := p + Tp M = {p + v | v ∈ Tp M } .
Ψ(t)γ̇(t) 6= γ̇ 0 (t).
M’
p
γ(t) = γ 0 (t) = p
and take Ψ(t) to be the identity on Tp M ⊥ and any curve of rotations on the
tangent space Tp M . As a concrete example with m = 2 and n = 3 one can
take M to be the sphere of radius one centered at the point (0, 1, 0) and p
to be the origin:
M := (x, y, z) ∈ R3 | x2 + (y − 1)2 + z 2 = 1 ,
p := (0, 0, 0).
Then M 0 is the (x, z)-plane and A(t) is any curve of rotations in the (x, z)-
plane, i.e. about the y-axis Tp M ⊥ . (See Figure 3.5.)
M’
p
M := (x, y, 0) ∈ R3 | x2 + (y − 1)2 = 1 ,
p := (0, 0, 0).
Then M 0 is the x-axis and Ψ(t) is any curve of rotations in the (y, z)-plane,
i.e. about the axis M 0 . (See Figure 3.6.)
M’ p
3.5.2 Sliding
When a train slides on the track (e.g. in the process of stopping suddenly),
there is a terrific screech. Since we usually do not hear a screech, this means
that the wheel moves along without sliding. In other words the velocity of
the point of contact in the train wheel M equals the velocity of the point
of contact in the track M 0 . But the track is not moving; hence the point of
contact in the wheel is not moving. One may explain the paradox this way:
the train is moving forward and the wheel is rotating around the axle. The
velocity of a point on the wheel is the sum of these two velocities. When
the point is on the bottom of the wheel, the two velocities cancel.
Definition 3.5.7. A motion (Ψ, γ, γ 0 ) is said to be without sliding iff it
satisfies Ψ(t)γ̇(t) = γ̇ 0 (t) for every t.
Here is the geometric picture of the no sliding condition. As explained
in Remark 3.5.2 we can view a motion as a smooth family of isometries
ψt (p) := γ 0 (t) + Ψ(t) p − γ(t)
acting on the manifold M with γ(t) ∈ M being the point of contact with M 0 .
Differentiating the curve t 7→ ψt (p) which describes the motion of the point
p ∈ M in the space Rn we obtain
d
ψt (p) = γ̇ 0 (t) − Ψ(t)γ̇(t) + Ψ̇(t) p − γ(t) .
dt
Taking p = γ(t0 ) we find
d
ψt (γ(t0 )) = γ̇ 0 (t0 ) − Ψ(t0 )γ̇(t0 ).
dt
t=t0
This expression vanishes under the no sliding condition. In general the
curve t 7→ ψt (γ(t0 )) will be non-constant, but (when the motion is without
sliding) its velocity will vanish at the instant t = t0 ; i.e. at the instant when
it becomes the point of contact. In other words the motion is without sliding
if and only if the point of contact is motionless.
We remark that, if the motion is without sliding, we have:
0
γ̇ (t) = |Ψ(t)γ̇(t)| = |γ̇(t)|
Exercise 3.5.8. Give an example of a motion where |γ̇(t)| = |γ̇ 0 (t)| for
every t but which is not without sliding.
M := (x, y) ∈ R2 | x2 + (y − 1)2 = 1 .
Choose
γ(t) := (cos(t − π/2), 1 + sin(t − π/2))
= (sin(t), 1 − cos(t)),
0
γ (t) := (t, 0),
The reader can easily verify that this is a motion without sliding. A fixed
point p0 on M , say p0 = (0, 0), sweeps out a cycloid with parametric equa-
tions
x = t − sin(t), y = 1 − cos(t).
(Check that (ẋ, ẏ) = (0, 0) when y = 0; i.e. for t = 2nπ.)
x2 + (y − 1)2 + z 2 = 1,
the plane is y = 0 and the line is the x-axis. The z-coordinate of a point is
unaffected by the motion. Note that the curve γ 0 traces out a straight line
in the plane M 0 and the curve γ traces out a great circle on the sphere M .
∇0 (ΨX) = Ψ∇X.
(iii) Ψ transforms parallel vector fields along γ into parallel vector fields
along γ 0 , i.e. for X ∈ Vect(γ)
∇X = 0 =⇒ ∇0 (ΨX) = 0.
Ψ(t)Π(γ(t)) = Π0 (γ 0 (t))Ψ(t)
for every t ∈ I. This restates the condition that Ψ(t) maps tangent vectors
of M to tangent vectors of M 0 and normal vectors of M to normal vectors
of M 0 . Differentiating the equation X 0 (t) = Ψ(t)X(t) we obtain
∇0 X 0 = Ψ∇X + Π0 (γ 0 )Ψ̇X.
We prove that (iii) implies (iv). Let t0 ∈ I and v0 ∈ Tγ(t0 ) M . Define the
vector field X ∈ Vect(γ) by X(t) := Φγ (t, t0 )v0 for t ∈ I and let X 0 := ΨX.
Then ∇X = 0, hence ∇0 X 0 = 0 by (iii), and hence
for all t ∈ I. Since X 0 (t) = Ψ(t)X(t) = Ψ(t)Φγ (t, t0 )v0 , this implies (iv).
We prove that (iv) implies (ii). Let X ∈ Vect(γ) and X 0 := ΨX. By (iv)
we have
Φ0γ 0 (t0 , t)X 0 (t) = Ψ(t0 )Φγ (t0 , t)X(t).
Differentiating this equation with respect to t at t = t0 and using Theo-
rem 3.3.6, we obtain ∇0 X 0 (t0 ) = Ψ(t0 )∇X(t0 ). This proves the lemma.
Ψ̇(t)Tγ(t) M ⊥ ⊂ Tγ 0 (t) M 0 .
(iii) Ψ transforms parallel normal vector fields along γ into parallel normal
vector fields along γ 0 , i.e. for Y ∈ Vect⊥ (γ)
⊥
∇⊥ Y = 0 =⇒ ∇0 (ΨY ) = 0.
3.5.4 Development
A development is an intrinsic version of motion without sliding or twisting.
Φ0 : Tp0 M → Tγ 0 (t0 ) M 0
(ii) Any two developments (Φ1 , γ1 , γ 0 |I1 ) and (Φ2 , γ2 , γ 0 |I2 ) as in (i) on two
intervals I1 and I2 agree on the intersection I1 ∩ I2 , i.e.
for every t ∈ I1 ∩ I2 .
(iii) If M is complete, then (i) holds with I = R.
γ(t0 ) = p0
by
Φ(t) := Φ0γ 0 (t, t0 )Φ0 Φγ (t0 , t). (3.5.8)
This is an orthogonal transformation for every t and it intertwines parallel
transport. However, in general Φ(t)γ̇(t) will not be equal to γ̇ 0 (t).
156 CHAPTER 3. THE LEVI-CIVITA CONNECTION
R → Rn×m : t 7→ e(t)
is smooth. In fact, the map t 7→ (γ(t), e(t)) is a smooth path in the frame
bundle F(M ). Define the smooth map ξ : R → Rm by
for every t ∈ I, where Bξ(t) ∈ Vect(F(M )) denotes the basic vector field
associated to ξ(t) ∈ Rm (see equation (3.4.9)).
The triple (Φ, γ, γ 0 ) is a development on I if and only if
Φ(t)γ̇(t) = γ̇ 0 (t)
Φ0γ 0 (t, t0 )Φ0 Φγ (t0 , t)γ̇(t) = γ̇ 0 (t) = Φ0γ 0 (t, t0 )Φ0 e0 ξ(t),
hence to
Φγ (t0 , t)γ̇(t) = e0 ξ(t),
and hence to
γ̇(t) = Φγ (t, t0 )e0 ξ(t) = e(t)ξ(t) (3.5.12)
for every t ∈ I. By (3.5.9) and the Gauß–Weingarten formula, we have
Remark 3.5.22. As any two developments (Φ1 , γ1 , γ 0 |I1 ) and (Φ2 , γ2 , γ 0 |I2 )
on two intervals I1 and I2 that satisfy the initial condition (3.5.7) agree
on I1 ∩ I2 there is a development defined on I1 ∪ I2 . Hence there is a
unique maximally defined development (Φ, γ, γ 0 |I ), defined on a maximal
interval I = I(t0 , p0 , Φ0 ), associated to the initial data t0 , p0 , Φ0 .
Denote the space of initial data by
t ∈ R, p ∈ M, Φ : Tp M → Tγ 0 (t) M 0
P := (t, p, Φ) , (3.5.13)
is an orthogonal transformation
and let
D → P : (t, t0 , p0 , Φ0 ) 7→ (t, γ(t), Φ(t)), (3.5.15)
be the map which assigns to each (t0 , p0 , Φ0 ) ∈ P and each t ∈ I(t0 , p0 , Φ0 )
the value at time t of the unique development (Φ, γ, γ 0 |I ) associated to the
inital condition (t0 , p0 , Φ0 ) on the maximal time interval I = I(t0 , p0 , Φ0 ).
Then the space P has a natural structure of a smooth manifold (in the
intrinsic setting), and it follows from Theorem 2.4.9 and the proof of Theo-
rem 3.5.21 that D is an open subset of R × P and the map (3.5.5) is smooth.
The smooth structure on P can be understood as follows. The space
Ψ0 Tp0 M = Tγ 0 (t0 ) M 0 .
Proof. We have already noted that the basic vector fields are all tangent to
the orthonormal frame bundle O(M ) ⊂ F(M ). Now note that if a smooth
curve I → F(M ) : t 7→ β(t) = (γ(t), e(t)) on an interval I ⊂ R satisfies the
differential equation
β̇(t) = Bξ(t) (β(t))
for all t, then so does the curve
for every a ∈ GL(m, R). Since any frame e0 : Rm → Tp0 M can be car-
ried to any other (in particular an orthonormal one) by a suitable ma-
trix a ∈ GL(m, R), this shows that (i) is equivalent to (ii).
That (i) implies (iii) was proved in Theorem 3.5.21.
3.5. MOTIONS AND DEVELOPMENTS 159
Φ0 := e−1
0 : Tp0 M → R
m
and Z t
0
γ (t) := ξ(s) ds ∈ Rm for t ∈ R.
0
By (ii) there exists a development (Φ, γ, γ 0 ) of M along Rm on all of R that
satisfies the initial conditions
γ(0) = p0 , Φ(0) = Φ0 .
Then
for t ∈ R. Then (γ, e) : R → F(M ) is a smooth curve that satisfies the initial
condition (γ(0), e(0)) = (p0 , e0 ) and the differential equation
φ:U →Ω
ψ := φ−1 : Ω → U ⊂ M.
x = (x1 , . . . , xm ) ∈ Ω.
X(t)
γ(t)
U
M ψ
φ
Ω
ξ(t)
c(t)
γ =ψ◦c:I →M
(see Figure 3.7). Our goal is to describe the operator X 7→ ∇X on the space
of vector fields along γ in local coordinates. Let X : I → Rn be a vector
field along γ. Then
ξ = (ξ 1 , . . . , ξ m ) : I → Rm
such that
m
X ∂ψ
X(t) = dψ(c(t))ξ(t) = ξ i (t) (c(t)). (3.6.4)
∂xi
i=1
We examine the projection ∇X(t) = Π(γ(t))Ẋ(t) of this vector onto the tan-
gent space of M at γ(t). The first summand on the right in (3.6.5) is already
162 CHAPTER 3. THE LEVI-CIVITA CONNECTION
tangent to M . For the second summand we simply observe that the vec-
tor Π(ψ(x))∂ 2 ψ/∂xi ∂xj lies in tangent space Tψ(x) M and can therefore be
expressed as a linear combination of the basis vectors ∂ψ/∂x1 , . . . , ∂ψ/∂xm .
The coefficients will be denoted by Γkij (x). Thus there exist smooth func-
tions Γkij : Ω → R for i, j, k = 1, . . . , m such that
m
∂2ψ X ∂ψ
Π(ψ(x)) i j
(x) = Γkij (x) k (x) (3.6.6)
∂x ∂x ∂x
k=1
for all x ∈ Ω and all i, j ∈ {1, . . . , m}. The coefficients Γkij : Ω → R are called
the Christoffel symbols associated to the coordinate chart φ : U → Ω. To
sum up we have proved the following.
γ := ψ ◦ c : I → M.
Our next goal is to understand how the Christoffel symbols are deter-
mined by the metric in local coordinates. Recall from equation (3.6.2) that
the inner products on the tangent spaces inherited from the standard Eu-
clidean inner product on the ambient space Rn are in local coordinates
represented by the matrix valued function
g = (gij )m
i,j=1 : Ω → R
m×m
given by
∂ψ ∂ψ
gij := , . (3.6.8)
∂xi ∂xj Rn
We shall see that the Christoffel symbols are completely determined by the
functions gij : Ω → R. Here are first some elementary observations.
3.6. CHRISTOFFEL SYMBOLS 163
Remark 3.6.2. The matrix g(x) ∈ Rm×m is symmetric and positive definite
for every x ∈ Ω. This follows from the fact that the matrix dψ(x) ∈ Rn×m
has rank m and the matrix g(x) is given by
g(x) = dψ(x)T dψ(x)
Thus ξ T g(x)ξ = |dψ(x)ξ|2 > 0 for all ξ ∈ Rm \ {0}.
Remark 3.6.3. For x ∈ Ω we have det(g(x)) > 0 by Remark 3.6.2 and
so the matrix g(x) is invertible. Denote the entries of the inverse ma-
trix g(x)−1 ∈ Rm×m by g k` (x). They are determined by the condition
m
X
jk k 1, if i = k,
gij (x)g (x) = δi =
0, if i 6= k.
j=1
Since g(x) is symmetric and positive definite, so is its inverse matrix g(x)−1 .
In particular, we have g k` (x) = g `k (x) for all x ∈ Ω and all k, ` ∈ {1, . . . , m}.
Remark 3.6.4. Suppose that X, Y ∈ Vect(γ) are vector fields along our
curve γ = ψ ◦ c : I → M and ξ, η : I → Rm are defined by
m m
X
i ∂ψ X ∂ψ
X(t) = ξ (t) i (c(t)), Y (t) = η j (t) (c(t)).
∂x ∂xj
i=1 j=1
If the Γkij are defined by (3.6.6) and the gij by (3.6.8), then the Γkij sat-
isfy (3.6.9) and hence are given by (3.6.10).
164 CHAPTER 3. THE LEVI-CIVITA CONNECTION
Proof. Suppose that the Γkij are given by (3.6.6) and the gij by (3.6.8). Let
c : I → Ω, ξ, η : I → Rm
be smooth functions and suppose that the vector fields X, Y along the curve
γ := ψ ◦ c : I → M
are given by
m m
X ∂ψ X ∂ψ
X(t) := ξ i (t) (c(t)), Y (t) := η j (t) (c(t)).
∂xi ∂xj
i=1 j=1
Dropping the argument t in each term, we obtain from Remark 3.6.4 and
Lemma 3.6.1 that
X
hX, Y i = gij (c)ξ i η j ,
i,j
X X
hX, ∇Y i = gik (c)ξ i η̇ k + Γkj` (c)η j ċ` ,
i,k j,`
X X
h∇X, Y i = gkj (c) ξ˙k + Γki` (c)ξ i ċ` η j .
j,k i,`
i,k i,j,k,`
X X
− gkj (c)ξ˙k η j − gkj (c)Γki` (c)ξ i η j ċ`
j,k i,j,k,`
!
X ∂gij X X
= (c) − gik (c)Γkj` (c) − gjk (c)Γki` (c) ξ i η j ċ` .
∂x`
i,j,` k k
This holds for all smooth maps c : I → Ω and ξ, η : I → Rm , so the Γkij satisfy
the second equation in (3.6.9). That they are symmetric in i and j is obvious.
3.6. CHRISTOFFEL SYMBOLS 165
∂gij
Γ`ij = Γ`ji , = Γij` + Γji` . (3.6.12)
∂x`
Γ`ij = Γ`ji
and
∂gij
= Γij` + Γji` ,
∂x`
∂g`i
= Γ`ij + Γi`j = Γ`ij + Γij` ,
∂xj
∂g`j
= Γ`ji + Γj`i = Γ`ij + Γji` .
∂xi
Take the sum of the last two minus the first of these equations to obtain
Each matrix gα (x) is symmetrix and positive definite. Note that the
tangent vectors v and w in (3.7.2) can also be written in the form
ξ = ei , η = ej
in Rm we obtain
∂
[α, ei ]p = dφα (p)−1 ei =: (p)
∂xi
and hence
∂ −1 ∂ −1
gα,ij (x) = (φ (x)), j (φα (x)) . (3.7.3)
∂xi α ∂x
For different coordinate charts the maps gα and gβ are related through the
transition map
φβα := φβ ◦ φ−1
α : φα (Uα ∩ Uβ ) → φβ (Uα ∩ Uβ )
via
gα (x) = dφβα (x)T gβ (φβα (x))dφβα (x) (3.7.4)
for α, β ∈ A.
gα : φα (Uα ) → Rm×m
with values in the set of positive definite symmetric matrices that satis-
fies (3.7.4) for all α, β ∈ A determines a global Riemannian metric via (3.7.2).
where the sum runs over all α ∈ A with p ∈ Uα and the inner product is
the standard inner product on Rm . Since supp(θα ) ⊂ Uα for each α and the
sum is locally finite we find that the function
M → R : p 7→ hX(p), Y (p)ip
is smooth for every pair of vector fields X, Y ∈ Vect(M ). Moreover, the right
hand side of (3.7.5) is symmetric in v and w and is positive for v = w 6= 0
because each summand is nonnegative and each summand with θα (p) > 0 is
positive. Thus equation (3.7.5) defines a Riemannian metric on M .
The second method is to define the functions
gα : φα (Uα ) → Rm×m
by
X
gα (x) := θγ (φ−1 T
α (x))dφγα (x) dφγα (x) (3.7.6)
γ∈A
Define the inner product on this tangent space to be the restriction of the
standard inner product on Cn×k to this subspace. Exercise: Prove that
the unitary group U(n) acts on Gk (Cn ) by isometries.
170 CHAPTER 3. THE LEVI-CIVITA CONNECTION
A = {(φα , Uα )}α∈A
gα = (gα,ij )m
i,j=1 : φα (Uα ) → R
m×m
X(q) ∈ Tf (q) M,
N → T M : q 7→ (f (q), X(q))
∇ : Vect(γ) → Vect(γ)
d
hX, Y i = h∇X, Y i + hX, ∇Y i. (3.7.9)
dt
∇s ∂t γ = ∇t ∂s γ. (3.7.10)
(see Lemma 3.6.5). Prove that they are related by the equation
X ∂φkβα
0
∂ 2 φkβα
0
X 0
∂φi0 ∂φj 0
βα βα
Γk = + Γkβ,i0 j 0 ◦ φβα .
∂xk α,ij ∂xi ∂xj ∂xi ∂xj
k i0 ,j 0
for all α, β ∈ A.
Exercise 3.7.10. Denote ψα := φ−1α : φα (Uα ) → M . Prove that the covari-
ant derivative of a vector field
m
X ∂ψα
X(t) = ξαi (t) (cα (t))
∂xi
i=1
along γ = ψα ◦ cα : I → M is given by
m m
X
ξ˙αk (t) +
X ∂ψα
∇X(t) = Γkα,ij (c(t))ξαi (t)ċjα (t) (cα (t)). (3.7.12)
∂xk
k=1 i,j=1
Geodesics
175
176 CHAPTER 4. GEODESICS
1 b
Z
E(γ) := |γ̇(t)|2 dt. (4.1.2)
2 a
A variation of γ is a family of smooth curves γs : I → M , where s
ranges over the reals, such that the map R × I → M : (s, t) 7→ γs (t) is
smooth and γ0 = γ. The variation {γs }s∈R is said to have fixed endpoints
iff γs (a) = γ(a) and γs (b) = γ(b) for all s ∈ R.
Remark 4.1.2. The length of a continuousPfunction γ : [a, b] → Rn can
be defined as the supremum of the numbers N i=1 |γ(ti ) − γ(ti−1 )| over all
partitions a = t0 < t1 < · · · < tN = b of the interval [a, b]. By a theorem in
first year analysis [47] this supremum is finite whenever γ is continuously
differentiable and is given by (4.1.1).
We shall sometimes suppress the notation for the endpoints of a, b ∈ I.
When γ(a) = p and γ(b) = q we say that γ is a curve from p to q. One
can always compose γ with an affine reparametrization t0 = a + (b − a)t to
obtain a new curve γ 0 (t) := γ(t0 ) on the unit interval 0 ≤ t ≤ 1. This new
curve satisfies L(γ 0 ) = L(γ) and E(γ 0 ) = (b − a)E(γ). More generally, the
length L(γ), but not the energy E(γ), is invariant under reparametrization.
Remark 4.1.3 (Reparametrization). Let I = [a, b] and I 0 = [a0 , b0 ] be
compact intevrals. If γ : I → Rn is a smooth curve and σ : I 0 → I is a
smooth function such that σ(a0 ) = a, σ(b0 ) = b, and σ̇(t) ≥ 0 for all t ∈ I 0 ,
then
L(γ ◦ σ) = L(γ). (4.1.3)
To see this, we compute
Z b0 Z b0
d 0
0 0
L(γ ◦ σ) = γ(σ(t )) dt = γ̇(σ(t ))σ̇(t0 ) dt0 = L(γ).
0
a0 dt
a0
Here second equation follows from the chain rule and the fact that σ̇(t0 ) ≥ 0
for all t0 ∈ [a0 , b0 ], and the third equation follows from the change of variables
formula for the Riemann integral. This proves equation (4.1.3).
4.1. LENGTH AND ENERGY 177
Remark 4.1.6. (i) The conditions (i) and (ii) in Theorem 4.1.4 are mean-
ingless when I is not compact because then the curve has at most one
endpoint and the length and energy integrals may be infinite. However, the
conditions (iii), (iv), and (v) in Theorem 4.1.4 are equivalent for smooth
curves on any interval, compact or not.
(ii) The function s 7→ E(γs ) associated to a smooth variation is always
smooth and so condition (i) in Theorem 4.1.4 is meaningful. However, more
care has to be taken in part (ii) because the function s 7→ L(γs ) need not be
differentiable. It is differentiable at s = 0 whenever γ̇(t) 6= 0 for all t ∈ I.
178 CHAPTER 4. GEODESICS
That every tangent vector of the path space Ωp,q at γ is a vector field along γ
vanishing at the endpoints follows from the above discussion. The converse
inclusion is the content of the next lemma.
4.1. LENGTH AND ENERGY 179
That (iii) implies (i) follows directly from this identity. To prove that (i) im-
plies (iv) we argue indirectly and assume that there exists a point t0 ∈ [0, 1]
such that γ̈(t0 ) is not orthogonal to Tγ(t0 ) M . Then there exists a vec-
tor v0 ∈ Tγ(t0 ) M such that hγ̈(t0 ), v0 i > 0. We may assume without loss
of generality that a < t0 < b. Then there exists a constant ε > 0 such
that a < t0 − ε < t0 + ε < b and
Now choose a smooth cutoff function β : I → [0, 1] such that β(t) = 0 for
all t ∈ I with |t − t0 | ≥ ε and β(t0 ) = 1. Define X ∈ Tγ Ωp,q by
Then hγ̈(t), X(t)i ≥ 0 for all t and hγ̈(t0 ), X(t0 )i > 0. Hence
Z b
dE(γ)X = − hγ̈(t), X(t)i dt < 0
a
and so γ does not satisfy (i). Thus (i) is equivalent to (iii) and (iv).
4.1. LENGTH AND ENERGY 181
We prove that (i) is equivalent to (ii). Assume first that γ satisfies (i).
Then γ also satisfies (iv) and hence γ̈(t) ⊥ Tγ(t) M for all t ∈ I. This implies
1d
0 = hγ̈(t), γ̇(t)i = |γ̇(t)|2 .
2 dt
Hence the function I → R : t 7→ |γ̇(t)|2 is constant. Choose c ≥ 0 such
that |γ̇(t)| ≡ c. If c = 0, then γ(t) is constant and so γ(t) ≡ p = q. If c > 0,
then
Z b
d
dL(γ)X = |γ̇s (t)| dt
ds s=0 a
Z b
∂
= |γ̇s (t)| dt
a ∂s s=0
Z b
−1 ∂
= |γ̇(t)| γ̇(t), γ̇ s (t) dt
a ∂s s=0
1 bD
Z E
= γ̇(t), Ẋ(t) dt
c a
1
= dE(γ)X.
c
Thus, in the case c > 0, γ is an extremal point of E if and only if it is an
extremal point of L. Hence (i) is equivalent to (ii).
We prove that (iii) is equivalent to (v). Let (Φ, γ, γ 0 ) be a development
of M along M 0 = Rm . Then γ̇ 0 (t) = Φ(t)γ̇(t) and dt d
Φ(t)X(t) = Φ(t)∇X(t)
for all X ∈ Vect(γ) and all t ∈ I. Take X = γ̇ to obtain γ̈ 0 (t) = Φ(t)∇γ̇(t) for
all t ∈ I. Thus ∇γ̇ ≡ 0 if and only if γ̈ 0 ≡ 0. This proves Theorem 4.1.4.
Remark 4.1.3 shows that reparametrization by a nundecreasing surjective
map σ : I 0 → I gives rise to map
Ωp,q (I) → Ωp,q (I 0 ) : γ 7→ γ ◦ σ
which preserves the length functional, i.e.
L(γ ◦ σ) = L(γ)
for all γ ∈ Ωp,q (I). Thus the chain rule in infinite dimensions should assert
that if γ◦σ is an extremal (i.e. critical) point of L, then γ is an extremal point
of L. Moreover, if σ is a diffeomorphism, the map γ 7→ γ ◦ σ is bijective and
should give rise to a bijective correspondence between the extremal points
of L on Ωp,q (I) and those on Ωp,q (I 0 ). Finally, if the tangent vector field γ̇
vanishes nowhere, then γ can be parametrized by the arclength. This is
spelled out in more detail in the next exercise.
182 CHAPTER 4. GEODESICS
γ̇(t) 6= 0
(i) Prove that there exists a unique diffeomorphism σ : [0, T ] → I such that
Z t
0 0
σ(t ) = t ⇐⇒ t = |γ̇(s)| ds
a
for all t0 ∈ [0, T ] and all t ∈ [a, b]. Prove that γ 0 := γ ◦ σ : [0, T ] → M is
parametrized by the arclength, i.e. |γ̇ 0 (t0 )| = 1 for all t0 ∈ [0, T ].
(ii) Prove that
Z b
dL(γ)X = − hV̇ (t), X(t)i dt, V (t) := |γ̇(t)|−1 γ̇(t). (4.1.11)
a
4.2 Distance
Assume that M ⊂ Rn is a connected smooth m-dimensional submanifold.
Two point p, q ∈ M are of distance |p − q| apart in the ambient Euclidean
space Rn . In this section we define a distance function which is more in-
timately tied to M by minimizing the length functional over the space of
curves in M with fixed endpoints. Thus it may happen that two points
in M have a very short distance in Rn but can only be joined by very long
curves in M (see Figure 4.1). This leads to the intrinsic distance in M .
Throughout we denote by I = [0, 1] the unit interval and, for p, q ∈ M , by
p
q
M p q
The inequality d(p, q) ≥ 0 holds because each curve has nonnegative length
and the inequality d(p, q) < ∞ holds because Ωp,q 6= ∅.
Remark 4.2.2. Every smooth curve γ : [0, 1] → Rn with endpoints γ(0) = p
and γ(1) = q satisfies the inequality
Z 1 Z 1
L(γ) = |γ̇(t)| dt ≥
γ̇(t) dt = |p − q| .
0 0
Thus d(p, q) ≥ |p − q|. For γ(t) := p + t(q − p) we have equality and hence
the straight lines minimize the length among all curves from p to q.
184 CHAPTER 4. GEODESICS
Hence d(p, r) < d(p, q) + d(q, r) + 2ε for every ε > 0. This proves part (iii)
and Lemma 4.2.3.
q
p
be the unit sphere in R3 and fix two points p, q ∈ S 2 . Then d(p, q) is the
length of the shortest curve on the 2-sphere connecting p and q. Such a
curve is a segment on a great circle through p and q (see Figure 4.2) and its
length is
d(p, q) = cos−1 (hp, qi), (4.2.3)
where hp, qi denotes the standard inner product, and we have
0 ≤ d(p, q) ≤ π.
(See Example 4.3.11 below for details.) By Lemma 4.2.3 this defines a metric
on S 2 . Exercise: Prove directly that (4.2.3) is a distance function on S 2 .
We now have two topologies on our manifold M ⊂ Rn , namely the topol-
ogy determined by the metric d in Lemma 4.2.3 and the relative topology
inherited from Rn . The latter is also determined by a distance function,
namely the extrinsic distance function defined as the restriction of the Eu-
clidean distance function on Rn to the subset M . We denote it by
d0 : M × M → [0, ∞), d0 (p, q) := |p − q| . (4.2.4)
A natural question is if these two metrics d and d0 induce the same topology
on M . In other words is a subset U ⊂ M open with respect to d0 if and only
if it is open with respect to d? Or, equivalently, does a sequence pν ∈ M
converge to p0 ∈ M with respect to d if and only if it converges to p0 with
respect to d0 ? Lemma 4.2.7 answers this question in the affirmative.
Exercise 4.2.6. Prove that every translation of Rn and every orthogonal
transformation preserves the lengths of curves.
Lemma 4.2.7. For every p0 ∈ M we have
d(p, q)
lim = 1.
p,q→p0 |p − q|
Proof. See page 187.
Lemma 4.2.8. Let p0 ∈ M and let φ0 : U0 → Ω0 be a coordinate chart onto
an open subset of Rm such that its derivative dφ0 (p0 ) : Tp0 M → Rm is an
orthogonal transformation. Then
d(p, q)
lim = 1.
p,q→p0 |φ0 (p) − φ0 (q)|
Proof. See page 188.
186 CHAPTER 4. GEODESICS
|p − q| ≤ d(p, q) (4.2.5)
Bρ (p0 , d) ⊂ U.
With
ρ
ρ0 := min δ,
1+ε
this implies Bρ0 (p0 , d0 ) ⊂ U . Namely, if p ∈ M satisfies
|p − p0 | < ρ0 ≤ δ,
then
d(p, p0 ) ≤ (1 + ε)|p − p0 | < (1 + ε)ρ0 ≤ ρ
and so p ∈ U . Thus U is d0 -open and this proves that (i) is equivalent to (ii).
That (ii) implies (iii) follows from the fact that each coordinate chart φ0
is a homeomorphism. To prove that (iii) implies (i), we argue indirectly
and assume that U is not d-open. Then there exists a sequence pν ∈ M \ U
that converges to an element p0 ∈ U . Let φ0 : U0 → Ω0 be a coordinate
chart with p0 ∈ U0 . Then limν→∞ |φ0 (pν ) − φ0 (p0 )| = 0 by Lemma 4.2.8.
Thus φ0 (U0 ∩ U ) is not open and so U does not satisfy (iii).
4.2. DISTANCE 187
Tp 0M ⊥
M
p0
Tp 0M
f : Tp0 M → Tp0 M ⊥
for all p ∈ W (see Figure 4.3). Moreover, by definition the map f satisfies
Hence there exists a constant δ > 0 such that, for every x ∈ Tp0 M , we have
|df (x)bx|
|x| < δ =⇒ x + f (x) ∈ W and kdf (x)k = sup < ε.
b∈Tp0 M
06=x |bx|
188 CHAPTER 4. GEODESICS
Define
U0 := {p ∈ M ∩ W | |x(p)| < δ} .
Given p, q ∈ U0 let γ : [0, 1] → M be the curve whose projection to the x-axis
is the straight line joining x(p) to x(q), i.e.
x(γ(t)) = x(p) + t(x(q) − x(p)) =: x(t),
y(γ(t)) = f (x(γ(t))) = f (x(t)) =: y(t).
Then γ(t) ∈ U0 for all t ∈ [0, 1] and
Z t
L(γ) = |ẋ(t) + ẏ(t)| dt
0
Z t
= |ẋ(t) + df (x(t))ẋ(t)| dt
0
Z t
≤ 1 + kdf (x(t))k |ẋ(t)| dt
0
Z t
≤ (1 + ε) |ẋ(t)| dt
0
= (1 + ε) |x(p) − x(q)|
= (1 + ε) |Π(p0 )(p − q)|
≤ (1 + ε) |p − q| .
Hence d(p, q) ≤ L(γ) ≤ (1 + ε) |p − q| and this proves Lemma 4.2.7.
Proof of Lemma 4.2.8. By assumption we have
|dφ0 (p0 )v| = |v|
for all v ∈ Tp0 M . Fix a constant ε > 0. Then, by continuity of the derivative,
there exists a d0 -open neighborhood M0 ⊂ M of p0 such that for all p ∈ M0
and all v ∈ Tp M we have
(1 − ε) |dφ0 (p)v| ≤ |v| ≤ (1 + ε) |dφ0 (p)v| .
Thus for every curve γ : [0, 1] → M0 we have
(1 − ε)L(φ0 ◦ γ)) ≤ L(γ) ≤ (1 + ε)L(φ0 ◦ γ).
One is tempted to take the infimum over all curves γ : [0, 1] → M0 joining
two pints p, q ∈ M0 to obtain the inequality
(1 − ε) |φ0 (p) − φ0 (q)| ≤ d(p, q) ≤ (1 + ε) |φ0 (p) − φ0 (q)| . (4.2.7)
However, we must justify these inequalities by showing that the infimum
over all curves in M0 agrees with the infimum over all curves in M joining
the points p and q.
4.2. DISTANCE 189
A next question one might ask is the following. Can we choose a coor-
dinate chart φ : U → Ω on M with values in an open set Ω ⊂ Rm so that
the length of each smooth curve γ : [0, 1] → U is equal to the length of the
curve c := φ ◦ γ : [0, 1] → Ω? We examine this question by considering the
inverse map ψ := φ−1 : Ω → U . Denote the components of x and ψ(x) by
Given a smooth curve [0, 1] → Ω : t 7→ c(t) = (c1 (t), . . . , cm (t)) we can write
the length of the composition γ = ψ ◦ c : [0, 1] → M in the form
1
Z
d
L(ψ ◦ c) = ψ(c(t)) dt
dt
0
v
Z 1u n 2
uX d ν
= t ψ (c(t)) dt
0 dt
ν=1
v !2
Z 1u n m ν
uX X ∂ψ
= t (c(t))ċi (t) dt
0 ∂xi
ν=1 i=1
v
u n X m
Z 1 uX
∂ψ ν ∂ψ ν
= t
i
(c(t)) j (c(t))ċi (t)ċj (t) dt
0 ∂x ∂x
ν=1 i,j=1
v
u m
Z 1uX
= t ċi (t)gij (c(t))ċj (t) dt.
0 i,j=1
190 CHAPTER 4. GEODESICS
for every smooth curve c : [0, 1] → Ω. Thus the condition L(ψ ◦ c) = L(c)
for every such curve is equivalent to
This means that ψ preserves angles and areas. The next example shows that
for M = S 2 it is impossible to find such coordinates.
α
β Aγ
α + β + γ = π + A.
Proof. We prove the inclusion “⊂” in (4.3.1). Let (b p, vb) ∈ T(p,v) T M and
choose a smooth curve R → T M : t 7→ (γ(t), X(t)) such that
and hence (1l − Π(γ(t)))Ẋ(t) = hγ(t) (γ̇(t), X(t)) for all t ∈ R. Take t = 0
to obtain (1l − Π(p))b
v = hp (b
p, v). This proves the inclusion “⊂” in (4.3.1).
Equality holds because both sides of the equation are 2m-dimensional linear
subspaces of Rn × Rn .
for all (p, v) ∈ T M . A special case is where S1 (p, v) = v. Such vector fields
correspond to second order differential equations on M .
Definition 4.3.2 (Spray). A vector field S ∈ Vect(T M ) is called a spray
iff it has the form S(p, v) = (v, S2 (p, v)) where S2 : T M → Rn is a smooth
map satisfying
p
M
V := {(p, v) | p ∈ M, v ∈ Vp } ⊂ T M
Ip,v = {t ∈ R | tv ∈ Vp }
Proof. Part (i) follows directly from Lemma 4.3.3 and Theorem 2.4.9. To
prove part (ii), fix an element p ∈ M and a tangent vector v ∈ Vp , and
let γ : Ip,v → M be the unique geodesic with γ(0) = p and γ̇(0) = v. Fix a
nonzero real number λ and define the map γλ : λ−1 Ip,v → M by
Then γ̇λ (t) = λγ̇(λt) ans γ̈λ (t) = λ2 γ̈(λt) and hence
λ ∈ Ip,v ⇐⇒ 1 ∈ Ip,λv ⇐⇒ λv ∈ Vp
and
γ(λ) = γλ (1) = expp (λv)
for λ ∈ Ip,v . This proves Lemma 4.3.6.
194 CHAPTER 4. GEODESICS
sin(|v|)
expp (v) = cos(|v|)p + v
|v|
Example 4.3.12. Consider the orthogonal group O(n) ⊂ Rn×n with the
standard inner product hv, wi := trace(v T w) on Rn×n . The orthogonal
projection Π(g) : Rn×n → Tg O(n) is given by
1
Π(g)v := v − gv T g
2
and the second fundamental form by
hg (v, v) = −gv T v.
for k = 1, . . . , m.
Proof. This follows immediately from the definition of geodesics and equa-
tion (3.6.7) in Lemma 3.6.1 with X = γ̇ and ξ = ċ.
We remark that Lemma 4.3.15 gives rise to another proof of Lemma 4.3.4
that is based on the existence and uniqueness of solutions of second order
differential equations in local coordinates.
Exercise 4.3.16. Let Ω ⊂ Rm be an open set and g = (gij ) : Ω → Rm×m
be a smooth map with values in the space of positive definite symmetric
matrices. Consider the energy functional
Z 1
E(c) := L(c(t), ċ(t)) dt
0
on the space of paths c : [0, 1] → Ω, where L : Ω × Rm → R is defined by
m
1 X i
L(x, ξ) := ξ gij (x)ξ j . (4.3.7)
2
i,j=1
Remark 4.4.3. (i) Condition (i) says that (the velocity |γ̇| is constant
and) L(γ) = d(p, q), i.e. that γ is a shortest curve from p to q. It is not
precluded that there be more than one such γ; consider for example the
case where M is a sphere and p and q are antipodal.
(ii) Condition (ii) implies that
d
E(γs ) = 0
ds s=0
Proof of Lemma 4.4.1. We prove that (i) implies (ii). Let (c) be the (con-
stant) value of |γ̇(t)|. Then
(b − a)c2
L(γ) = (b − a)c, E(γ) = .
2
Then, for every smooth curve γ 0 : I → M with γ 0 (a) = p and γ 0 (b) = q, we
have
4E(γ)2 = c2 L(γ)2
≤ c2 L(γ 0 )2
Z b 2
2
0
=c γ̇ (t) dt
a
Z b
2 γ̇ 0 (t)2 dt
≤ c (b − a)
a
= 2(b − a)c2 E(γ 0 )
= 4E(γ)E(γ 0 ).
Here the fourth step follows from the Cauchy–Schwarz inequality. Now
divide by 4E(γ) to obtain E(γ) ≤ E(γ 0 ).
We prove that (ii) implies (i). We have already shown in Remark 4.4.3
that (ii) implies that γ is a geodesic. It is easy to dispose of the case where
M is one-dimensional. In that case any γ minimizing E(γ) or L(γ) must be
monotonic onto a subarc; otherwise it could be altered so as to make the
integral smaller. Hence suppose M is of dimension at least two. Suppose,
by contradiction, that L(γ 0 ) < L(γ) for some curve γ 0 from p to q. Since
the dimension of M is bigger than one, we may approximate γ 0 by a curve
whose tangent vector nowhere vanishes, i.e. we may assume without loss of
generality that γ̇ 0 (t) 6= 0 for all t. Then we can reparametrize γ 0 proportional
to arclength without changing its length, and by a further transformation
we can make its domain equal to I. Thus we may assume without loss of
generality that γ 0 : I → M is a smooth curve with γ 0 (a) = p and γ 0 (b) = q
such that |γ 0 (t)| = c0 and
(b − a)c0 2 (b − a)c2
E(γ 0 ) = < = E(γ).
2 2
This contradicts (ii) and proves Lemma 4.4.1.
4.4. MINIMAL GEODESICS 199
and a curve γ ∈ Ωp,q has minimal length L(γ) = |v| if and only if there is a
smooth map β : [0, 1] → [0, 1] satisfying
β(0) = 0, β(1) = 1, β̇ ≥ 0
Ur
∂α ∂2α
∇s = Π(α) 2 ≡ 0.
∂s ∂s
By Theorem 4.1.4, the function
∂α
s 7→ (s, t)
∂s
is constant for every t, so that
∂α
(s, t) = ∂α (0, t) = |w(t)| = r
∂s ∂s for (s, t) ∈ R × I.
This implies
∂ ∂α ∂α ∂α ∂α ∂α ∂α
, = ∇s , + , ∇s
∂s ∂s ∂t ∂s ∂t ∂s ∂t
2
∂α ∂ α
= , Π(α)
∂s ∂s∂t
∂α ∂ 2 α
= Π(α) ,
∂s ∂s∂t
∂α ∂ 2 α
= ,
∂s ∂s∂t
1 ∂ ∂α 2
=
2 ∂t ∂s
= 0.
is open in the relative topology of (0, 1]. Thus I is a union of open intervals
in (0, 1) and one half open interval containing 1. Define β : [0, 1] → [0, 1]
and w : I → Tp M by
|v(t)| v(t)
β(t) := , w(t) := ε .
ε |v(t)|
and
|w(t)| = ε, γ(t) = expp (β(t)w(t))
for all t ∈ I. We prove that L(γ) ≥ ε. To see this let α : [0, 1] × I → M be
the map of Lemma 4.4.5, i.e.
∂α ∂α
γ̇(t) = β̇(t) (β(t), t) + (β(t), t)
∂s ∂t
for every t > 0. Hence it follows from Lemma 4.4.5 that
2 2
2 ∂α
2
∂α
|γ̇(t)| = β̇(t) (β(t), t) + (β(t), t) ≥ β̇(t)2 ε2
∂s ∂t
Here the last equation follows by applying the fundamental theorem of cal-
culus to each interval in I and using the fact that β(0) = 0 and β(1) = 1.
If L(γ) = ε, we must have
∂α
(β(t), t) = 0, β̇(t) ≥ 0 for all t ∈ I.
∂t
Thus I is a single half open interval containing 1 and on this interval the
condition ∂α∂t (β(t), t) = 0 implies ẇ(t) = 0. Since w(1) = v we have w(t) = v
for every t ∈ I. Hence γ(t) = expp (β(t)v) for every t ∈ [0, 1]. It follows
that β is smooth on the closed interval [0, 1] (and not just on I). Thus we
have proved that every γ ∈ Ωp,q with values in U ε has length L(γ) ≥ ε with
equality if and only if γ is a reparametrized geodesic. But if γ ∈ Ωp,q does not
take values only in U ε , there must be a T ∈ (0, 1) such that γ([0, T ]) ⊂ U ε
and γ(T ) ∈ ∂Uε . Then L(γ|[0,T ] ) ≥ ε, by what we have just proved,
and L(γ|[T,1] ) > 0 because the restriction of γ to [T, 1] cannot be constant;
so in this case we have L(γ) > ε. This proves Theorem 4.4.4.
The next corollary gives a partial answer to our problem of finding length
minimizing curves. It asserts that geodesics minimize the length locally.
Corollary 4.4.6. Let M ⊂ Rn be a smooth m-manifold, let I ⊂ R be an
open interval, and let γ : I → M be a geodesic. Fix a point t0 ∈ I. Then
there exists a constant ε > 0 such that
t0 − ε < s < t < t0 + ε =⇒ L(γ|[s,t] ) = d(γ(s), γ(t)).
Proof. Since γ is a geodesic its derivative has constant norm |γ̇(t)| ≡ c (see
Theorem 4.1.4). Choose δ > 0 so small that the interval [t0 − δ, t0 + δ] is
contained in I. Then there is a constant r > 0 such that r ≤ inj(γ(t))
whenever |t − t0 | ≤ δ. Choose ε > 0 such that
ε < δ, 2εc < r.
If t0 − ε < s < t < t0 + ε, then
γ(t) = expγ(s) ((t − s)γ̇(s))
and
|(t − s)γ̇(s)| = |t − s| c < 2εc < r ≤ inj(γ(s)).
Hence it follows from Theorem 4.4.4 that
L(γ|[s,t] ) = |t − s| c = d(γ(s), γ(t)).
This proves Corollary 4.4.6.
4.4. MINIMAL GEODESICS 203
for 0 < r < inj(p; M ). Hence Theorem 4.5.3 below asserts that Br (p) is
convex for r > 0 sufficiently small.
(i) Show that it can happen that a geodesic in Br (p) is not minimal. Hint:
Take M to be the hemisphere {(x, y, z) ∈ R3 | x2 + y 2 + z 2 = 1, z > 0} to-
gether with the disc {(x, y, z) ∈ R3 | x2 + y 2 ≤ 1, z = 0}, but smooth the cor-
ners along the circle x2 + y 2 = 1, z = 0. Take p = (0, 0, 1) and r = π/2.
(ii) Show that, if r > 0 is sufficiently small, then the unique geodesic γ
in Br (p) joining two points q, q 0 ∈ Br (p) is minimal and that in fact any
curve γ 0 from q to q 0 which is not a reparametrization of γ is strictly longer,
i.e. L(γ 0 ) > L(γ) = d(q, q 0 ).
204 CHAPTER 4. GEODESICS
for all t ∈ I. (We remark that the hypothesis γ̇(t) 6= 0 implies that σ is
actually a diffeomorphism, i.e. σ̇(t) > 0 for all t ∈ I.)
(iii) The curve γ minimizes the length functional locally, i.e. there ex-
ists an ε > 0 such that L(γ|[s,t] ) = d(γ(s), γ(t)) for every closed subinter-
val [s, t] ⊂ I of length t − s < ε.
It is often convenient to consider curves γ where γ̇(t) is allowed to vanish
for some values of t; then γ cannot (in general) be parametrized by arclength.
Such a curve γ : I → M can be smooth (as a map) and yet its image may
have corners (where γ̇ necessarily vanishes). Note that a curve with corners
can never minimize the distance, even locally.
Exercise 4.4.12. Show that conditions (ii) and (iii) in Exercise 4.4.11 are
equivalent, even without the assumption that γ̇ is nowhere vanishing. De-
duce that, if γ : I → M is a shortest curve joining p to q, i.e. L(γ) = d(p, q),
then γ is a reparametrized geodesic.
Show by example that one can have a variation {γs }s∈R of a reparame-
trized geodesic γ0 = γ for which the map s 7→ L(γs ) is not even differentiable
at s = 0. (Hint: Take γ to be constant. See also Exercise 4.1.9.)
Show, however, that conditions (i), (ii) and (iii) in Exercise 4.4.11 remain
equivalent if the hypothesis that γ̇ is nowhere vanishing is weakened to the
hypothesis that γ̇(t) 6= 0 for all but finitely many t ∈ I. Conclude that a bro-
ken geodesic is a reparametrized geodesic if and only if it minimizes arclength
locally. (A broken geodesic is a continuous map γ : I = [a, b] → M for
which there exist a = t0 < t1 < · · · < tn = b such that γ|[ti−1 ,ti ] is a geodesic
for i = 1, . . . , n. It is thus a geodesic if and only if γ̇ is continuous at the
break points, i.e. γ̇(t− +
i ) = γ̇(ti ) for i = 1, . . . , n − 1.)
4.5. CONVEX NEIGHBORHOODS 205
is geodesically convex.
206 CHAPTER 4. GEODESICS
φ := ψ −1 : U → Ω (4.5.5)
c(t) := φ(γ(t))
Hence
d2 |c|2 d 2 |ċ|2
= hċ, ci = | ċ| + hc̈, ci = Qc ( ċ) ≥ ≥0
dt2 2 dt 2
and so the function t 7→ |φ(γ(t))|2 is convex. Thus, if γ(0), γ(1) ∈ Ur for
some r > 0, it follows that γ(t) ∈ Ur for all t ∈ [0, 1].
Consider the exponential map
in Lemma 4.3.6. Its domain V is open and the exponential map is smooth.
Since it sends the pair (p0 , 0) ∈ V to expp0 (0) = p0 ∈ U , it follows from con-
tinuity that there exist constants ε > 0 and r > 0 such that
Moreover, we have
by Corollary 4.3.7. Hence the Implicit Function Theorem 2.6.15 asserts that
the constants ε > 0 and r > 0 can be chosen such that (4.5.6) holds and there
exists a smooth map h : Ur × Ur → Rn that satisfies the conditions
h(p0 , p0 ) = 0
d : M × M → [0, ∞)
Exercise 4.6.4. Let (M, d) be a metric space. Prove that every compact
subset K ⊂ M is closed and bounded. Find an example of a metric space
that contains a closed and bounded subset that is not compact.
210 CHAPTER 4. GEODESICS
Theorem 4.6.6 implies Theorem 4.6.5. That (i) implies (ii) follows directly
from the definitions.
We prove that (ii) implies (iii). Thus assume that M is geodesically
complete at the point p0 ∈ M and let K ⊂ M be a closed and bounded
subset. Then r := supq∈K d(p0 , q) < ∞. Hence Theorem 4.6.6 asserts that,
for every q ∈ K, there exists a vector v ∈ Tp0 M such that |v| = d(p0 , q) ≤ r
and expp0 (v) = q. Thus
K ⊂ expp0 (B r (p0 )), B r (p0 ) = {v ∈ Tp0 M | |v| ≤ r} .
Then B := {v ∈ Tp0 M | |v| ≤ r, expp0 (v) ∈ K} is a closed and bounded sub-
set of the Euclidean space Tp0 M . Hence B is compact and K = expp0 (B).
Since the exponential map expp0 : Tp0 M → M is continuous it follows that K
is compact. This shows that (ii) implies (iii).
4.6. COMPLETENESS AND HOPF–RINOW 211
We prove that (iii) implies (iv). Thus assume that every closed and
bounded subset of M is compact and choose a Cauchy sequence pi ∈ M .
Choose i0 ∈ N such that d(pi , pj ) ≤ 1 for all i, j ∈ N with i, j ≥ i0 . Define
c := max d(p1 , pi ) + 1.
1≤i≤i0
Then d(p1 , pi ) ≤ d(p1 , pi0 ) + d(pi0 , pi ) ≤ d(p1 , pi0 ) + 1 ≤ c for all i ≥ i0 and
so d(p1 , pi ) ≤ c for all i ∈ N. Hence the set {pi | i ∈ N} is bounded and so
is its closure. By (iii) this implies that the sequence pi has a convergent
subsequence. Since pi is a Cauchy sequence, this implies that pi converges.
Thus we have proved that (iii) implies (iv).
We prove that (iv) implies (v). Fix a smooth curve ξ : R → Rm and
an element (p0 , e0 ) ∈ F(M ). Assume, by contradiction, that there exists
a real number T > 0 such that there exists a solution β : [0, T ) → F(M )
of equation (4.6.1) that cannot be extended to the interval [0, T + ε) for
any ε > 0. Write β(t) =: (γ(t), e(t)) so that γ and e satisfy the equations
γ̇(t) = e(t)ξ(t), ė(t) = hγ(t) (γ̇(t))e(t), γ(0) = p0 , e(0) = e0 .
⊥ M for all η ∈ Rm and therefore
This implies e(t)η ∈ Tγ(t) M and ė(t)η ∈ Tγ(t)
d d
hη, e(t)T e(t)ζi = he(t)η, e(t)ζi = hė(t)η, e(t)ζi + he(t)η, ė(t)ζi = 0
dt dt
for all η, ζ ∈ Rm and all t ∈ [0, T ). Thus the function t 7→ e(t)T e(t) is con-
stant, hence
|e(t)η|
e(t)T e(t) = eT
0 e0 , ke(t)k = sup = ke0 k (4.6.2)
06=η∈Rm |η|
for 0 ≤ t < T , hence
|γ̇(t)| = |e(t)ξ(t)| ≤ ke0 k |ξ(t)| ≤ ke0 k sup |ξ(s)| =: cT
0≤s≤T
and so d(γ(s), γ(t)) ≤ L(γ|[s,t] ) ≤ (t − s)cT for 0 ≤ s < t < T . Since (M, d)
is a complete metric space, this shows that the limit p1 := limt%T γ(t) ∈ M
exists. Thus the set K := γ([0, T )) ∪ {p1 } ⊂ M is compact and so is the set
n o
Ke := (p, e) ∈ F(M ) | p ∈ K, eT e = eT e0 ⊂ F(M ).
0
By equation (4.6.2) the curve [0, T ) → R × F(M ) : t 7→ (t, γ(t), e(t)) takes
values in the compact set [0, T ] × K
e and is the integral curve of a vector field
on the manifold R × F(M ). Hence Corollary 2.4.15 asserts that [0, T ) cannot
be the maximal existence interval of this integral curve, a contradiction. This
shows that (iv) implies (v).
212 CHAPTER 4. GEODESICS
This shows that Sε (p) = expp (εΣ1 (p)) and, since ε is smaller than the injec-
tivity radius, the map
Σ1 (p) → Sε (p) : v 7→ expp (εv)
is a diffeomorphism.
4.6. COMPLETENESS AND HOPF–RINOW 213
r := d(p, q) > ε.
Fix a constant δ > 0 and choose a smooth curve γ : [0, 1] → M such that
Choose t0 > 0 such that γ(t0 ) is the last point of the curve on Sε (p), i.e.
Then
This shows that d(Sε (p), q) ≤ r + δ − ε for every δ > 0 and therefore
d(Sε (p), q) ≤ r − ε.
Moreover,
d(p0 , q) ≥ d(p, q) − d(p, p0 ) = r − ε
d(Sε (p), q) = r − ε.
d(γ(t), q) = r − t.
and
d(γ(s), q) ≥ d(p, q) − d(p, γ(s)) ≥ r − s.
Hence d(γ(s), q) = r − s and hence s ∈ I. This proves (4.6.4).
216 CHAPTER 4. GEODESICS
We prove that I is open (in the relative topology of [0, r]). Let t ∈ I
be given with t < r. Choose a constant ε > 0 smaller than the injectivity
radius of M at γ(t) and smaller than r − t. Then, by Lemma 4.6.7 with p
replaced by γ(t), we have
d(Sε (γ(t)), q) = r − t − ε.
Next we choose w ∈ Tγ(t) M such that
|w| = 1, d(expγ(t) (εw), q) = r − t − ε.
Then
d(γ(t − ε), expγ(t) (εw)) ≥ d(γ(t − ε), q) − d(expγ(t) (εw), q)
= (r − t + ε) − (r − t − ε)
= 2ε.
The converse inequality is obvious, because both points have distance ε
to γ(t) (see Figure 4.8).
Sε (γ (t)) exp (ε w)
γ(t)
γ(t)
ε ε q
p r−t−ε
Let x0 := φ(p0 ) ∈ Ω and choose r > 0 such that B r (x0 ) ⊂ Ω. Then there is
a constant δ ∈ (0, 1] such that
q
δ |ξ| ≤ ξ T g(x)ξ ≤ δ −1 |ξ| (4.7.3)
L(γ) ≥ δr.
∇γ̇ = 0,
as in Definition 4.1.5. Then all the above results about geodesics, as well
as their proofs, carry over almost verbatim to the intrinsic setting. In
particular, geodesics are in local coordinates described by equation (4.3.6)
(Lemma 4.3.15) and they are the critical points of the energy functional
1 1
Z
E(γ) := |γ̇(t)|2 dt
2 0
on the space Ωp,q of all paths γ : [0, 1] → M with fixed endpoints γ(0) = p
and γ(1) = q. Here we use the fact that Lemma 4.1.7 extends to the in-
trinsic setting via the Embedding Theorem 2.9.12. So for every vector
field X ∈ Vect(γ) along γ with X(0) = 0 and X(1) = 0 there exists a curve
of curves R → Ωp,q : s 7→ γs with γ0 = γ and ∂s γs |s=0 = X. Then, by the
properties of the Levi-Civita connection, we have
1 1
Z
dE(γ)X = ∂s |∂t γs (t)|2 dt
2 0
Z 1
= hγ̇(t), ∇t X(t)i dt
0
Z 1
=− h∇t γ̇(t), X(t)i dt.
0
The right hand side vanishes for all X if and only if ∇γ̇ ≡ 0 (Theorem 4.1.4).
With this understood, we find that, for all p ∈ M and v ∈ Tp M , there exists
a unique geodesic γ : Ip,v → M on a maximal open interval Ip,v ⊂ R con-
taining zero that satisfies γ(0) = p and γ̇(0) = v (Lemma 4.3.4). This gives
rise to a smooth exponential map expp : Vp = {v ∈ Tp M | 1 ∈ Ip,v } → M as
in §4.3 which satisfies d expp (0) = id : Tp M → Tp M as in Corollary 4.3.7.
This leads directly to the injectivity radius, the Gauß Lemma 4.4.5, the local
length minimizing property of geodesics in Theorem 4.4.4, and the Convex
Neighborhood Theorem 4.5.3. Also the proof of the equivalence of metric
and geodesic completeness in Theorem 4.6.5 and of the Hopf–Rinow Theo-
rem 4.6.6 carry over verbatim to the intrinsic setting of general Riemannian
manifolds. The only place where some care must be taken is in the proof of
the Curve Shortening Lemma 4.6.8 as is spelled out in Exercise 4.7.2 below.
220 CHAPTER 4. GEODESICS
Refine the estimate (4.7.1) in the proof of Lemma 4.7.1 and show that
d(p, q)
lim = 1.
p,q→p0 |φ(p) − φ(q)|
This is the intrinsic analogue of Lemma 4.2.8. Use this to prove that equa-
tion (4.6.3) continues to hold for all Riemannian manifolds, i.e.
d(expp (δv), expp (δw))
lim = |v − w|
δ→0 δ
for p ∈ M and v, w ∈ Tp M . With this understood, the proof of the Curve
Shortening Lemma 4.6.8 carries over verbatim to the intrinsic setting.
Exercise 4.7.3. The real projective space RPn inherits a Riemannian met-
ric from S n as it is a quotient of S n by an isometric involution. Prove that
each geodesic in S n with its standard metric descends to a geodesic in RPn .
Exercise 4.7.4. Let f : S 3 → S 2 be the Hopf fibration defined by
f (z, w) = |z|2 − |w|2 , 2Re z̄w, 2Im z̄w
hX, Y i := trace(X T Y )
Λ = im D, DT D
b = 0, Λ b : Rk → Λ⊥ = ker DT .
b ◦D =D
Prove that the group O(n) acts on Gk (Rn ) by isometries. Which subgroup
acts trivially?
Exercise 4.7.8. Carry over Exercises 4.7.6 and 4.7.7 to the complex Grass-
mannian Gk (Cn ). Prove that the group U(n) acts on Gk (Cn ) by isometries.
Which subgroup acts trivially?
222 CHAPTER 4. GEODESICS
Chapter 5
Curvature
5.1 Isometries
Characterization of Isometries
Let M and M 0 be connected submanifolds of Rn . An isometry is an isomor-
phism of the intrinsic geometries of M and M 0 . Recall the definition of the
intrinsic distance function
d : M × M → [0, ∞)
in §4.2 by Z 1
d(p, q) := inf L(γ), L(γ) = |γ̇(t)| dt
γ∈Ωp,q 0
for p, q ∈ M . Let d0 denote the intrinisic distance function on M 0 .
223
224 CHAPTER 5. CURVATURE
for all p, q ∈ M .
(ii) φ is a diffeomorphism and
dφ(p) : Tp M → Tφ(p) M 0
L(φ ◦ γ) = L(γ)
Lemma 5.1.2. For every p ∈ M there exists a constant ε > 0 such that,
for all v, w ∈ Tp M with 0 < |w| < |v| < ε, we have
|w|
d(expp (w), expp (v)) = |v| − |w| =⇒ w= v. (5.1.1)
|v|
Remark 5.1.3. It follows from the triangle inequality and Theorem 4.4.4
that
d(expp (v), expp (w)) ≥ d(expp (v), p) − d(expp (w), p) = |v| − |w|
whenever 0 < |w| < |v| < inj(p). Lemma 5.1.2 asserts that equality can
only hold when w is a positive multiple of v or, to put it differently, that the
distance between expp (v) and expp (w) must be strictly bigger that |v| − |w|
whenever w is not a positive multiple of v.
5.1. ISOMETRIES 225
By Theorem 4.4.4 and the definition of the injectivity radius, the exponential
map at p is a diffeomorphism expp : Bε (p) → Uε (p) for ε < inj(p). Choose
0 < r < inj(p). Then the closure of Ur (p) is a compact subset of M . Hence
there is a constant ε > 0 such that ε < r and ε < inj(p0 ) for every p0 ∈ Ur (p).
Since ε < r we have
Following first the shortest geodesic from p to p1 and then the shortest
geodesic from p1 to p2 we obtain (after suitable reparametrization) a smooth
γ : [0, 2] → M such that
and
Since w and β(1)v are both elements of Bε (p) and expp is injective on Bε (p),
this implies w = β(1)v. Since β(1) ≥ 0 we have β(1) = |w| / |v|. This
proves (5.1.1) and Lemma 5.1.2.
226 CHAPTER 5. CURVATURE
Proof of Theorem 5.1.1. That (ii) implies (iii) follows from the definition of
the length of a curve. Namely
Z b
d
L(φ ◦ γ) = dt φ(γ(t)) dt
a
Z b
= |dφ(γ(t))γ̇(t)| dt
a
Z b
= |γ̇(t)| dt
a
= L(γ).
In the third equation we have used (ii). That (iii) implies (i) follows imme-
diately from the definition of the intrinsic distance functions d and d0 .
We prove that (i) implies (ii). Fix a point p ∈ M and choose ε > 0 so
small that ε < inj(p) and that the assertion of Lemma 5.1.2 holds for the
point p0 := φ(p) ∈ M 0 . Then there is a unique homeomorphism
Φp : Bε (p) → Bε (φ(p))
such that the following diagram commutes.
Φp
Tp M ⊃ Bε (p) / Bε (φ(p)) ⊂ Tφ(p) M 0 .
expp exp0φ(p)
φ
M ⊃ Uε (p) / Uε (φ(p)) ⊂ M0
Here the vertical maps are diffeomorphisms and φ : Uε (p) → Uε (φ(p)) is a
homeomorphism by (i). Hence Φp : Bε (p) → Bε (φ(p)) is a homeomorphism.
Claim 1. The map Φp satisfies the following equations for every v ∈ Bε (p)
and every t ∈ [0, 1]:
exp0φ(p) (Φp (v)) = φ(expp (v)), (5.1.3)
|Φp (v)| = |v| , (5.1.4)
Φp (tv) = tΦp (v). (5.1.5)
Equation (5.1.3) holds by definition. To prove (5.1.4) we observe that, by
Theorem 4.4.4, we have
|Φp (v)| = d0 (φ(p), exp0φ(p) (Φp (v)))
= d0 (φ(p), φ(expp (v)))
= d(p, expp (v))
= |v| .
5.1. ISOMETRIES 227
Here the second equation follows from (5.1.3) and the third equation from (i).
Equation (5.1.5) holds for t = 0 because Φp (0) = 0 and for t = 1 it is a
tautology. Hence assume 0 < t < 1. Then
d0 (exp0φ(p) (Φp (tv)), exp0φ(p) (Φp (v))) = d0 (φ(expp (tv)), φ(expp (v)))
= d(expp (tv), expp (v))
= |v| − |tv|
= |Φp (v)| − |Φp (tv)| .
Here the first equation follows from (5.1.3), the second equation from (i),
the third equation from Theorem 4.4.4 and the fact that |v| < inj(p), and
the last equation follows from (5.1.4). Since 0 < |Φp (tv)| < |Φp (v)| < ε we
can apply Lemma 5.1.2 and obtain
|Φp (tv)|
Φp (tv) = Φp (v) = tΦp (v).
|Φp (v)|
This proves Claim 1.
By Claim 1, Φp extends to a bijective map Φp : Tp M → Tφ(p) M 0 via
1
Φp (v) := Φp (δv),
δ
where δ > 0 is chosen so small that δ |v| < ε. The right hand side of
this equation is independent of the choice of δ. Hence the extension is well
defined. It is bijective because the original map Φp is a bijection from Bε (p)
to Bε (φ(p)). The reader may verify that the extended map satisfies the
conditions (5.1.4) and (5.1.5) for all v ∈ Tp M and all t ≥ 0.
Claim 2. The extended map Φp : Tp M → Tφ(p) M 0 is linear and preserves
the inner product.
It follows from the equation (4.6.3) in the proof of Lemma 4.6.8 that
d(expp (tv), expp (tw))
|v − w| = lim
t→0 t
d0 (φ(expp (tv)), φ(expp (tw)))
= lim
t→0 t
d0 (exp0φ(p) (Φp (tv))), exp0φ(p) (Φp (tw))))
= lim
t→0 t
d (expφ(p) (tΦp (v))), exp0φ(p) (tΦp (w))))
0 0
= lim
t→0 t
= |Φp (v) − Φp (w)| .
228 CHAPTER 5. CURVATURE
Here the second equation follows from (i), the third from (5.1.3), the fourth
from (5.1.4), and the last equation follows again from (4.6.3). By polariza-
tion we obtain
Since Φp is linear, this shows that the restriction of φ to the open set Uε (p)
is smooth. Moreover, for every v ∈ Tp M we have
d d
dφ(p)v = φ(expp (tv)) = exp0φ(p) (tΦp (v)) = Φp (v).
dt t=0 dt t=0
Here we have used equations (5.1.3) and (5.1.5) as well as Lemma 4.3.6.
This proves Claim 3 and Theorem 5.1.1.
5.1. ISOMETRIES 229
ψ(p) = Ap + b
|v| = |w|
and
|v − ei | = |w − ei | for i = 1, . . . , n
must be equal.
II. Given M and M 0 when are they intrinsically isomorphic, i.e. when is
there an isometry φ : M → M 0 from M onto M 0 ?
230 CHAPTER 5. CURVATURE
As we have noted, both the first and second fundamental forms are
preserved by extrinsic isomorphisms while only the first fundamental form
need be preserved by an intrinsic isomorphism (i.e. an isometry).
A question which occurred to Gauß (who worked for a while as a cartog-
rapher) is this: Can one draw a perfectly accurate map of a portion of the
earth? (i.e. a map for which the distance between points on the map is pro-
portional to the distance between the corresponding points on the surface
of the earth). We can now pose this question as follows: Is there an isom-
etry from an open subset of a sphere to an open subset of a plane? Gauß
answered this question negatively by associating an invariant, the Gaußian
curvature
K : M → R,
to a surface M ⊂ R3 . According to his Theorema Egregium
K0 ◦ φ = K
for an isometry φ : M → M 0 . The sphere has positive curvature; the plane
has zero curvature; hence the perfectly accurate map does not exist. Our
aim is to explain these ideas.
Local Isometries
We shall need a concept slightly more general than that of “isometry”.
Definition 5.1.6. A smooth map φ : M → M 0 is called a local isometry
iff its derivative
dφ(p) : Tp M → Tφ(p) M 0
is an orthogonal linear isomorphism for every p ∈ M .
0
Remark 5.1.7. Let M ⊂ Rn and M 0 ⊂ Rm be manifolds and φ : M → M 0
be a map. The following are equivalent.
(i) φ is a local isometry.
(ii) For every p ∈ M there are open neighborhoods U ⊂ M and U 0 ⊂ M 0
such that the restriction of φ to U is an isometry from U onto U 0 .
That (ii) implies (i) follows immediately from Theorem 5.1.1. On the other
hand (i) implies that dφ(p) is invertible so that (ii) follows from the inverse
function theorem.
Example 5.1.8. The map
R → S 1 : θ 7→ eiθ
is a local isometry but not an isometry.
5.1. ISOMETRIES 231
Rp : Tp M × Tp M → L(Tp M, Tp M )
We must prove that R is well defined, i.e. that the right hand side of
equation (5.2.1) is independent of the choice of γ and Z. This follows from
the Gauß–Codazzi formula which we prove next. Recall that the second fun-
damental form can be viewed as a linear map hp : Tp M → L(Tp M, Tp M ⊥ )
and that, for u ∈ Tp M , the linear map hp (u) ∈ L(Tp M, Tp M ⊥ ) and its
dual hp (u)∗ ∈ L(Tp M ⊥ , Tp M ) are given by
hp (u)∗ w = dΠ(p)u w
hp (u)v = dΠ(p)u v,
for v ∈ Tp M and w ∈ Tp M ⊥ .
5.2. THE RIEMANN CURVATURE TENSOR 233
Theorem 5.2.2. The Riemann curvature tensor is well defined and given
by the Gauß–Codazzi formula
for u, v ∈ Tp M .
∇t Z = ∂t Z − hγ (∂t γ)Z
= ∂t Z − dΠ(γ)∂t γ Z
= ∂t Z − ∂t Π ◦ γ Z.
Hence
∂s ∇t Z = ∂s ∂t Z − ∂s ∂t Π ◦ γ Z
= ∂s ∂t Z − ∂s ∂t Π ◦ γ Z − ∂t Π ◦ γ ∂s Z
= ∂s ∂t Z − ∂s ∂t Π ◦ γ Z − dΠ(γ)∂t γ ∇s Z + hγ (∂s γ)Z
= ∂s ∂t Z − ∂s ∂t Π ◦ γ Z − hγ (∂t γ)∇s Z − hγ (∂t γ)∗ hγ (∂s γ)Z.
Here the first two terms on the right are tangent to M and the last two
terms on the right are orthogonal to Tγ M . Hence
∇s ∇t Z − ∇t ∇s Z = Π(γ) ∂s ∇t Z − ∂t ∇s Z
= hγ (∂s γ)∗ hγ (∂t γ)Z − hγ (∂t γ)∗ hγ (∂s γ)Z.
This proves the Gauß–Codazzi equation and shows that the left hand side
is independent of the choice of γ and Z. This proves Theorem 5.2.2.
234 CHAPTER 5. CURVATURE
∇v X(p) := Π(p)dX(p)v ∈ Tp M,
∇X : Vect(M ) → Vect(M ),
∇f X (Y ) = f ∇X Y, (5.2.6)
∇X (f Y ) = f ∇X Y + (LX f ) Y (5.2.7)
LX hY, Zi = h∇X Y, Zi + hY, ∇X Zi, (5.2.8)
∇Y X − ∇X Y = [X, Y ], (5.2.9)
for all X, Y, Z ∈ Vect(M ) and f ∈ F (M ), where LX f = df ◦ X and
[X, Y ] ∈ Vect(M ) denotes the Lie bracket of the vector fields X and Y .
The next lemma asserts that the Levi-Civita connection is uniquely deter-
mined by (5.2.8) and (5.2.9).
The same equation holds for the Levi-Civita connection and hence
ξ = (ξ 1 , . . . , ξ m ) : Ω → Rm
defined by
ξ(φ(p)) = dφ(p)X(p)
m
X
k` 1 ∂g`i ∂g`j ∂gij
Γkij := g j
+ i
− , (5.2.11)
2 ∂x ∂x ∂x`
`=1
where gij is the metric tensor and g ij is the inverse matrix so that
X
gij g jk = δik
j
(see Lemma 3.6.5). This formula can be used to prove the existence state-
ment in Lemma 5.2.7 and hence define the Levi-Civita connection in the
intrinsic setting.
Exercise 5.2.9. In the proof of Lemma 5.2.7 we did not actually use
that the operator DX : Vect(M ) → Vect(M ) is linear nor that the oper-
ator X 7→ DX is linear. Prove directly that if a map
DX : L(M ) → L(M )
satisfies (5.2.8) for all Y, Z ∈ Vect(M ), then DX is linear. Prove that every
map
Vect(M ) → L(Vect(M ), Vect(M )) : X 7→ DX
Proof. Fix a point p ∈ M . Then the right hand side of equation (5.2.12) at
p remains unchanged if we multiply each of the vector fields X, Y, Z by a
smooth function f : M → [0, 1] that is equal to one near p. Choosing f with
compact support we may therefore assume that the vector fields X and Y
are complete. Let φs denote the flow of X and ψ t the flow of Y . Define the
map γ : R2 → M by
γ(s, t) := φs ◦ ψ t (p), s, t ∈ R.
Then
∂s γ = X(γ), ∂t γ = (φs∗ Y )(γ).
This implies
∇s ∇t (Z ◦ γ) = ∇∂s γ ∇φs∗ Y Z (γ) + ∇∂s φs∗ Y Z (γ).
Since
∂
φs Y = [X, Y ]
∂s s=0 ∗
and ∂s γ = X(γ) we obtain
Hence
Rp (X(p), Y (p))Z(p) = ∇s ∇t (Z ◦ γ) − ∇t ∇s (Z ◦ γ) (0, 0)
= ∇X ∇Y Z(p) − ∇Y ∇X Z(p) + ∇[X,Y ] Z(p).
5.2.4 Symmetries
Theorem 5.2.13. The Riemann curvature tensor satisfies
R(Y, X) = −R(X, Y ) = R(X, Y )∗ , (5.2.16)
R(X, Y )Z + R(Y, Z)X + R(Z, X)Y = 0, (5.2.17)
hR(X, Y )Z, W i = hR(Z, W )X, Y i, (5.2.18)
for X, Y, Z, W ∈ Vect(M ). Equation (5.2.17) is the first Bianchi identity.
Proof. The first equation in (5.2.16) is obvious from the definition and the
second follows from the Gauß–Codazzi formula (5.2.3). Alternatively, choose
a smooth map γ : R2 → M and two vector fields Z, W along γ. Then
0 = ∂s ∂t hZ, W i − ∂t ∂s hZ, W i
= ∂s h∇t Z, W i + ∂s hZ, ∇t W i − ∂t h∇s Z, W i − ∂t hZ, ∇s W i
= h∇s ∇t Z, W i + hZ, ∇s ∇t W i − h∇t ∇s Z, W i − hZ, ∇t ∇s W i
= hR(∂s γ, ∂t γ)Z, W i − hZ, R(∂s γ, ∂t γ)W i.
This proof has the advantage that it carries over to the intrinsic setting. We
prove the first Bianchi identity using (5.2.9) and (5.2.12):
R(X, Y )Z + R(Y, Z)X + R(Z, X)Y
= ∇X ∇Y Z − ∇Y ∇X Z + ∇[X,Y ] Z + ∇Y ∇Z X − ∇Z ∇Y X + ∇[Y,Z] X
+∇Z ∇X Y − ∇X ∇Z Y + ∇[Z,X] Y
= ∇[Y,Z] X − ∇X [Y, Z] + ∇[Z,X] Y − ∇Y [Z, X] + ∇[X,Y ] Z − ∇Z [X, Y ]
= [X, [Y, Z]] + [Y, [Z, X]] + [Z, [X, Y ]].
The last term vanishes by the Jacobi identity. We prove (5.2.18) by com-
bining the first Bianchi identity with (5.2.16):
hR(X, Y )Z, W i − hR(Z, W )X, Y i
= −hR(Y, Z)X, W i − hR(Z, X)Y, W i − hR(Z, W )X, Y i
= hR(Y, Z)W, Xi + hR(Z, X)W, Y i + hR(W, Z)X, Y i
= hR(Y, Z)W, Xi − hR(X, W )Z, Y i
= hR(Y, Z)W, Xi − hR(W, X)Y, Zi.
Note that the first line is related to the last by a cyclic permutation. Re-
peating this argument we find
hR(Y, Z)W, Xi − hR(W, X)Y, Zi = hR(Z, W )X, Y i − hR(X, Y )Z, W i.
Combining the last two identities we obtain (5.2.18). This proves Theo-
rem 5.2.13.
240 CHAPTER 5. CURVATURE
on the ambient space gl(n, R) = Rn×n . This fits into the extrinsic setting
used throughout most of this book. By Theorem 2.5.26 every Lie subgroup
of O(n) is a closed subset of O(n) and hence is compact.
Example 5.2.17. Let G ⊂ O(n) be a Lie subgroup and let
g := Lie(G) = T1l G
For G = O(n) we have seen this in Example 4.3.12 and in the general
case this follows from equation (5.2.20) with X = γ̇. Hence the exponential
map exp : g → G defined by the exponential matrix (as in §2.5) agrees with
the time-1-map of the geodesic flow (as in §4.3).
(d) The Riemann curvature tensor on G is given by
1
g −1 Rg (u, v)w = − [[g −1 u, g −1 v], g −1 w] (5.2.22)
4
for g ∈ G and u, v, w ∈ Tg G. Note that the first Bianchi identity is equivalent
to the Jacobi identity. (Exercise: Prove equation (5.2.22).)
242 CHAPTER 5. CURVATURE
Exercise 5.2.25. Prove that κ([ξ, η], ζ) = κ(ξ, [η, ζ]) for all ξ, η, ζ ∈ g.
Exercise 5.2.26. Assume that g admits an invariant inner product. For
each ξ ∈ g prove that the derivation ad(ξ) is skew-adjoint with respect to this
inner product and deduce that κ(ξ, ξ) = −trace(ad(ξ)∗ ad(ξ)) = −|ad(ξ)|2 .
Deduce that the Killing form is nondegenerate whenever g has a trivial
center and admits an invariant inner product.
Example 5.2.27. Let G ⊂ GL(n, R) be any Lie subgroup (not necessarily
contained in O(n)), and let g := Lie(G) = T1l G be its Lie algebra. Fix
any inner product on the Lie algebra g (not necessarily invariant under
conjugation) and consider the Riemannian metric on G defined by
hv, wig := hvg −1 , wg −1 i (5.2.28)
for v, w ∈ Tg G. This metric is called right invariant.
(a) The map g 7→ ga is an isometry of G for every a ∈ G.
(b) Define the linear map A : g → End(g) by
1
hA(ξ)η, ζi = hξ, [η, ζ]i − hη, [ζ, ξ]i − hζ, [ξ, η]i (5.2.29)
2
for ξ, η, ζ ∈ g. Then A is the unique linear map that satisfies
A(ξ) + A(ξ)∗ = 0, A(η)ξ − A(ξ)η = [ξ, η]
for all ξ, η ∈ g, where A(ξ)∗ is the adjoint operator with respect to the inner
product on g. Let γ : R → G be a smooth curve and X ∈ Vect(γ) be a
smooth vector field along γ. Then the covariant derivative of X is given by
d −1 −1 −1
∇X = (Xγ ) + A(γ̇γ )Xγ γ. (5.2.30)
dt
(Exercise: Prove this. Moreover, if the inner produt on g is invariant, prove
that A(ξ)η = − 21 [ξ, η] for all ξ, η ∈ g.)
(c) A smooth curve γ : R → G is a geodesic if and only if it satisfies
d
(γ̇γ −1 ) + A(γ̇γ −1 )γ̇γ −1 = 0. (5.2.31)
dt
(Exercise: G is complete.)
(d) The Riemann curvature tensor on G is given by
Rg (ξg, ηg)ζg g −1 = A ([ξ, η]) + [A (ξ) , A (η)] ζ
(5.2.32)
5.3.1 Pushforward
0
We assume throughout this section that M ⊂ Rn and M 0 ⊂ Rn are smooth
submanifolds of the same dimension m. As in §5.1 we denote objects on M 0
by the same letters as objects in M with primes affixed. In particular, g 0
denotes the first fundamental form on M 0 and R0 denotes the Riemann
curvature tensor on M 0 .
Let φ : M → M 0 be a diffeomorphism. Using φ we can move objects
on M to M 0 . For example the pushforward of a smooth curve γ : I → M is
the curve
φ∗ γ := φ ◦ γ : I → M 0 ,
the pushforward of a smooth function f : M → R is the function
φ∗ f := f ◦ φ−1 : M 0 → R,
(v) φ∗ R = R0 .
246 CHAPTER 5. CURVATURE
Hence it follows from Lemma 3.6.5 that Γ0 kij = Γkij for all i, j, k. This implies
that the covariant derivative of φ∗ X is given by
m m
X X ∂ψ 0
∇0 (φ∗ X) = ξ˙k + Γkij (c)ċi ξ j k (c)
∂x
k=1 i,j=1
m m
X
ξ˙k +
X ∂ψ
= dφ(ψ(c)) Γkij (c)ċi ξ j k (c)
∂x
k=1 i,j=1
= φ∗ ∇X.
This proves (5.3.1). Equation (5.3.2) follows immediately from (5.3.1) and
Remark 5.2.4.
5.3. GENERALIZED THEOREMA EGREGIUM 247
DX Y := φ∗ (∇φ∗ X φ∗ Y ) .
E(φ ◦ γ) = E(γ)
for every smooth map γ : [0, 1] → M . Hence γ is a critical point of the energy
functional if and only if φ ◦ γ is a critical point of the energy functional.
Alternatively it follows from (ii) that
0 d
∇ φ ◦ γ = ∇0 φ∗ γ̇ = φ∗ ∇γ̇
dt
By (ii) the vector fields X 0 and φ∗ X along φ ◦ γ are both parallel and they
agree at t = t0 . Hence X 0 (t) = φ∗ X(t) for all t ∈ I and this proves (5.3.3).
248 CHAPTER 5. CURVATURE
γ 0 = φ ◦ γ : R2 → M 0 , Z 0 := φ∗ Z ∈ Vect(γ 0 ).
The next corollary establishes the basic relation between local isome-
tries and developments and lies at the heart of the existence criterion for
isometries in the Cartan–Ambrose–Hicks Theorem 6.1.7 in §6.1 below.
Corollary 5.3.4. Let φ : M → M 0 be a local isometry and let γ : I → M be
a smooth curve on an interval I. Fix an element t0 ∈ I and define
p0 := γ(t0 ), q0 := φ(p0 ), Φ0 := dφ(p0 ). (5.3.6)
Then there exists a unique development (Φ, γ, γ 0 ) of M along M 0 on the
entire interval I satisfying the initial conditions
γ 0 (t0 ) = q0 , Φ(t0 ) = Φ0 . (5.3.7)
This development is given by
γ 0 (t) = φ(γ(t)), Φ(t) = dφ(γ(t)) (5.3.8)
for t ∈ I.
250 CHAPTER 5. CURVATURE
Proof. Define
γ 0 (t) := φ(γ(t)), Φ(t) := dφ(γ(t))
for t ∈ I. Then
γ̇ 0 (t) = Φ(t)γ̇(t)
for all t ∈ I by the chain rule, and every vector field X along γ satisfies
Φ∇X = ∇0 (ΦX)
M ⊂ Rm+1 ,
ν(p) ⊥ Tp M, |ν(p)| = 1.
Such a map always exists locally (see Example 3.1.3). Note that ν(p) is
an element of the unit sphere in Rm+1 for every p ∈ M and hence we can
regard ν as a map from M to S m :
ν : M → Sm.
dν(p) : Tp M → Tν(p) S m = Tp M
Here we use the fact that Tν(p) S m = ν(p)⊥ and, by definition of the Gauß
map ν, the tangent space of M at p is also equal to ν(p)⊥ . Thus dν(p) is a
linear map from the tangent space of M at p to itself.
5.3. GENERALIZED THEOREMA EGREGIUM 251
AreaS (ν(B))
|K(p)| = lim .
B→p AreaM (B)
This says that the curvature at p is roughly the ratio of the (m-dimensional)
area of the spherical image ν(B) to the area of B where B is a very small open
neighborhood of p in M . The sign of K(p) is positive when the linear map
dν(p) : Tp M → Tp M preserves orientation and negative when it reverses
orientation.
Remark 5.3.7. We see that the Gaußian curvature is a natural general-
ization of Euler’s curvature for a plane curve. Indeed if M ⊂ R2 is a
1-manifold and p ∈ M , we can choose a curve γ = (x, y) : (−ε, ε) → M such
that γ(0) = p and |γ̇(s)| = 1 for every s. This curve parametrizes M by the
arclength and the unit normal vector pointing to the right with respect to
the orientation of γ is ν(x, y) = (ẏ, −ẋ). This is a local Gauß map and its
derivative (ÿ, −ẍ) is tangent to the curve. The inner product of the latter
with the unit tangent vector γ̇ = (ẋ, ẏ) is the Gaußian curvature. Thus
dx d2 y dy d2 x dθ
K := 2
− 2
=
ds ds ds ds ds
where s is the arclength parameter and θ is the angle made by the normal
(or the tangent) with some constant line. With this convention K is positive
at a left turn and negative at a right turn.
252 CHAPTER 5. CURVATURE
Here the first summand is the second fundamental form, which maps Tp M
to Tp M ⊥ , and the second summand is its dual, which maps Tp M ⊥ to Tp M .
Thus
T
hp (v) = ν(p) dν(p)v : Tp M → Tp M ⊥ ,
hp (u)∗ = dν(p)u ν(p)T : Tp M ⊥ → Tp M.
5.3. GENERALIZED THEOREMA EGREGIUM 253
K = K 0 ◦ φ : M → R.
Proof. That (ii) implies (i) follows directly from the definitions. We prove in
three steps that (i) implies (ii). For p ∈ M and r > 0 denote by Vp ⊂ Tp M
the domain of the exponential map of M at p (Definition 4.3.5), and de-
fine Ur (p) := {q ∈ M | d(p, q) < r} and Br (p) := {v ∈ Tp M | |v| < r}.
Step 1. Assume (5.4.5) holds and choose a real number 0 < r ≤ inj(p0 ; M ).
Then r ≤ inj(q0 ; M ) and φi converges on the open set Ur (p0 ) in the C ∞
topology to the isometry φ0 := expq0 ◦ Φ0 ◦ exp−1
p0 : Ur (p0 ) → Ur (q0 ).
Br (φi (p0 )) = dφi (p0 )Br (p0 ) ⊂ dφi (p0 )Vp0 ⊂ Vφi (p0 )
and φi ◦ expp0 = expφi (p0 ) ◦ Φi : Br (p0 ) → Ur (φi (p0 )). Thus the map
φ0 := expq0 ◦ Φ0 ◦ exp−1
p0 : Ur (p0 ) → expq0 (Br (q0 )).
258 CHAPTER 5. CURVATURE
The map φ0 satisfies d(φ0 (p), φ0 (q)) = limi→∞ d(φi (p), φi (q)) = d(p, q) for
all p, q ∈ Ur (p0 ) and so by Theorem 5.1.1 is a difffeomorphism from Ur (p0 )
onto the open set expq0 (Br (q0 )). Hence the map
expq0 = φ0 ◦ expp0 ◦ Φ−1
0 : Br (q0 ) → expq0 (Br (q0 ))
Namely, the sequence d(φ−1 i (q0 ), p0 ) = d(q0 , φi (p0 )) converges to zero by as-
sumption, and by Step 1 the derivatives dφi converge to dφ uniformly in
some neighborhood U0 ⊂ M of p0 . Since φ−1 i (q0 ) ∈ U0 for i sufficiently
large, this implies that the sequence dφi (φ−1 i (q 0 )) converges to dφ(p0 ) = Φ0
and hence the sequence dφi (φ−1 i (q 0 ))−1 = dφ−1 (q ) converges to Φ−1 . This
i 0 0
proves (5.4.6). By (5.4.6) and Step 2 the sequence φ−1 i converges in the C ∞
topology to a smooth map ψ : M → M . By uniform convergence on com-
pact subsets of M we have ψ ◦ φ = id and φ ◦ ψ = id, so φ : M → M is a
diffeomorphism. Moreover, since φi is an isometry for each i and converges
pointwise to φ, we have d(φ(p), φ(q)) = limi→∞ d(φi (p), φi (q)) = d(p, q) for
all p, q ∈ M , so φ is an isometry. This proves Step 3 and Lemma 5.4.2.
5.4. THE GROUP OF ISOMETRIES* 259
The next goal is to verify that the spaces Ip in (5.4.4) are diffeomorphic
to each other. For p0 , p1 ∈ M define the map Fp1 ,p0 : Ip0 → Ip1 by
Fp1 ,p0 (φ(p0 ), dφ(p0 )) := (φ(p1 ), dφ(p1 )), φ ∈ I(M ). (5.4.7)
This map is well-defined by Lemma 5.1.10, because M is connected. Col-
lectively, these maps give rise to a map F : M × I → I defined by
I := (φ(p), dφ(p), p) p ∈ M, φ ∈ I(M ) ⊂ G,
(5.4.8)
F(p1 , (φ(p0 ), dφ(p0 ), p0 )) := (φ(p1 ), dφ(p1 ), p1 )
for p0 , p1 ∈ M and φ ∈ I(M ). The next lemma uses the concept of a smooth
map on an arbitrary subset of Euclidean space as in the beginning of §2.1.
Lemma 5.4.3. The map F : M × I → I is smooth. In particular, for each
pair of points p0 , p1 ∈ M , the map Fp1 ,p0 : Ip0 → Ip1 is a diffeomorphism.
Proof. Let p0 , p1 ∈ M and choose a smooth path γ : I = [0, 1] → M with
the endpoints γ(0) = p0 and γ(1) = p1 . Let Uγ ⊂ Gp0 be the set of all
pairs (q0 , Φ0 ) ∈ Gp0 such that there exists a development (Φ, γ, γ 0 ) on the
interval I that satisfies
γ 0 (0) = q0 , Φ(t) = Φ0 . (5.4.9)
By Remark 3.5.22, the set Uγ is open in Gp0 and the map Uγ → G that sends
the pair (q0 , Φ0 ) ∈ Uγ to the pair (γ 0 (1), Φ(1)) ∈ Gp1 is smooth. This shows
that there is a unique diffeomorphism
Fγ : Uγ → Uγ −1 , γ −1 (t) := γ(1 − t), Uγ ⊂ Gp0 , Uγ −1 ⊂ Gp1
that satisfies the condition
Fγ (γ 0 (0), Φ(0)) = (γ 0 (1), Φ(1)) (5.4.10)
for every development (Φ, γ, γ 0 )
of M along M on the interval I. The inverse
of Fγ is the smooth map Fγ −1 : Uγ −1 → Uγ . If φ ∈ I(M ), then by Corol-
lary 5.3.4 there exists a development (Φ, γ, γ 0 ) on I satisfying the initial
conditions γ 0 (0) = φ(p0 ) and Φ(0) = dφ(p0 ), and it is given by
γ 0 (t) = φ(γ(t)), Φ(t) = dφ(γ(t)) (5.4.11)
for t ∈ I. Hence (φ(p0 ), dφ(p0 )) ∈ Uγ and by (5.4.10) and (5.4.11) we have
Fγ (φ(p0 ), dφ(p0 )) = (φ(p1 ), dφ(p1 )) = Fp1 ,p0 (φ(p0 ), dφ(p0 ))
for every φ ∈ I(M ). Thus Ip0 ⊂ Uγ and Ip1 ⊂ Uγ −1 and Fγ |Ip0 = Fp1 ,p0 .
Hence Fp1 ,p0 is smooth. The smoothness of F follows from the smooth
dependence of the map Fγ on the curve γ, the verification of which we leave
to the reader. This proves Lemma 5.4.3.
260 CHAPTER 5. CURVATURE
The next lemma shows that the space of Killing vector fields is a finite-
dimensional Lie subalgebra of the Lie algebra Vect(M ) of vector fields on M .
Part (ii) is the linear counterpart of Lemma 5.1.10.
Lemma 5.4.6. Let M ⊂ Rn be a connected smooth m-manifold. Then the
following holds.
(i) If X is a Killing vector field on M and ψ : M → M is an isometry, then
the pullback ψ ∗ X is a Killing vector field. If X and Y are Killing vector
fields on M , then so is their Lie bracket [X, Y ].
(ii) If X is a Killing vector field on M and there exists a p0 ∈ M such that
ψ −1 ◦ φt ◦ ψ : ψ −1 (Dt ) → ψ −1 (D−t )
Now let v ∈ Tp M and define γ(t) := expp (tv) for t ∈ Ip,v . Then
φi (γ(t)) − γ(t) φi (p) − p
X(γ(t)) − X(p) = lim − lim
i→∞ τi i→∞ τi
φi (γ(t)) − φi (γ(0)) − γ(t) + γ(0)
= lim
i→∞ τi
Z t
dφi (γ(s))γ̇(s) − γ̇(s)
= lim ds
i→∞ 0 τi
Z t
= A(γ(s))γ̇(s) ds.
0
Here the last step uses uniform convergence on compact sets in (5.4.23).
Divide by t and use the continuity of the curve t 7→ A(γ(t))γ̇(t) to obtain
1 t X(expp (tv)) − X(p)
Z
A(p)v = lim A(γ(s))γ̇(s) ds = lim .
t→0 t 0 t→0 t
Hence, by Exercise 2.2.16, we deduce that, for all p ∈ M and all v ∈ Tp M ,
A(p)v = dX(p)v.
By (5.4.20) and (5.4.23) this shows that X satisfies (5.4.21) and (5.4.22).
By (5.4.22) and Exercise 2.2.4 we have (X(p), dX(p)) ∈ T(p,1l) Gp and this
implies hv, dX(p)vi = 0 for all v ∈ Tp M . Here is a more direct proof of this
crucial fact. Namely, by (5.4.22) we have
|dφi (p)v − v|2
lim = |dX(p)v|2 .
i→∞ τi2
Hence
hv, dφi (p)v − vi
hv, dX(p)vi = lim
i→∞ τi
hv, dφi (p)vi − |v|2
= lim
i→∞ τi
2hv, dφi (p)vi − |v|2 − |dφi (p)v|2
= lim
i→∞ 2τi
|dφi (p)v − v|2 τi
= − lim
i→∞ τi2 2
=0
for all p ∈ M and all v ∈ Tp M . This shows that X is a Killing vector field
and so it remains to prove that X is complete.
266 CHAPTER 5. CURVATURE
the maximal existence interval for the solution γ of the initial value prob-
lem γ̇(t) = X(γ(t)), γ(0) = p. We prove in three steps that X is complete.
Step 1. Let p ∈ M and let T > 0 such that T |X(p)| < inj(p; M ). Then
[−T, T ] ⊂ Ip .
Define γ(t) := φt (p) for t ∈ Ip . Since X is a Killing vector field, the diffeo-
morphism φt : Dt → D−t is an isometry by Lemma 5.4.4, and hence
for all t ∈ Ip . This implies d(p, γ(t)) ≤ t|X(p)| for all t ∈ Ip , and hence γ
cannot leave the compact set UT |X(p)| (p) = {q ∈ M | d(p, q) ≤ T |X(p)|} on
any subinterval I ⊂ [−T, T ]. Hence [−T, T ] ⊂ Ip by Corollary 2.4.15 and
this proves Step 1.
Step 2. Let p ∈ M . If t ∈ R satisfies |tX(p)| < inj(p; M ) and the sequence
of integers mi ∈ Z is chosen such that
Let T > 0 such that T |X(p)| < inj(p; M ). Then [0, T ] ⊂ Ip by Step 1 and
the set
K := φt (p) 0 ≤ t ≤ T
is compact and |X(q)| = |X(p)| for all q ∈ K. By Lemma 4.2.7 there exists
a constant δ > 0 such that, for all q, q0 , q1 ∈ M ,
d(q0 , q1 )
q ∈ K, d(q, q0 ) ≤ δ, d(q, q1 ) ≤ δ =⇒ ≤ 2. (5.4.25)
|q0 − q1 |
Fix a real number 0 < ε ≤ 1 and choose i0 ∈ N such that, for all i ≥ i0 ,
τi |X(p)| < δ, (T + τi )|X(p)| < inj(p; M ), sup d(q, φi (q)) < δ, (5.4.26)
q∈K
τ
φi (q) − q φ i (q) − q
sup
− X(q) < ε, sup
− X(q) < ε. (5.4.27)
q∈K τi q∈K τi
5.4. THE GROUP OF ISOMETRIES* 267
Then, for all i ≥ i0 and all q ∈ K, we have d(q, φτi (q)) ≤ τi |X(p)| < δ
and d(q, φi (q)) < δ by (5.4.26), and hence by (5.4.25) and (5.4.27),
sup d(φi (q), φτi (q)) ≤ 2 sup |φi (q) − φτi (q)| ≤ 4τi ε. (5.4.28)
q∈K q∈K
Here the last inequality holds for i ≥ i0 by (5.4.28). By induction this implies
d(φm i
i (p), φ
mi τi
(p)) ≤ 4mi τi ε ≤ 4T ε.
for all i ≥ i0 . Hence
lim φm i
i (p) = lim φ
m i τi
(p) = φt (p)
i→∞ i→∞
and this proves Step 2 for t ≥ 0. For t ≤ 0 the argument is similar.
Step 3. Ip = R for every p ∈ M .
Fix an element p ∈ M and real number T > 0 such that T |X(p)| < inj(p; M ).
Choose the sequence mi ∈ N0 such that τi mi ≤ T < (mi + 1)τi . Then Step 2
asserts that φT (p) = limi→∞ φm mi
i (p). Since φi : M → M is an isometry, it
i
mi
follows that T |X(p)| < inj(φi (p); M ) for all i ∈ N and hence
T |X(p)| < inj φT (p); M .
By Step 1 this implies [0, T ] ⊂ IφT (p) . Hence [0, 2T ] ⊂ Ip and, by Step 2,
φ2T (p) = lim φm φT (p) .
i
i→∞ i
Thus [0, ∞) ⊂ Ip and the same argument shows that (−∞, 0] ⊂ Ip . Hence X
is complete. This proves Step 3 and Lemma 5.4.11.
268 CHAPTER 5. CURVATURE
X := t1 Y1 + · · · + tk Yk , (5.4.31)
(q, Φ) := ι tk+1 , . . . , t`
for t1 , . . . , t` ∈ R` . Then
and
dΘ(0) t̂1 , . . . t̂` = t̂1 η1 + · · · + t̂` η`
for t̂1 , . . . , t̂` ∈ R` . Thus the derivative dΘ(0) : R` → T(q0 ,Φ0 ) Gp is a bi-
jective linear map and so the Inverse Function Theorem asserts that the
map Θ restricts to a diffeomorphism from a sufficiently small open neigh-
borhood Ω ⊂ R` of the origin onto its image ι(Ω) ⊂ Gp which is an open
neighborhood in Gp of the point Θ(0) = ιp (φ0 ) ∈ Ip . With this understood,
the assertion that Ip is a smooth submanifold of Gp is a direct consequence
of the following Claim.
Claim. There exists an open set Ω0 ⊂ R` such that
Θ Ω0 ∩ (Rk × {0}) = U0 ∩ Ip ,
0 ∈ Ω0 ⊂ Ω, U0 := Θ(Ω0 ). (5.4.32)
Suppose, by contradiction, that such an open set Ω0 does not exist. Then
there exist sequences ti = t1i , . . . , t`i ∈ R` and φi ∈ I(M ) such that
ti ∈ Ω \ Rk × {0} ,
lim ti = 0, Θ(ti ) = ιp (φi ) ∈ Ip .
i→∞
Define
Then φi ◦ φ−1
0 converges to the identity in the C
∞ topology by Lemma 5.4.2.
Since φi 6= φ0 , we have
|d(φi ◦ φ−1
0 )(p)v − v|
τi := (φi ◦ φ−1
0 )(p) − p + sup >0 (5.4.34)
06=v∈Tp M |v|
for all i by Lemma 5.1.10, and limi→∞ τi = 0. Hence, by Exercise 2.2.4 there
exists a tangent vector (v0 , A0 ) ∈ T(p,1l) Gp and a subsequence, still denoted
by φi , such that, for all v ∈ Tp M ,
(φi ◦ φ−1
0 )(p) − p d(φi ◦ φ−1
0 )(p)v − v
lim = v0 , lim = A0 v. (5.4.35)
i→∞ τi i→∞ τi
It follows from (5.4.35) and Lemma 5.4.11 that there exists a complete
Killing vector field X ∈ VectK,c (M ) such that
X(p) = v0 , dX(p) = A0 .
φi (p) − φ0 (p)
X(φ0 (p)) = lim ,
i→∞ τi
dφi (p)v − dφ0 (p)v
dX(φ0 (p))dφ0 (p)v = lim
i→∞ τi
for v ∈ Tp M . Since by (5.4.33) the sequence
converges to (φ0 (p), dφ0 (p)) = ι(0), the pair (X(φ0 (p)), dX(φ0 (p))dφ0 (p))
belongs to the image of the derivative dι(0) by Exercise 2.2.4, and hence
is a nonzero linear combination of the vectors ηk+1 , . . . , η` . It is also a
linear combination of the vectors η1 , . . . , ηk , because X is a complete Killing
vector field and so is a linear combination of the vector fields Y1 , . . . , Yk .
This is a contradiction, and this contradiction proves the Claim. Thus there
does, after all, exist an open set Ω0 ⊂ R` that satisfies (5.4.32), and the
map Θ−1 : U0 → Ω0 is then a coordinate chart on Gp which satisfies
Proof of Theorem 5.4.1. By Lemma 5.4.3 and Lemma 5.4.12 there exists
a unique smooth structure on I(M ) such that the map ιp : I(M ) → Gp
in (5.4.3) is an embedding for every p ∈ M . That the topology induced
by this smooth structure agrees with the C ∞ topology was established in
Lemma 5.4.2. That it also agrees with the compact open topology (i.e. with
the C 0 topology of uniform convergence on conpact sets) is the content of
Exercise 5.1.11. This proves part (i).
We prove part (ii). Recall the definition of the map
F :M ×I →I
in (5.4.8) and the target map t : G → M , the inverse map i : G → G, and the
composition map m : (s × t)−1 (∆) → G in the beginning of §5.4.1. These
maps are all smooth and turn I into a smooth subgroupoid of G. For
each p ∈ M they endow the submanifold Ip ⊂ Gp with the structure of a Lie
group as follows. The unit is the element
ιp : I(M ) → Ip
ιp (id) = ep ,
mp (ιp (ψ), ιp (φ)) = ιp (ψ ◦ φ), (5.4.39)
−1
ip (ιp (φ)) = ιp (φ )
for all φ, ψ ∈ I(p). The first equation in (5.4.39) follows directly from the
definitions. To prove the remaining equations, fix two isometries φ and ψ
and define ξ, η, ζ ∈ Ip by
Here the last equality follows from the definition of the map mp in (5.4.37).
This proves the second equation in (5.4.39).
To verify the third equation in (5.4.39), we compute
Here the last equality follows from the definition of the map ip in (5.4.38).
This proves the third equation in (5.4.39) and part (ii).
That the Lie algebra of Ip is isomorphic to the space VectK,c (M ) of
complete Killing vector fields under the Lie algebra isomorphism
follows from Lemma 5.4.12. This proves part (iii) and Theorem 5.4.1.
Corollary 5.4.13. Let (p, e) ∈ O(M ). Then there exists a constant ε > 0
with the following significance. If φ : M → M is an isometry that satisfies
then there exists a complete Killing vector field X ∈ VectK,c (M ) whose flow
has the time-1-map ψX = φ.
Proof. Use the construction of the map Θ : Ω → Gp and the Claim in the
proof of Lemma 5.4.12 with φ0 = id to obtain φ ∈ Θ(Ω ∩ (Rk × {0})).
5.4. THE GROUP OF ISOMETRIES* 273
The proof of Theorem 5.4.14 is based on four lemmas. The heart of the
proof is Lemma 5.4.16. It equires the following preparatory result, which
makes use of the Killing form in Example 5.2.24. The case where the Lie
algebra has a nontrivial center is then dealt with in Lemma 5.4.18.
Proof. Part (i) was explained in Exercise 5.2.26. To prove part (ii), assume
that g has a nondegenerate Killing form. Then the map ad : g → Der(g)
is injective because any ξ ∈ ker(ad) satisfies κ(ξ, η) = 0 for all η ∈ g. Now
choose a basis ξ1 , . . . , ξk of g. Since the Killing form is nondegenerate, there
exists a dual basis η1 , . . . , ηk ofP g such that κ(ξi , ηj ) = δij for all i and j.
With these basesPwe have ζ = iκ(ηi , ζ)ξi for all ζ ∈ g. Now let δ ∈ Der(g)
and define ξ := i trace δ ad(ξi ) ηi ∈ g. Then, for all ζ ∈ g,
X
trace δ ad(ζ) = κ(ηi , ζ)trace δ ad(ξi ) = κ(ξ, ζ) = trace ad(ξ)ad(ζ) .
i
Lemma 5.4.17. The assertion of Theorem 5.4.14 holds under the additional
assumption that the center of the Lie algebra g = Lie(G) is trivial.
Proof. Let φ ∈ I0 (G) and choose a smooth isotopy [0, 1] → I0 (G) : t 7→ φt
joining the identity φ0 = id to φ1 = φ. For 0 ≤ t ≤ 1 define the diffeomor-
phism ψt : G → G by ψt (g) := φt (1l)−1 φt (g). Then each ψt is an isometry
and the path [0, 1] → I0 (G) : t 7→ ψt satisfies
ψ0 = id, ψt (1l) = 1l
d
Ψt = Ψt ad(ξt ), Ψ0 = 1l. (5.4.47)
dt
Now let [0, 1] → G : t 7→ bt be the solution of the differential equation
d
bt = bt ξt , b0 = 1l, (5.4.48)
dt
d
and define Φt := Ad(bt ) (Example 2.5.22). Then dt Φt = Φt ad(ξt ) for all t
and Φ0 = 1l. Thus Φt = Ψt and so
Ψt η = bt ηb−1
t (5.4.49)
for all t and η. Take t = 1, define b := b1 , and use Lemma 5.1.10 (uniqueness
of local isometries) to obtain ψ1 (g) = bgb−1 for all g ∈ G. Hence
ḡ := g/Z(g) ∼
= Z(g)⊥
Then it follows from equation (5.4.50) in Lemma 5.4.18 that φ(g) = agb−1
for all g ∈ G. This proves Theorem 5.4.14.
G = Tn = Rn /Zn
Aut(Tn ) = GL(n, Z)
of integer matrices with determinant ±1, while the group of isometries that
fix the origin is the finite subgroup
O(n, Z) ⊂ GL(n, Z)
SO(4) ∼
= SU(2) × SU(2) /{±1l}.
Example 5.4.23. The isometry group of Rm is the group of all affine trans-
formations of Rm with orthogonal linear part (Exercise 5.1.4).
Example 5.4.24. We will see in §6.4.3 that the isometry group of the m-
sphere S m ⊂ Rm+1 is the group
I(S m ) = O(m + 1)
I(Hm ) = O(m, 1)
The preceding three examples are the constant sectional curvature man-
ifolds discussed in §6.2 and §6.4 below. In all three cases the isometry group
has the maximal dimension
m(m + 1)
dim(I(M )) =
2
and is diffeomorphic to the orthonormal frame bundle O(M ), so these exam-
ples satisfy the condition Ip = Gp in the notation of §5.4.1. We will prove in
Corollary 6.4.13 that a complete, connected, simply connected manifold M
satisfies Ip = Gp if and only if it has constant sectional curvature.
280 CHAPTER 5. CURVATURE
M0 := R2 \ {(0, 0)}, M1 := R2 \ Z2 .
I0 (S 2 × S 2 ) ,→ Diff 0 (S 2 × S 2 )
φx,θ : S 2 → S 2
denotes the rotation about the axis through x ∈ S 2 by the angle θ ∈ R/2πZ
and the diffeomorphism ψθ : S 2 × S 2 → S 2 × S 2 is defined by
φ = ψ −1 : U → Ω
∂ψ
Ei (x) := (x) ∈ Tψ(x) M.
∂xi
These vector fields form a basis of Tψ(x) M for every x ∈ Ω and the coeffi-
cients gij : Ω → R of the first fundamental form are
gij = hEi , Ej i .
282 CHAPTER 5. CURVATURE
Recall from Lemma 3.6.5 that the Christoffel Γkij : Ω → R are the coefficients
of the Levi-Civita connection, defined by
m
X
∇i Ej = Γkij Ek
k=1
` : Ω → R and R
Define the coefficients Rijk ijk` : Ω → R of the Riemann cur-
vature tensor by
m
X
`
R(Ei , Ej )Ek = Rijk E` (5.5.1)
`=1
and
m
X
ν
Rijk` := hR(Ei , Ej )Ek , E` i = Rijk gν` . (5.5.2)
ν=1
` ` `
Rijk + Rjki + Rkij = 0, Rijk` + Rjki` + Rkij` = 0. (5.5.5)
Warning: Care must be taken with the ordering of the indices. Some
`
authors use the notation Rkij `
for what we call Rijk and R`kij for what we
call Rijk` .
Exercise 5.5.1. Prove equations (5.5.3), (5.5.4), and (5.5.5). Use (5.5.3)
to give an alternative proof of Theorem 5.3.1.
5.5. CURVATURE IN LOCAL COORDINATES* 283
Gauß
If M ⊂ Rn is a 2-manifold (not necessarily embedded in R3 ), we can use
equation (5.3.9) as the definition of the Gaußian curvature K : M → R.
Let ψ : Ω → U be a local parametrization of an open set U ⊂ M defined on
an open set Ω ⊂ R2 . Denote the coordinates in R2 by (x, y) and define the
functions E, F, G : Ω → R by
∂y F − 12 ∂x G
E F
1 1
K ◦ψ = det F G 2 ∂y G
D2 1 1 1 2 1 2
2 ∂x E ∂x F − 2 ∂y E − 2 ∂y E + ∂x ∂y F − 2 ∂x G
1
E F 2 ∂ y E
1 1
− 2 det F G 2 ∂x G
D 1 1
∂y E 2 ∂x G 0
2
1 ∂ E∂x G − F ∂y E
= − √ √
2 D ∂x E D
1 ∂ 2E∂x F − F ∂x E − E∂y E
+ √ √ .
2 D ∂y E D
This expression simplifies dramatically when F = 0 and we get
1 ∂ ∂x G ∂ ∂y E
K ◦ψ =− √ √ + √ (5.5.6)
2 EG ∂x EG ∂y EG
Exercise 5.5.2. Prove that the Riemannian metric
4
E=G= , F = 0,
(1 + x2 + y 2 )2
6.1.1 Homotopy
Definition 6.1.1. Let M be a manifold and let I = [a, b] be a compact
interval. A (smooth) homotopy of maps from I to M is a smooth map
γ : [0, 1] × I → M . We often write γλ (t) = γ(λ, t) for λ ∈ [0, 1] and t ∈ I
and call γ a (smooth) homotopy between γ0 and γ1 . We say the
285
286 CHAPTER 6. GEOMETRY AND TOPOLOGY
homotopy has fixed endpoints if γλ (a) = γ0 (a) and γλ (b) = γ0 (b) for all
λ ∈ [0, 1]. (See Figure 6.1.)
We remark that a homotopy and a variation are essentially the same
thing, namely a curve of maps (curves). The difference is pedagogical. We
used the word “variation” to describe a curve of maps through a given
map; when we use this word we are going to differentiate the curve to find
a tangent vector (field) to the given map. The word “homotopy” is used
to describe a curve joining two maps; it is a global rather than a local
(infinitesimal) concept.
γ1
M
γ0
γn (t) := e2πint , 0 ≤ t ≤ 1,
p0
M’
Example 6.1.8. Before giving the proof let us interpret the conditions in
case M and M 0 are two-dimensional spheres of radius r and r0 respectively in
three-dimensional Euclidean space R3 . Imagine that the spheres are tangent
at p0 = p00 . Clearly the spheres will be isometric exactly when r = r0 .
Condition (ii) says that if the spheres are rolled along one another without
sliding or twisting, then the endpoint γ 0 (1) of one curve of contact depends
only on the endpoint γ(1) of the other and not on the intervening curve γ(t).
By Theorem 5.3.10 the Riemann curvature of a 2-manifold at p is determined
by the Gaußian curvature K(p); and for spheres we have K(p) = 1/r2 .
288 CHAPTER 6. GEOMETRY AND TOPOLOGY
Exercise 6.1.9. Let γ be the closed curve which bounds an octant as shown
in the diagram for Example 6.1.8. Find γ 0 .
as required.
We prove that (ii) implies (iii) when M 0 is complete. Choose develop-
ments (Φi , γi , γi0 ) for i = 0, 1 as in (iii). Define a curve γ : [0, 1] → M by
“composition”
γ0 (2t), 0 ≤ t ≤ 1/2,
γ(t) :=
γ1 (2 − 2t), 1/2 ≤ t ≤ 1,
so that γ is continuous and piecewise smooth and γ(1) = p0 . By Theo-
rem 3.5.21 there is a development (Φ, γ, γ 0 ) on the interval [0, 1] satisfy-
ing (6.1.2) (because M 0 is complete). Since γ(1) = p0 it follows from (ii)
that
γ 0 (1) = p00 , Φ(1) = Φ0 .
By the uniqueness of developments and the invariance under reparametriza-
tion, we have
for 0 ≤ s ≤ 1 is a development.
To prove (b), choose a vector v ∈ Tp M such that
for every t, by Corollary 5.3.4. Hence it follows from part (iv) of Corol-
lary 5.3.2 (Theorema Egregium for local isometries) that
from γ0 to γ1 with endpoints fixed. By Theorem 3.5.21 there is, for each λ,
a development (Φλ , γλ , γλ0 ) on the interval [0, 1] with initial conditions
To see this we will show that, for each fixed t, the curve
as required.
First choose a basis e1 , . . . , em of Tp0 M and extend it to obtain vector
fields Ei ∈ Vect(γ) along the homotopy γ by imposing the conditions that
the vector fields t 7→ Ei (λ, t) be parallel, i.e.
Then the vectors E1 (λ, t), . . . Em (λ, t) form a basis of Tγλ (t) M for all λ and t.
Second, define the vector fields Ei0 along γ 0 by
Here the second equation follows from (6.1.4) and the fact that Φλ ∂t γ = ∂t γ 0 .
6.1. THE CARTAN–AMBROSE–HICKS THEOREM 291
i=1 i=1
and
∇0t Yi0 = ∇0t ∇0λ Ei0 − ∇0λ ∇0t Ei0 = R0 (∂t γ 0 , ∂λ γ 0 )Ei0 .
To sum up we have X 0 (λ, 0) = Yi0 (λ, 0) = 0 and
m
X
∇0t X 0 = ∂λ ξ i Ei0 + ξ i Yi0 , ∇0t Yi0 = R0 (∂t γ 0 , ∂λ γ 0 )Ei0 .
(6.1.7)
i=1
Remark 6.1.11. The proof of Theorem 6.1.7 shows that the various im-
plications in the weak version of the theorem (where φ is only a local
isometry) require the following conditions on M and M 0 :
(i) always implies (ii), (iii), and (iv);
(ii) implies (iii) whenever M 0 is complete;
(iii) implies (i) whenever M 0 is complete and M is connected;
(iv) implies (iii) whenever M 0 is complete and M is simply connected.
Remark 6.1.12. The proof that (iii) implies (i) in Theorem 6.1.7 can be
slightly shortened by using the following observation. Let φ : M → M 0 be
a map between smooth manifolds. Assume that φ ◦ γ is smooth for every
smooth curve γ : [0, 1] → M . Then φ is smooth.
Proof. This follows by combining the weak and strong versions of the global
C-A-H Theorem 6.1.7. Fix a point p0 ∈ M and define
Then the tuple M, M 0 , p0 , p00 , Φ0 satisfies condition (i) of the weak version
of Theorem 6.1.7. Hence this tuple also satisfies condition (iv) of Theo-
rem 6.1.7. Since M and M 0 are connected, simply connected, and complete
we may apply the strong version of Theorem 6.1.7 to obtain an isometry
ψ : M → M0
satisfying
ψ(p0 ) = p00 , dψ(p0 ) = Φ0 .
Since every isometry is also a local isometry and M is connected it follows
from Lemma 5.1.10 that φ(p) = ψ(p) for all p ∈ M . Hence φ is an isometry,
as required.
Remark 6.1.14. Refining the argument in the proof of Corollary 6.1.13 one
can show that a local isometry φ : M → M 0 must be surjective whenever M
is complete and M 0 is connected. None of these assumptions can be removed.
(Take an isometric embedding of a disc in the plane or an embedding of
a complete space M into a space with two components, one of which is
isometric to M .)
6.1. THE CARTAN–AMBROSE–HICKS THEOREM 293
(iii) If (Φ0 , γ0 , γ00 ) and (Φ1 , γ1 , γ10 ) are developments as in (ii), then
γ(t) := expp0 (tv), γ 0 (t) := exp0p0 (tΦ0 v), Φ(t) := Φ0γ 0 (t, 0)Φ0 Φγ (0, t),
0
Lemma 6.1.16. Let p ∈ M and v, w ∈ Tp M such that |v| < inj(p). For
0 ≤ t ≤ 1 define
∂
γ(t) := exp(tv), X(t) := exp p t(v + λw) ∈ Tγ(t) M.
∂λ λ=0
Then
A vector field along γ satisfying the first equation in (6.1.9) is called a Ja-
cobi field along γ.
Proof. Define
for all λ and t. Since γ(λ, 0) = p for all λ we have X(λ, 0) = 0 and
d
∇t X(λ, 0) = ∇t ∂λ γ(λ, 0) = ∇λ ∂t γ(λ, 0) = v + λw = w.
dλ
Moreover, ∇t ∂t γ = 0 and hence
∇t ∇t X = ∇t ∇t ∂λ γ
= ∇t ∇λ ∂t γ − ∇λ ∇t ∂t γ
= R(∂t γ, ∂λ γ)∂t γ
= R(∂t γ, X)∂t γ.
Proof of Theorem 6.1.15. The proofs (i) =⇒ (ii) =⇒ (iii) =⇒ (i) =⇒ (iv)
are as before; the reader might note that when L(γ) ≤ r we also have
L(γ 0 ) ≤ r for any development so that there are plenty of developments
with γ : [0, 1] → Ur and γ 0 : [0, 1] → Ur0 . The proof that (iv) implies (i) is
a little different since (iv) here is somewhat weaker than (iv) of the global
theorem: the equation Φ∗ R = R0 is only assumed for certain developments.
Hence assume (iv) and define φ : Ur → Ur0 by
γ 0 := φ ◦ γ, dφ(γ)X = X 0 . (6.1.11)
Φ(t)γ̇(t) = γ̇ 0 (t)
for every t. Moreover, it follows from (iv) that Φ∗ Rγ = Rγ0 0 . Combining this
with (6.1.9) we obtain
Hence the vector field ΦX along γ 0 also satisfies the initial value prob-
lem (6.1.12) and thus
ΦX = X 0 = dφ(γ)X.
Here we have also used (6.1.11). Using (6.1.10) we find
and so
dφ(q)u = dφ(γ(1))X(1) = X 0 (1) = Φ(1)u.
Since Φ(1) : Tγ(1) M → Tγ 0 (1) M 0 is an orthogonal transformation this gives
Proof. Assertion (i) follows immediately from Theorem 6.1.15 and (ii) fol-
lows immediately from Theorem 6.1.7.
is flat.
M ⊂ Rm+1
M = {q + tv | q ∈ Γ, t ∈ R}
where H and Γ are as in (i) and v is a fixed vector not parallel to H. (This
is a cone with the cone point p at infinity.)
(iii) The tangent developable to a space curve γ : R → R3 , i.e.
where γ̇(t0 ) and γ̈(t0 ) are linearly independent and ε > 0 is sufficiently small.
(iv) The paper model of a Möbius strip (see Figure 6.3).
x2 y 2
3
M := (x, y, z) ∈ R z = 2 − 2 . (6.2.5)
a b
(both with two straight lines through every point in M ), Plücker’s conoid
3 2
2 2xy
M := (x, y, z) ∈ R x + y 6= 0, z = 2 , (6.2.6)
x + y2
300 CHAPTER 6. GEOMETRY AND TOPOLOGY
the helicoid
( )
x + iy
3 iαz
M := (x, y, z) ∈ R p =e , (6.2.7)
x2 + y 2
and the Möbius strip
cos(s) cos(s/2) cos(s)
t s ∈ R and
M := sin(s) + cos(s/2) sin(s) . (6.2.8)
2 −1 < t < 1
0 sin(s/2)
These five surfaces have negative Gaußian curvature. The Möbius strip
in (6.2.8) is not developable, while the paper model of the Möbius strip is.
The helicoid in (6.2.7) is a minimal surface, i.e. its mean curvature (the
trace of the second fundamental form) vanishes. A minimal surface which
is not ruled is the catenoid
M := {(x, y, z) ∈ R3 | x2 + y 2 = c2 cosh (z/c)}.
(Exercise: Prove all this.)
φ:M →M
satisfying
φ(p) = p, dφ(p) = −id; (6.3.1)
M is called a symmetric space iff it is symmetric about each of its points.
A Riemannian manifold M is called locally symmetric about the point
p ∈ M if, for r > 0 sufficiently small, there exists an isometry
Remark 6.3.2. The proof of Theorem 6.3.4 below will show that, if M is
locally symmetric, the isometry φ : Ur (p, M ) → Ur (p, M ) with φ(p) = p
and dφ(p) = −id exists whenever 0 < r ≤ inj(p).
Corollary 6.3.5. Let M and M 0 be locally symmetric spaces and fix two
points p0 ∈ M and p00 ∈ M 0 , and let Φ0 : Tp0 M → Tp00 M 0 be an orthogonal
linear isomorphism. Let r > 0 be less than the injectivity radius of M at p0
and the injectivity radius of M 0 at p00 . Then the following holds.
(i) There exists an isometry φ : Ur (p0 , M ) → Ur (p00 , M 0 ) with φ(p0 ) = p00
and dφ(p0 ) = Φ0 if and only if Φ0 intertwines R and R0 , i.e.
(Φ0 )∗ Rp0 = Rp0 0 . (6.3.3)
0
such that
(∇R)(X)(X1 , X2 )Y = ∇X R(X1 , X2 )Y − R(∇X X1 , X2 )Y
(6.3.4)
− R(X1 , ∇X X2 )Y − R(X1 , X2 )∇X Y
defined by the right hand side of equation (6.3.4), is multi-linear over the
ring of functions F (M ). Hence it follows as in Remark 5.2.12 that ∇R is
well defined, i.e. that the right hand side of (6.3.4) at p ∈ M depends only
on the tangent vectors X(p), X1 (p), X2 (p), Y (p).
Remark 6.3.10. Let γ : I → M be a smooth curve on an interval I ⊂ R
and
X1 , X2 , Y ∈ Vect(γ)
be smooth vector fields along γ. Then equation (6.3.4) continues to hold
with X replaced by γ̇ and each ∇X on the right hand side replaced by the
covariant derivative of the respective vector field along γ:
(ii) The covariant derivative of the Riemann curvature tensor satisfies the
second Bianchi identity
(∇X R)(Y, Z) + (∇Y R)(Z, X) + (∇Z R)(X, Y ) = 0. (6.3.7)
Proof. We prove (i). Let v1 , v2 , w ∈ Tp M and choose parallel vector fields
X1 , X2 , Y ∈ Vect(γ) along γ satisfying the initial conditions X1 (0) = v1 ,
X2 (0) = v2 , Y (0) = w. Thus
X1 (t) = Φγ (t, 0)v1 , X2 (t) = Φγ (t, 0)v2 , Y (t) = Φγ (t, 0)w.
Then the last three terms on the right vanish in equation (6.3.5) and hence
(∇v R)(v1 , v2 )w = ∇(R(X1 , X2 )Y )(0)
d
= Φ R (X1 (t), X2 (t))Y (t)
dt t=0 γ(0,t) γ(t)
d
= Φ R (Φγ (t, 0)v1 , Φγ (t, 0)v2 )Φγ (t, 0)w
dt t=0 γ(0,t) γ(t)
d
= Φγ (0, t)∗ Rγ(t) (v1 , v2 )w.
dt t=0
Here the second equation follows from Theorem 3.3.6. This proves (i).
We prove (ii). Choose a smooth function γ : R3 → M and denote by
(r, s, t) the coordinates on R3 . If Y is a vector field along γ, we have
(∇∂r γ R)(∂s γ, ∂t γ)Y = ∇r R(∂s γ, ∂t γ)Y − R(∂s γ, ∂t γ)∇r Y
−R(∇r ∂s γ, ∂t γ)Y − R(∂s γ, ∇r ∂t γ)Y
= ∇r (∇s ∇t Y − ∇t ∇s Y ) − (∇s ∇t − ∇t ∇s ) ∇r Y
+R(∂t γ, ∇r ∂s γ)Y − R(∂s γ, ∇t ∂r γ)Y.
Permuting the variables r, s, t cyclically and taking the sum of the resulting
three equations we obtain
(∇∂r γ R)(∂s γ, ∂t γ)Y + (∇∂s γ R)(∂t γ, ∂r γ)Y + (∇∂t γ R)(∂r γ, ∂s γ)Y
= ∇r (∇s ∇t Y − ∇t ∇s Y ) − (∇s ∇t − ∇t ∇s ) ∇r Y
+∇s (∇t ∇r Y − ∇r ∇t Y ) − (∇t ∇r − ∇r ∇t ) ∇s Y
+∇t (∇r ∇s Y − ∇s ∇r Y ) − (∇r ∇s − ∇s ∇r ) ∇t Y.
The terms on the right cancel out. This proves Theorem 6.3.11.
6.3. SYMMETRIC SPACES 305
Proof of Theorem 6.3.4. We prove that (iii) implies (i). This follows from
the local Cartan–Ambrose–Hicks Theorem 6.1.15 with
p00 = p0 = p, Φ0 = −id : Tp M → Tp M.
This isomorphism satisfies
(Φ0 )∗ Rp = Rp .
Hence it follows from the discussion in the beginning of this section that
Φ∗ R = R 0
for every development (Φ, γ, γ 0 ) of M along itself satisfying
γ(0) = γ 0 (0) = p, Φ(0) = −id.
Hence, by the local C-A-H Theorem 6.1.15, there is an isometry
φ : Ur (p, M ) → Ur (p, M )
satisfying
φ(p) = p, dφ(p) = −id
whenever 0 < r < inj(p; M ).
We prove that (i) implies (ii). By Theorem 5.3.1 (Theorema Egregium),
every isometry φ : M → M 0 preserves the Riemann curvature tensor and
covariant differentiation, and hence also the covariant derivative of the Rie-
mann curvature tensor, i.e.
φ∗ (∇R) = ∇0 R0 .
Applying this to the local isometry φ : Ur (p, M ) → Ur (p, M ) we obtain
∇dφ(p)v R φ(p) (dφ(p)v1 , dφ(p)v2 ) = dφ(p) (∇v R) (v1 , v2 )dφ(p)−1
Given i, j, k, ` ∈ {1, . . . , m} we can express the vector field (∇Ei R)(Ej , Ek )E`
along ψ for each x ∈ Ω as a linear combination of the basis vectors Ei (x).
This gives rise to functions
ν
∇i Rjk` :Ω→R
defined by X
ν
(∇Ei R)(Ej , Ek )E` =: ∇i Rjk` Eν . (6.3.8)
ν
These functions are given by
µ
X
ν ν
∇i Rjk` = ∂i Rjk` + Γνiµ Rjk`
µ
(6.3.9)
Γµij Rµk` Γµik Rjµ` Γµi` Rjkµ
X X X
ν ν ν
− − − .
µ µ µ
Example 6.3.15. The flat tori of Exercise 6.2.7 in the previous section are
symmetric (but not simply connected). This shows that the hypothesis of
simple connectivity cannot be dropped in part (ii) of Corollary 6.3.5.
Example 6.3.18. The real projective space RPn with the metric inherited
from S n is a symmetric space and the orthogonal group O(n+1) acts on it by
isometries. The complex projective space CPn with the Fubini–Study metric
in Example 3.7.5 is a symmetric space and the unitary group U(n + 1) acts
on it by isometries. The complex Grassmannian Gk (Cn ) in Example 3.7.6
is a symmetric space and the unitary group U(n) acts on it by isometries.
(Exercise: Prove this.)
This follows from equation (5.3.11) in the proof of Theorem 5.3.10 which
holds in all dimensions. In particular, when M = S m , we have ν(p) = p
and hence K(p, E) = 1 for all p and E. For a sphere of radius r we have
ν(p) = p/r and hence K(p, E) = 1/r2 .
6.4. CONSTANT CURVATURE 309
Example 6.4.4. Let G ⊂ O(n) be a Lie subgroup equipped with the Rie-
mannian metric
hv, wi := trace(v T w)
for v, w ∈ Tg G ⊂ Rn×n . Then, by Example 5.2.17, the sectional curvature
of G at the identity matrix 1l is given by
1
|[ξ, η]|2
K(1l, E) =
4
for every 2-dimensional linear subspace E ⊂ g = Lie(G) = T1l G with an
orthonormal basis ξ, η.
Exercise 6.4.5. Let E ⊂ Tp M be a 2-dimensional linear subspace, let r > 0
be smaller than the injectivity radius of M at p, and let N ⊂ M be the 2-
dimensional submanifold given by
N := expp ({v ∈ E | |v| < r}) .
Show that the sectional curvature K(p, E) of M at (p, E) agrees with the
Gaußian curvature of N at p.
Exercise 6.4.6. Let p ∈ M ⊂ Rn and let E ⊂ Tp M be a 2-dimensional lin-
ear subspace. For r > 0 let L denote the ball of radius r in the (n − m + 2)-
dimensional affine subspace of Rn through p and parallel to the vector sub-
space E + Tp M ⊥ :
n o
L = p + v + w | v ∈ E, w ∈ Tp M ⊥ , |v|2 + |w|2 < r2 .
for all v1 , v2 , v3 , v4 ∈ Tp M .
310 CHAPTER 6. GEOMETRY AND TOPOLOGY
Proof. That (ii) implies (i) follows directly from the definition of the sec-
tional curvature in (6.4.1) by taking v1 = v4 = u and v2 = v3 = v in (6.4.4).
Conversely, assume (i) and define the multi-linear map Q : Tp M 4 → R by
Q(v1 , v2 , v3 , v4 ) := hRp (v1 , v2 )v3 , v4 i − k hv1 , v4 ihv2 , v3 i − hv1 , v3 ihv2 , v4 i .
0 = Q(u1 + u2 , v1 , u1 + u2 , v2 )
= Q(u1 , v1 , u2 , v2 ) + Q(u2 , v1 , u1 , v2 )
for all u1 , u2 , v1 , v2 ∈ T pM . Hence
Q(v1 , v2 , v3 , v4 ) = −Q(v3 , v2 , v1 , v4 )
= Q(v2 , v3 , v1 , v4 )
= −Q(v3 , v1 , v2 , v4 ) − Q(v1 , v2 , v3 , v4 ).
Here the second equation follows from (6.4.5) and the last from (6.4.6). Thus
1 1
Q(v1 , v2 , v3 , v4 ) = − Q(v3 , v1 , v2 , v4 ) = Q(v1 , v3 , v2 , v4 )
2 2
for all v1 , v2 , v3 , v4 ∈ Tp M and, repeating this argument,
1
Q(v1 , v2 , v3 , v4 ) = Q(v1 , v2 , v3 , v4 ).
4
Hence Q ≡ 0 as claimed. This proves Theorem 6.4.8.
6.4. CONSTANT CURVATURE 311
a = id + (12)
c = id + (123) + (132)
b = id − (34)
d = id + (13) + (24) + (13)(24)
φ : Ur (p, M ) → Ur (p0 , M 0 )
such that
φ(p) = p0 , dφ(p) = Φ.
Proof. This follows from Corollary 6.3.5 and Corollary 6.4.10. Alternatively
one can use Theorem 6.4.8 and the local C-A-H Theorem 6.1.15.
312 CHAPTER 6. GEOMETRY AND TOPOLOGY
φ(p) = q, dφ(p) = Φ.
Proof. That (i) implies (ii) follows immediately from Theorem 6.4.8 and the
global C-A-H Theorem 6.1.7. Conversely assume (ii). Then, for every pair
of points p, q ∈ M and every orthogonal linear isomorphism
Φ : Tp M → Tq M,
Φ∗ Rp = Rq
and so
K(p, E) = K(q, ΦE)
for every 2-dimensional linear subspace E ⊂ Tp M . Since, for every pair of
points p, q ∈ M and of 2-dimensional linear subspaces
E ⊂ Tp M, F ⊂ Tq M,
ΦE = F,
M = Rm
with its standard metric is, up to isometry, the unique connected, simply
connected, complete Riemannian m-manifold with constant sectional curva-
ture
k = 0.
M = Sm
with its standard metric is, up to isometry, the unique connected, simply
connected, complete Riemannian m-manifold with constant sectional curva-
ture
k = 1.
Hence, by Corollary 6.4.12, every connected simply connected, complete
Riemannian manifold with positive sectional curvature k = 1 is compact.
Moreover, by Corollary 6.4.13, the isometry group I(S m ) is isomorphic to
the group O(m + 1) of orthogonal linear transformations of Rm+1 . Thus,
by Corollary 6.4.13, the orthonormal frame bundle O(S m ) is diffeomorphic
to O(m + 1). This follows also from the fact that, if
v1 , . . . , v m
p, v1 , . . . , vm
Tp Hm = v ∈ Rm+1 | Q(p, v) = 0 .
gp (v, w) := Q(v, w)
= hξ, ηi − ξ0 η0
(6.4.10)
hξ, xihη, xi
= hξ, ηi −
1 + |x|2
We remark that the manifold Hm does not quite fit into the extrinsic
framework of most of this book as it is not exhibited as a submanifold
of Euclidean space but rather of “pseudo-Euclidean space”: the positive
definite inner product hv, wi of the ambient space Rm+1 is replaced by a
nondegenerate symmetric bilinear form Q(v, w). However, all the theory
developed thus far goes through (reading Q(v, w) for hv, wi) provided we
make the additional hypothesis (true in the example M = Hm ) that the
first fundamental form gp = Q|Tp M is positive definite. For then Q|Tp M is
nondegenerate and we may define the orthogonal projection Π(p) onto Tp M
as before. The next lemma summarizes the basic observations; the proof is
an exercise in linear algebra.
Lemma 6.4.20. Let Q be a symmetric bilinear form on a vector space V
and for each subspace E of V define its orthogonal complement by
E ⊥Q := {u ∈ V | Q(u, v) = 0 ∀v ∈ E} .
V = E ⊕ E ⊥Q ⇐⇒ E ∩ E ⊥Q = {0},
Q(p, p) = −1.
for v ∈ Rm+1 .
316 CHAPTER 6. GEOMETRY AND TOPOLOGY
∇X(t) = Π(γ(t))Ẋ(t)
= Ẋ(t) + Q(Ẋ(t), γ(t))γ(t)
= Ẋ(t) − Q(X(t), γ̇(t))γ(t).
where the first summand on the right is tangent to Hm and the second
summand is Q-orthogonal to Tp Hm . Hence
Ẋ = ∇X + hγ (γ̇)X
extends to the present setting. The reader may verify that the operators
∇ : Vect(γ) → Vect(γ)
thus defined satisfy the axioms of Theorem 3.7.8 and hence define the Levi-
Civita connection on Hm .
6.4. CONSTANT CURVATURE 317
d
Q(γ̇, γ̇) = 2Q(γ̈, γ̇) = 0
dt
because γ̈ is a scalar multiple of γ, and so Q(γ̇, γ̇) is constant. Let p ∈ Hm
and v ∈ Tp Hm be given with Q(v, v) = 1. Then the geodesic γ : R → Hm
with γ(0) = p and γ̇(0) = v is given by
where
et + e−t et − e−t
cosh(t) := , sinh(t) := .
2 2
In fact we have γ̈(t) = γ(t) ⊥Q Tγ(t) Hm . It follows that the geodesics
exist for all time and hence Hm is geodesically complete. Moreover, being
diffeomorphic to Euclidean space, Hm is connected and simply connected.
It remains to prove that Hm has constant sectional curvature k = −1.
To see this we use the Gauß–Codazzi formula in the hyperbolic setting, i.e.
Exercise 6.4.21. Prove that the pullback of the metric on Hm under the
diffeomorphism p
Rm → Hm : x 7→ 1 + |x|2 , x
is given by s
hx, ξi2
|ξ|x = |ξ|2 −
1 + |x|2
or, equivalently, by the metric tensor,
xi xj
gij (x) = δij − (6.4.14)
1 + |x|2
for x = (x1 , . . . , xm ) ∈ Rm .
Exercise 6.4.22. The Poincaré model of hyperbolic space is the open
unit disc Dm ⊂ Rm equipped with the Poincaré metric
2 |η|
|η|y =
1 − |y|2
for y ∈ Dm and η ∈ Rm = Ty Dm . Thus the metric tensor is given by
4δij
gij (y) = 2 , y ∈ Dm . (6.4.15)
2
1 − |y|
ξ(0) = 0, ˙
ξ(0) 6= 0, ξ(t) 6= 0 ∀ t > 0.
ξ(0) = 0, ¨ = 0.
ξ(0)
Prove that there exist constants ε > 0 and c > 0 such that, for all t ∈ R,
Hint: Write
ξ(t) = tv + η(t), ˙ = v + η̇(t)
ξ(t)
with η(t) = O(t3 ) and η̇(t) = O(t2 ). Show that the terms of order 2 and 4
cancel in the Taylor expansion at t = 0.
Proof of Theorem 6.5.2. We prove that (i) implies (ii). Fix a point p ∈ M
and two tangent vectors v, vb ∈ Tp M . Assume without loss of generality
that vb 6= 0 and define the curve γ : R → M and the vector field X ∈ Vect(γ)
along γ by
∂
γ(t) := expp (tv), X(t) := v )) ∈ Tγ(t) M
expp (t(v + sb (6.5.1)
∂s s=0
for t ∈ R. Then
It follows from Exercise 6.5.3 with ξ(t) := Φγ (0, t)X(t) that the function
[0, ∞) → R : t 7→ |X(t)|
is smooth and
d
|X(t)| = |∇X(0)| = |b
v| .
dt t=0
Moreover, for t > 0, we have
d2 d hX, ∇Xi
2
|X| =
dt dt |X|
|∇X|2 + hX, ∇∇Xi hX, ∇Xi2
= −
|X| |X|3 (6.5.4)
|X|2 |∇X|2 − hX, ∇Xi2 hX, R(γ̇, X)γ̇i
= +
|X|3 |X|
≥ 0.
Here the third equality follows from the fact that X is a Jacobi field along γ,
and the inequality follows from the nonpositive sectional curvature condition
in (i) and from the Cauchy–Schwarz inequality. Thus the second derivative
of the function [0, ∞) → R : t 7→ |X(t)| − t |b v | is nonnegative; so its first
derivative is nondecreasing and it vanishes at t = 0; thus
|X(t)| − t |b
v| ≥ 0
X(0) = 0, ∇X(0) = vb 6= 0.
Hence it follows from Exercise 6.5.4 with ξ(t) := Φγ (0, t)X(t) that there is a
constant c > 0 such that, for t > 0 sufficiently small, we have the inequality
X(t)
lim γ̇(t) = v, lim = vb.
t&0 t&0 t
Hence, by (6.5.5) there exist constants δ > 0 and ε > 0 such that
Integrate this inequality over an interval [0, t] with ct2 < εδ 2 to obtain
d d
|X(t)| < |X(t)| = |∇X(0)|
dt dt t=0
Integrating this inequality again gives |X(t)| < t |∇X(0)| for small positive t.
Hence it follows from (6.5.2) that
v = |X(t)| < t |∇X(0)| = t |b
d expp (tv)tb v|
ψ : Rm → M
by
m
!
X
i
ψ(x) := expp x ei
i=1
for x = (x1 , . . . , xm ) ∈ Rm .
Define the metric tensor by
∂ψ ∂ψ
gij (x) := (x), j (x) , i, j = 1, . . . , m.
∂xi ∂x
324 CHAPTER 6. GEOMETRY AND TOPOLOGY
R → Rm : t 7→ tξ
expp : Tp M → M
Here the third inequality follows from (ii). This shows that (ii) implies (iii).
6.5. NONPOSITIVE SECTIONAL CURVATURE 325
q := expp (v).
φ := exp−1
q ◦ expp : Tp M → Tq M.
|φ(v + vb))| = |w| = d(q, expq (w)) = d(expp (v), expp (v + vb)) ≥ |b
v| .
Thus we have proved that (iii) implies (ii) and this completes the proof of
Theorem 6.5.2.
326 CHAPTER 6. GEOMETRY AND TOPOLOGY
Ω ⊂ B(p0 , rΩ ) ∩ B(p1 , rΩ ).
1
m := expp0 2 v0 .
Ω ⊂ B(pΩ , rΩ ).
Ω = gΩ ⊂ B(gpΩ , rΩ )
hS1 , S2 iP := trace S1 P −1 S2 P −1
(6.5.8)
for P ∈ P and S1 , S2 ∈ S = TP P.
g∗ P := gP g T (6.5.9)
P0 := {P ∈ P | det(P ) = 1} (6.5.10)
then there is a constant ε > 0 such that γ(t) ∈ L for |t| < ε.
(ii) If γ : I → L is a smooth curve on an open interval I and Φγ denotes
parallel transport along γ in M , then
∇s ∂t P = ∇t ∂s P
330 CHAPTER 6. GEOMETRY AND TOPOLOGY
for every smooth map R2 → P : (s, t) 7→ P (s, t) and the Leibniz rule
for any two vector fields S1 and S2 along P . These two conditions determine
the covariant derivative uniquely (see Lemma 3.6.5 and Theorem 3.7.8).
Step 2. The geodesics in P are given by
γ(t) = P exp(tP −1 S)
= exp(tSP −1 )P (6.5.12)
1/2 −1/2 −1/2 1/2
=P exp(tP SP )P
R(S, T )A = ∇s ∇t A − ∇t ∇s A
Hence
1 1
R(S, T )A = ∂s ∇t A − (∇t A)P −1 S − SP −1 (∇t A)
2 2
1 1
− ∂t ∇s A + (∇s A)P T + T P −1 (∇s A).
−1
2 2
Now Step 3 follows by a direct calculation which we leave to the reader.
Step 4. The manifold P has nonpositive sectional curvature.
By Step 3 with A = T and equation (6.5.8) we have
= trace SP −1 (RP (S, T )T )P −1
hS, RP (S, T )T iP
1
= − trace SP −1 SP −1 T P −1 T P −1
4
1
− trace SP −1 T P −1 T P −1 SP −1
4
1
+ trace SP −1 T P −1 SP −1 T P −1
4
1
+ trace SP −1 T P −1 SP −1 T P −1
4
1
= − trace SP −1 T P −1 T P −1 SP −1
2
1
+ trace SP −1 T P −1 SP −1 T P −1
2
1 1
= − trace X T X + trace X 2 ,
2 2
where X := P −1/2 SP −1 T P −1/2 . Write X =: (xij )i,j=1,...,n . Then, by the
Cauchy–Schwarz inequality, we have
X X
trace(X 2 ) = xij xji ≤ x2ij = trace(X T X)
i,j i,j
for every matrix X ∈ Rn×n . Hence hS, RP (S, T )T iP ≤ 0 for all P ∈ P and
all S, T ∈ S . This proves Step 4.
Step 5. P is a symmetric space.
Fix an element A ∈ P and define the map φ : P → P by
φ(P ) := AP −1 A
for P ∈ P. This map is a diffeomorphism, fixes the matrix A = φ(A), and
its derivative at P ∈ P is given by
dφ(P )S = −AP −1 SP −1 A for S ∈ S .
332 CHAPTER 6. GEOMETRY AND TOPOLOGY
Ricp : Tp M × Tp M → R
defined by
m
X
Ricp (u, v) := hRp (ei , u)v, ei i, (6.6.1)
i=1
where e1 , . . . , em is an orthonormal basis of Tp M . The Ricci tensor is inde-
pendent of the choice of this orthonormal frame and is symmetric by equa-
tions (5.2.16) and (5.2.18) in Theorem 5.2.13.
for all s ∈ R (see equation (4.1.10)). Differentiate this equation again with
respect to s and use the identity ∇s ∂t = ∇t ∂s to obtain
Z 1
d2
d
E(γ s ) = − h∇t ∂t γs , ∂s γs i dt
ds2 ds s=0 0
Z 1 Z 1
=− h∇s ∇t ∂t γs , ∂s γs i dt − h∇t ∂t γs , ∇s ∂s γs i dt
0 0
Z 1
=− hR(∂s γs , ∂t γs )∂t γs , ∂s γs i dt
0
Z 1 Z 1
− h∇t ∇t ∂s γs , ∂s γs i dt − h∇t ∂t γs , ∇s ∂s γs i dt.
0 0
The example of the m-sphere shows that the estimate in Corollary 6.6.4
is sharp. Namely, M := S m has sectional curvature K = 1 and diameter π.
The paraboloid M := (x, y, z) ∈ R3 | z = x2 + y 2 } has positive Gaußian
curvature and so positive Ricci curvature (Lemma 6.7.2) but is noncompact.
Remark 6.6.5 (Sphere Theorem). The Topological Sphere Theorem
asserts that every complete, connected, simply connected Riemannian m-
manifold M whose sectional curvature satisfies the estimate
1/4 < K(p, E) ≤ 1
for every p ∈ M and every 2-dimensional linear subspace E ⊂ Tp M must be
homeomorphic to the m-sphere. The Differentiable Sphere Theorem
asserts under the same assumptions that M is diffeomorphic to S m .
The problem goes back to a question posed by Heinz Hopf [24, 25] in the
1920s. After intermediate results by Rauch [43] (with 1/4 replaced by 3/4)
and others, the Toplogical Sphere Theorem was proved in 1961 by Berger [5]
and Klingenberg [30]. The Differentiable Sphere Theorem was proved in
2007 by Brendle and Schoen [6, 7, 8]. They even weakened the assumption
to 0 < maxE K(p, E) < 4 minE K(p, E) for all p ∈ M , where the maximum
and minimum are taken over all 2-dimensional linear subspaces E ⊂ Tp M .
The Topological Sphere Theorem is sharp, as the suitably scaled Fubini–
Study metric on complex projective space satisfies 1/4 ≤ K(p, E) ≤ 1 for
all p and E. The Differentiable Sphere Theorem is a significant improve-
ment, because in many dimensions there exist smooth m-manifolds that are
homeomorphic, but not diffeomorphic, to S m . These are the so-called ex-
otic spheres and many of those do not even admit metrics of positive scalar
curvature [23]. (For the definition of scalar curvature see §6.7.)
6.7. SCALAR CURVATURE* 337
Qp : Tp M → Tp M
by X
Qp u := Rp (u, ei )ei
i
for u ∈ Tp M . The right hand side of this equation is independent of the
choice of the orthonormal basis and the endomorphism Qp satusfies
trace(Qp ) = S(p)
and
hQp u, vi = Ricp (u, v)
for all u, v ∈ Tp M . With this notation equation (6.7.5) takes the form
S(p)
Rp (u, v)w = Ricp (v, w)u + hv, wi Qp u − u
2
(6.7.6)
S(p)
− Ricp (u, w)v − hu, wi Qp v − v .
2
Rp (u, v)w = hRp (u, v)w, uiu + hRp (u, v)w, viv
= Ricp (v, w)u − Ricp (u, w)v,
Proof. The skew-symmetry of the Weyl tensor and equation (6.8.2) follow
directly from the definition and Theorem 5.2.13. It then follows from (6.8.2)
that Wp (u, v) is a skew-adjoint endomorphism of Tp M for all u, v ∈ Tp M .
The verification of the Bianchi identity (6.8.3) is a straight forward compu-
tation which we leave as an exercise. To prove (6.8.4), let u, v ∈ Tp M and
choose an orthonormal basis e1 , . . . , em of Tp M . Then
m
X m
X
hWp (ei , u)v, ei i = hRp (ei , u)v, ei i
i=1 i=1
m
1 X
− Ricp (u, v)hei , ei i − Ricp (ei , v)hu, ei i
m−2
i=1
m
1 X
− Ricp (ei , ei )hu, vi − Ricp (u, ei )hei , vi
m−2
i=1
m
S(p) X
+ hei , ei ihu, vi − hu, ei ihei , vi
(m − 1)(m − 2)
i=1
= Ricp (u, v)
m−1
− Ricp (u, v)
m−2
S(p) 1
− hu, vi + Ricp (u, v)
m−2 m−2
S(p)
+ hu, vi
m−2
= 0.
Remark 6.8.4. This discussion shows that the Weyl tensor vanishes for
every Riemannian metric that is locally conformally flat. In fact, it turns
out that in dimension m ≥ 4 the Weyl tensor of M vanishes if and only if
the Riemannian metric on M is locally conformally flat (see [32]). More-
over, a theorem of Kuiper [26, 31] asserts that every compact, connected,
simply connected Riemannian m-manifold that is locally conformally flat is
conformally diffeomorphic to the m-sphere with its constant sectional curva-
ture metric. Thus every compact, connected, simply connected Riemannian
manifold of dimension m ≥ 4 that is not diffeomorphic to S m must have a
nonvanishing Weyl tensor.
Ω2 (M ) = Ω2,+ (M ) ⊕ Ω2,− (M ).
Thus ε = 0 and so
Proof. We prove that (i) implies (ii). Thus assume (i) and choose a positive
orthonormal basis e0 , e1 , e2 , e3 of Tp M . Let τ be the 2-form defined by
τ (e0 , e1 ) := 1, τ (e2 , e3 ) := −1
Notes
Then
(a) f is injective ⇐⇒ it has a left inverse g : Y → X (i.e. g ◦ f = idX );
(b) f is surjective ⇐⇒ it has a right inverse g : Y → X (i.e. f ◦ g = idY );
(c) f is bijective ⇐⇒ it has a two sided inverse f −1 : Y → X.
(Item (b) is the Axiom of Choice.)
351
352 APPENDIX A. NOTES
The analogous principle holds for linear maps: if A ∈ Rm×n , then the
linear map Rn → Rm : x 7→ Ax is
(Here 1lk is the k × k identity matrix.) However, this principle fails com-
pletely for continuous maps: the map f : [0, 2π) → S 1 defined by f (θ) =
(cos θ, sin θ) is continuous and bijective but its inverse is not continuous.
(Here S 1 ⊂ R2 is the unit circle x2 + y 2 = 1.)
The Normal Form Theorem from Linear Algebra says that if A ∈ Rm×n has
rank r, then there are invertible matrices P ∈ Rm×m and Q ∈ Rn×n such
that
−1 1lr 0r×(n−r)
P AQ = .
0(m−r)×r 0(m−r)×(n−r)
By the Fundamental Idea we can expect an analogous theorem for smooth
maps.
for (x, y) ∈ U1 × U2 .
U0 × V0 ⊂ U, g(x0 ) = y0
and
F (x, y) = 0 ⇐⇒ y = g(x)
for x ∈ U0 and y ∈ V0 .
354 APPENDIX A. NOTES
E n × E n → En : (p, q) 7→ p − q,
E n × En → E n : (p, v) 7→ p + v
p + 0 = p, p + (v + w) = (p + v) + w, q + (p − q) = p
E n = o + En
[1] Ralph Herman Abraham & Joel William Robbin, Transversal Mappings and
Flows. Benjamin Press, 1967.
[2] Lars Valerian Ahlfors & Leo Sario, Riemann Surfaces. Princeton University
Press, 1960.
[3] Michael Francis Atiyah & Nigel James Hitchin & Isodore Manuel Singer, Self-
duality in four-dimensional Riemannian geometry. Prococeedings of the Royal
Society of London, Series A 362 (1978), 425–461.
[5] Marcel Berger, Les variétés Riemanniennes 1/4-pincées, Annali della Scuola
Normale Superiore di Pisa 14 (1960) , 161–170.
[6] Simon Brendle, Ricci Flow and the Sphere Theorem. Graduate Studies in
Mathematics 111. AMS, Providence, RI, 2010.
[7] Simon Brendle & Richard Melvin Schoen, Manifolds with 1/4-pinched curva-
ture are space forms. Journal of the American Mathematical Society 22 (2009),
287–307. https://arxiv.org/pdf/0705.0766.pdf
[8] Simon Brendle & Richard Melvin Schoen, Curvature, Sphere Theorems, and
the Ricci Flow. Bulletin of the American Mathematical Society 48 (2011),
1–32. https://arxiv.org/pdf/1001.2278.pdf
[9] Élie Joseph Cartan, La théorie des groupes finis et continus et l’Analysis Situs.
Mémorial des Sciences Mathématiques 42 (1930), 1–61.
357
358 BIBLIOGRAPHY
[11] Simon Kirwan Donaldson, Lie algebra theory without algebra. In “Alge-
bra, Arithmetic, and Geometry, In honour of Yu. I. Manin”, edited by
Yuri Tschinkel and Yuri Zarhin, Progress in Mathematics, Birkhäuser 269,
2009, pp. 249–266. https://arxiv.org/pdf/math/0702016.pdf
[13] David Bernard Alper Epstein, Foliations with all leaves compact. Annales de
l’institute Fourier 26 (1976), 265–282.
[15] Michael Leonidovich Gromov & Herbert Blaine Lawson, The classification of
simply connected manifolds of positive scalar curvature, Annals of Mathemat-
ics 111 (1980), 423–434.
[16] Victor Guillemin & Alan Pollack, Differential Topology. Prentice-Hall, 1974.
[18] Sigurour Helgason, Differential Geometry, Lie Groups, and Symmetric Spaces.
Graduate Studies in Mathematics 35, AMS, 2001.
[19] Noel Justin Hicks, Notes on Differential Geometry, Van Nostrand Rheinhold
Mathematical Studies, 1965.
[20] David Hilbert & Stephan Cohn-Vossen, Geometry and the Imagination.
Chelsea Publishing Company, 1952.
[21] Joachim Hilgert & Karl-Hermann Neeb, Structure and Geometry of Lie groups.
Springer Monographs in Mathematics, Springer-Verlag, 2012.
[26] Ralph Howard, Kuiper’s theorem on conformally flat manifolds. Lecture Notes,
University of South Caroline, Columbia, July 1996. http://ralphhoward.
github.io/SemNotes/Notes/conformal.pdf
BIBLIOGRAPHY 359
[27] Mustafa Kalafat, Locally conformally flat and self-dual structures on simple
4-manifolds. Proceedings of the Gökova Geometry-Topology Conference 2012,
International Press, Somerville, MA, 2013, pp 111–122. http://gokovagt.
org/proceedings/2012/ggt12-kalafat.pdf
[28] John Leroy Kelley, General Topology. Graduate Texts in Mathematics, 27.
Springer 1975.
[29] Wilhelm Karl Joseph Killing, Über die Grundlagen der Geometrie. Journal
für die reine und angewandte Mathematik 109 (1892), 121–186.
[31] Nicolaas Hendrik Kuiper, On conformally flat spaces in the large. Annals of
Mathematics 50 (1949), 916–924.
[35] Anatolij Ivanovich Malcev, On the theory of the Lie groups in the large.
Matematichevskii Sbornik N. S. 16 (1945), 163–189.
Corrections to my paper ”On the theory of Lie groups in the large”. Matem-
atichevskii Sbornik N. S. 19 (1946), 523–524.
[36] Dusa McDuff & Dietmar Arno Salamon, Introduction to Symplectic Topology,
Third Edition. Oxford University Press, 2017.
[37] John Willard Milnor, Topology from the Differentiable Viewpoint. The Uni-
versity Press of Virginia 1969.
[38] James Raymond Munkres, Topology, Second Edition. Prentice Hall, 2000.
[39] Sumner Byron Myers & Norman Earl Steenrod, The group of isometries of a
Riemannian manifold. Annals of Mathematics 40 (1939), 400–416.
[40] John von Neumann, Über die analytischen Eigenschaften von Gruppen lin-
earer Transformationen und ihrer Darstellungen. Mathematische Zeitschrift
30 (1929), 3–42.
[41] Takushiro Ochiai & Tsunero Takahashi, The group of isometries of a left
invariant Riemannian metric on a Lie group. Mathematische Annalen 223
(1976), 91–96.
360 BIBLIOGRAPHY
[42] Roger Penrose, Nonlinear gravitons and curved twistor theory. The riddle of
gravitation – on the occasion of the 60th birthday of Peter G. Bergmann
(Proc. Conf., Syracuse Univ., Syracuse, NY., 1975), General Relativity and
Gravitation 7 (1976), 31–52.
[45] Mary Ellen Rudin, A new proof that metric spaces are paracompact. Proceed-
ings of the American Mathematical Society 20 (1969), 603.
[47] Dietmar Arno Salamon, Analysis II. Lecture Course, ETHZ, FS 2017.
https://people.math.ethz.ch/~salamon/PREPRINTS/analysis2.pdf
[48] Dietmar Arno Salamon, Spin Geometry and Seiberg–Witten invariants. Un-
published Manuscript, 1999. https://people.math.ethz.ch/~salamond/
PREPRINTS/witsei.pdf
[51] Dietmar Arno Salamon & Thomas Walpuski, Notes on the Octonions. Pro-
ceedings of the 23rd Gökova Geometry-Topology Conference 2016, edited
by S. Akbulut, D. Auroux, and T. Önder, International Press, Somerville,
Massachusetts, 2017, pp 1-85. https://people.math.ethz.ch/~salamon/
PREPRINTS/Octonions.pdf
[53] Richard Melvin Schoen & Shing-Tung Yau, On the structure of manifolds with
positive scalar curvature. Manuscripta Mathematica 28 (1979), 159–183.
[57] Neil Sidney Trudinger, Remarks concerning the conformal deformation of Rie-
mannian structures on compact manifolds. Annali della Scuola Normale Su-
periore di Pisa 22 (1968), 265–274.
362
INDEX 363