You are on page 1of 75

RIEMANNIAN GEOMETRY

MARC BURGER
STEPHAN TORNIER

Abstract. These are notes of the course Differential Geometry II held at


ETH Zurich in 2016.

Disclaimer. This is a preliminary version. Please report any typos, mistakes,


comments etc. to stephan.tornier@math.ethz.ch.

Acknowledgements. Thanks to those who pointed out typos and mistakes


as these notes were written, in particular S. Ammann, V. Junet, W. Liu, M.
Wasem and J. Wettstein.

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1. Review of Differential Geometry . . . . . . . . . . . . . . . . . . . . . 4
2. Riemannian Metrics, Covariant Derivative and Geodesics . . . . . . . 28
3. Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4. What’s Beyond . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

Introduction
In this introduction, we outline how Bernhard Riemann resolved three important
problems of his time with the definition of what we now call a Riemannian manifold
in his habilitation treatise [Rie54]. The three problems revolve around how to deal
with the geometry of curves and surfaces after Gauss, Euclid’s fifth postulate and
manifolds which “can’t be seen”. This is a largely informal section. Its notation is
deliberately old-fashioned and should not be taken too seriously.
Given a curve − →c : R → E2 , t 7→ −→c (t) in Euclidean space and points a, b ∈ R
with a < b one can talk about the length of the segment [− →c (a), −

c (b)]:
E2
Z b
length(−
→ k−

b

c [a, b]) = c ′ (t)k dt.


b
−c
→ a

This defines the intrinsic metric of −



c . Also, a curve −

c with − →c ′ (t) 6= 0 for all


t ∈ R can be reparametrized which is really saying that c (R) with its intrinsic
metric is isomorphic to R. This already finishes the metric classification of curves.
However, there are interesting invariants describing how the curve sits in Eu-
clidean space such as curvature: A circle of radius r is defined to have constant
curvature 1/r. Given a curve − →c : R → E2 and an orientation of − → c via normal

− 2 →
− →

vectors n : R → E the curvature of c at the point c (t) is defined as
±1
K(−

c (t)) =
radius of osculating circle at →
−c (t)

Date: June 2, 2016.


1
2 MARC BURGER STEPHAN TORNIER

where the sign depends on whether − →


n (t) points towards or away from the osculating
circle which is the unique circle tangent of order three at −→
c (t).

E2

In other words, let t1 ≤ t ≤ t2 . Then if −



c (t1 ), −

c (t) and −

c (t2 ) are not collinear,
there is a unique circle through the three of them. Take the limit as t1 and t2 tend
to t from below and above respectively.
Given a surface S in E3 and points p and q on S the intrinsic distance of p and
q is defined as

q
b

p
  
 →
− 
  c ([0, 1]) ⊆ S 

c ) −
dS (p, q) = inf length(−
→ c : [0, 1] → E3 smooth : →
−c (0) = p

 
−
→ 

c (1) = q
One may then ask what the shortest path between p and q is. This question
was studied at least since the work of the Bernoulli’s on variational questions. The
study of the distance on S and the determination of shortest paths – geodesics –
amount to the intrinsic geometry of S.
A surface S as above, or rather a piece of it, can – being two-dimensional – be
parametrized by two real parameters, say u and v. A curve − →c (t) lying on S can


then be given by c (t) = (x(u(t), v(t)), y(u(t), v(t)), z(u(t), v(t)))T and we compute
k−
→c ′ (t)k2 = E u′ (t)2 + 2F u′ (t)v ′ (t) + G v ′ (t)2 .
The expression ds2 = E du2 + 2F dudv + G dv 2 was used to refer to the first
fundamental form. This notation should not be confused with notation from the
theory of differential forms but rather be granted as moonshine from the 18th
century. The point is that the inner distance on S determines and is determined by
the first fundamental form.
What can be said about curvature in the case of surfaces? The mathematician’s
way to deal with it is to reduce the question to the case of curves discussed above.
Let m ∈ S and let − →
n denote a normal vector to S at p. Further, let − →u ∈ Tm S be
a unit tangent vector and let P denote the plane spanned by n and −

− →
u . Then the
intersection of S and P is a curve −
→c and we define
Π : T S → R, −
m
→u 7→ K(−→
c at m).
The map Π assumes both a minimum K1 (m) and a maximum K2 (m) which may
take arbitrary signs. Gauss defined curvature to be the product of these two prin-
cipal curvatures: K(m) := K1 (m)K2 (m). This makes sense geometrically:
RIEMANNIAN GEOMETRY 3

mb

m
b

m
K1 (m), K2 (m) > 0 K1 (m), K2 (m) < 0 K1 (m) < 0, K2 (m) > 0
He proved the following Theorema Egregium.
Theorem (Gauss). The curvature KS of a surface S only depends on the intrinsic
geometry of S. It admits a formula in terms of the coefficients E, F and G of the
first fundamental form.
He was also conscious of the fact that this notion of curvature assumes that we
can see the surface under consideration inside Euclidean space E3 and hence does
not lend itself to investigate e.g. projective space P2 (R) which is defined as
P2 R := R3 \{0}/ ∼
where x ∼ y if and only if there is λ ∈ R \{0} such that x = λy. The Klein bottle
is another example.
We now turn to Euclid’s fifth postulate, the parallel postulate. Euclid’s approach
to the geometry was to define the objects under consideration, namely points and
lines, and to postulate “self-evident” statements, such as that there should be a
unique line going through two distinct points, with the help of which further objects
could be introduced and statements made. His fifth postulate stated that given a
line l and a point P not on l there is a unique line through P not intersecting l.
In 1829, Lobatchevsky constructed a geometry that did not satisfy Euclid’s fifth
postulate. In fact Gauss had already known about this but had not published
anything as he feared his reputation.
Finally, Bernhard Riemann introduced n-dimensional manifolds and their tan-
gent spaces in his habilitation treatise in 1854. He suggested to smoothly put a
scalar product h−, −im on each tangent space Tm M of a manifold of M which
allows one to define the notion of length of a curve in M and geodesics. He also
introduced what is now called the Riemann curvature tensor, a revolutionary way
to measure curvature. It is related to sectional curvature, Ricci curvature and scalar
curvature. Sectional curvature is a curvature notion that generalizes Gauss curva-
ture: Let P ⊆ Tm M be a two-dimensional vector subspace. Given ε > 0, consider
the “circle” CP (m, ε) of points in M that are reached from m by following the ge-
odesic flow along unit tangent vectors in P . Then the length of the closed curve
CP (m, ε) is given by
π
length(CP (m, ε)) = 2πε − σm (P )ε3 + O(ε4 )
3
where σm (P ) is the sectional curvature which, as a consequence, is an intrinsic
object. By the way, note that it is a big mystery why there never is a term in ε2 .
Riemann Pended his habilitation treatise with the example of the open unit disk
n
{x ∈ Rn | i=1 x2i < 1} in Rn on which he defines the metric
dx21 + · · · + dx2n
ds2 = P 2 .
(1 − ni=1 x2i )
This coincides exactly with Lobatchevsky’s non-Euclidean plane. Overall, Riemann’s
habilitation treatise dealt with many problems in one stroke.
4 MARC BURGER STEPHAN TORNIER

Course Outline. In this course, we will define Riemannian metrics on smooth


manifolds and use them to study geodesics. We also study derivates of vector fields
with respect to each other, leading to the notion of connection. In general, there
are many possible connections. However, on a Riemannian manifold there is a pre-
ferred one, the Levi-Civita connection. Using the framework of connections we will
extremely efficiently identify geodesics as curves whose acceleration vanishes. After
introducing various notions of curvature we move on to the relation between local
curvature properties of a Riemannian manifold and its global properties, e.g. prop-
erties of its fundamental group or de Rham cohomology spaces. For instance, there
is the following classical theorem.
Theorem (Berger-Klingenberg). Let M be a simply-connected smooth manifold and
assume that 1/4 < σm (P ) ≤ 1 for all m ∈ M and P ≤ Tm M be two-dimensional.
Then M is homeomorphic to S n where n = dim M .
In the context of dynamical systems and ergodic theory we shall see that the
negative curvature world is very different from the above.

1. Review of Differential Geometry


In this chapter, we recall some fundamentals from differential geometry. When-
ever possible, we refer to the lecture notes [BT15] of the Differential Geometry I
course of the fall semester 2015. The main new topics include a more thorough
treatment of vector fields, basic concepts of vector bundles, a short section on Lie
groups and one on covering maps and fibrations.

1.1. Smooth Manifolds and Smooth Maps. Recall that a topological manifold
of dimension n is a topological space M which is Hausdorff, second-countable and
locally homeomorphic to Rn . ASsmooth atlas A on M is a collection of charts A =
{(Uα , ϕα ) | α ∈ A} such that α∈A Uα = M and all coordinate transformations
θβα = ϕβ ◦ ϕ−1 α : ϕα (Uα ∩ Uβ ) → ϕβ (Uα ∩ Uβ ) is smooth.

Uα Uβ

ϕα ϕβ

ϕα (Uα ) θβα ϕβ (Uβ )

θαβ

One shows that every smooth atlas is contained in a unique maximal one. A
smooth manifold of dimension n is a topological n-manifold together with a max-
imal smooth atlas. Recall that an exponent of a manifold typically indicates its
dimension. Let M m and N n be smooth manifolds and let f : M → N be a map.
We say that f is differentiable at p ∈ M if there are charts (U, ϕ) at p ∈ M and
(V, ψ) at f (p) such that f (U ) ⊆ V and ψf ϕ−1 : ϕ(U ) → ψ(V ) is differentiable at
ϕ(p). Further, f is smooth if it is continuous and if for all charts (U, ϕ) of M and
(V, ψ) of N satisfying f (U ) ⊆ V the map ϕ(U ) → ψ(V ) is smooth. Note that the
continuity assumption in the last definition is crucial as a map that is wild to the
extent that there are no pairs of charts as above should not be termed smooth.
Given a map f : M → N that is differentiable at p ∈ M the rank rankp (f ) of f at
RIEMANNIAN GEOMETRY 5

p is defined to be the rank of the linear map Dp f : Tp M → Tf (p) N . We further


recall that a smooth map f : M m → N n is an immersion if rankp (f ) = m for all
p ∈ M , an submersion if rankp (f ) = n for all p ∈ M and an embedding if it is an
immersion and a homeomorphism onto its image. In the last situation, f (N ) is a
regular submanifold and f is a diffeomorphism onto its image.
An important tool to construct manifolds is the following.
Theorem 1.1. Let M m and N n be smooth manifolds and let f : M m → N n be a
smooth map of constant rank k. Then for any y ∈ f (M ) the subset f −1 (y) ⊆ M is
a regular submanifold of dimension m − k.
By applying Theorem 1.1 in the case where M and N are vector spaces and f is
a linear map, note that it is a non-linear version of the first isomorphism theorem
from linear algebra.
Example 1.2. (Orthogonal Groups). Let q : Rn → R be a quadratic form. Then there
is a matrix A ∈ Sym(n, R) such that q(x) = xT Ax. Recall Silvester’s classification
of quadratic forms via their signature and assume from now on that q is non-
degenerate, equivalently, that A ∈ GL(n, R). For example, if n = 2 we may have
q(x1 , x2 )T = x21 − x22 . Notice that q producing vectors of length zero says nothing
about it being degenerate or non-degenerate. The orthogonal group of q is
O(q) := {h ∈ GL(n, R) | q(hx) = q(x) ∀x ∈ Rn }
= {h ∈ GL(n, R) | hT Ah = A}
Also recall that O(q) is compact if and only if q is either positive or negative definite.
Regardless of the signature of q, as long as it is non-degenerate, O(q) has dimension
n(n − 1)/2 as can be seen using Theorem 1.1: Namely, consider the map
f : GL(n, R) → GL(n, R), g 7→ g T Ag.
In order to compute the derivative of f we recall that GL(n, R) acquires its manifold
structure as an open subset of Mn,n (R) ∼ = Rn·n . As such we identify Tg GL(n, R)
with Mn,n (R) for every g ∈ GL(n, R). We compute
(Dg f )(Y ) = Y T Ag + g T AY
for all Y ∈ Tg GL(n, R). Observe, that the map Dg f takes values in Sym(n, R). In
fact every S ∈ Sym(n, R) is in the image of Dg f as
 
1 −1 −1 T
Dg f A (g ) S = S.
2
Consequently, rank(Dg f ) = n(n + 1)/2 and the constant rank theorem implies that
O(q) = f −1 (A) is a submanifold of dimension n2 − n(n + 1)/2 = n(n − 1)/2.
The case in which q is degenerate is left to the reader.
Example 1.3. (Symplectic Group). Let V be a finite-dimensional real vector space
and let ω : V × V → R be an alternating bilinear 2-form. If ω is non-degenerate, i.e.
{x ∈ V | ω(x, y) = {0} ∀y ∈ V } = 0, it is called symplectic. In this case, dim V = 2n
and there is a basis {e1 , . . . , en , f1 , . . . , fn } such that ω(x, y) = xT Jy where
 
0 Id
J=
− Id 0
Note the sharp contrast to the variety of quadratic forms. The symplectic group is
given by
Sp(2n, R) = {g ∈ GL(2n, R) | ω(gx, gy) = ω(x, y) ∀x, y ∈ R2n }
= {g ∈ GL(2n, R) | g T Jg = J}.
6 MARC BURGER STEPHAN TORNIER

It is an exercise to show that Sp(2n, R) is a submanifold of GL(n, R) of dimension


n(2n + 1).

We shall see later that admitting a a symplectic 2-form is an interesting invariant


for manifolds, in contrast to being Riemannian.

1.2. Tangent Bundle. Given a smooth manifold M m and p ∈ M , the tangent


space Tp M of M at p is defined as the quotient of
Ap := {(U, ϕ, ξ) | (U, ϕ) is a chart at p, ξ ∈ Rn }
by the equivalence relation
(U, ϕ, ξ) ∼p (V, ψ, η) ⇔ Dϕ(p) (ψϕ−1 )ξ = η.
In this way, Tp M := Ap / ∼p is an m-dimensional vector space canonically attached
to a point p ∈ M .

U V

ϕ ψ

ϕ(U ) ψ ◦ ϕ−1 η ψ(V )


b b

Note that every chart (U, ϕ) of M at p gives rise to a vector space isomorphism
θ(U,ϕ,p) : Rm → Tp M by setting ξ 7→ [(U, ϕ, ξ)].
If f : M → N is a map which is differentiable at p ∈ M then there is a canonical
linear map Dp f : Tp M → Tf (p) N such that whenever (U, ϕ) is a chart at p and
(V, ψ) is a chart at f (p) with f (U ) ⊆ V , the diagram
θ(U,ϕ,p)
Rm / Tp M

Dϕ(p) ψf ϕ−1 Dp f
 
Rm /T N
θ(V,ψ,f (p))f (p)

commutes. There is an alternative, more classical and convenient way to define


tangent vectors. It fits into the paradigm of reducing new problems to the case of
curves. Consider the set Cp of pairs (I, c) consisting of an open interval I ⊆ R
containing 0 ∈ R and smooth curve c : I → M with c(0) = p. We declare (I, c) ∼p
(J, γ) if there is a chart (U, ϕ) at p such that (ϕ ◦ c)′ (0) = (ϕ ◦ γ)′ (0).

Lemma 1.4. Let M be a smooth manifold and let p ∈ M . Further, let (U, ϕ) be a
chart of M at p. Then the map
Cp → Ap , (I, c) 7→ (U, ϕ, (ϕ ◦ c)′ (0))
induces a well-defined bijection Cp / ∼p → Ap / ∼p .

The verification of Lemma 1.4 is left to the reader.


RIEMANNIAN GEOMETRY 7

1.2.1. Tangent bundle. Finally, we now turn to the tangent bundle, see [BT15, Sec.
2.3]. As a set, the tangent bundle of a smooth manifold M is defined as
[
TM := {x} × Tx M = {(x, v) | x ∈ M, v ∈ Tx M }.
x∈M

It comes with the canonical projection map π : TM → M, (x, v) 7→ x. Given an


open set U ⊆ M we set TU = π −1 (U ). Further, if (U, ϕ) is a chart of M , we define
the map Dϕ : TU → ϕ(U ) × Rn by (x, v) 7→ (ϕ(x), (Dx ϕ)v).
Lemma 1.5. Let M m be a smooth manifold. Then TM admits a topological 2m-
manifold structure for which {(TU, Dϕ) | (U, ϕ) chart} is a smooth atlas.
1.2.2. Vector bundles. The tangent bundle is an example of the more general notion
of vector bundle which we introduce now.
Definition 1.6. A vector bundle is a triple (π, E, B) consisting of a smooth map
π : E → B of smooth manifolds such that
(i) π is surjective,
(ii) there is an open cover (Ui )i∈I of B and a collection of diffeomorphisms hi
(i ∈ I)
hi : π −1 (Ui ) → Ui × Rn
such that hi (π −1 (x)) = {x} × Rn , and
(iii) for all i, j ∈ I, the map hi (hj |Ui ∩Uj )−1 : (Ui ∩ Uj ) × Rn → (Ui ∩ Uj ) × Rn
given by (x, v) 7→ (x, gij (x)v), where gij : Ui ∩ Uj → GL(n, R), is smooth.
It is an exercise to show that the tangent bundle of a manifold is in fact a vector
bundle in the sense of Definition 1.6. Part (iii) of said definition entails that each
fiber π −1 (x) (x ∈ B) has a well-defined vector space structure.
Example 1.7. As another example of vector bundles, recall the Grassmannian
G(k, n) := {L ⊆ Rn | dim L = k} and define
E(k, n) = {(L, v) | L ∈ G(k, n), v ∈ L}
In a sense to be made precise, these vector bundles are universal.
Definition 1.8. Let (π, E, B) be a vector bundle.
(i) A (smooth) section of (π, E, B) is a (smooth) map s : B → E such that
π ◦ s = idB .
(ii) The vector bundle (π, E, B) is trivial if there is a diffeomorphism h : E →
B×Rn such that h|π−1 (x) : π −1 (x) → {x}×Rn is a vector space isomorphism
for every x ∈ B.
Note that for any vector bundle, the zero section is smooth.

1.3. Vector Fields.


Definition 1.9. Let M be a smooth manifold. A (smooth) vector field on M is a
(smooth) section X : M → TM of the tangent bundle.
Let Γ(TM ) denote the span of all smooth vector fields on M . This space is not
only a real vector space but in fact a module over C ∞ (M ): Given f ∈ C ∞ (M ) and
X ∈ Γ(TM ) we define (f X)(x) := f (x)X(x) for all x ∈ M .
Proposition 1.10. Let M be a smooth manifold. Then TM is trivial if and only if
there are smooth vector fields X1 , . . . , Xm on M such that (X1 (x), . . . , Xm (x)) is a
basis of Tx M for all x ∈ M .
8 MARC BURGER STEPHAN TORNIER

A manifold is parallelizable if its tangent bundle is trivial. When mathemati-


cians first looked at this notion they were surprised to find that there are non-
parallelizable manifolds, for instance almost all spheres.
Example 1.11.
(i) It is evident that Rn is parallelizable for all n.
(ii) We now show that also all tori T n := (S 1 )n are parallelizable. To this end,
we view S 1 as the abelian group S 1 := {z ∈ C | |z| = 1}. Recall that
for every ̺ ∈ T n the left multiplication map L̺ : T n → T n , z 7→ ̺z is a
diffeomorphism. A trivialization of T(T n ) is thus given by
T(T n ) → T n × T1 (T n ), (̺, v) 7→ (̺, D̺ L̺−1 (v))
This argument applies in the more general context of Lie groups.
(iii) As to the spheres, it is a fact that S n is parallelizable if and only if
n ∈ {1, 3, 7}. In these cases, the respective sphere can be identified as the
unit length elements in a real division algebra, namely either the complex
numbers, the quaternions or the octonions. The case of S 1 is covered above.
The key parallelizability does not lie in a manifold’s fundamental group but in
its higher-dimensional fundamental groups. Given Proposition 1.10, the following
theorem shows that all even-dimensional spheres, except S 0 , are not parallelizable.
The proof we give is due to Milnor and is not very telling but nicely uses the theory
developed in part one of this course.
Theorem 1.12. The sphere S n admits an everywhere non-zero smooth vector field
if and only if n is odd.
Proof. We first show that the condition is sufficient: If n = 2m−1 we may construct
a nowhere vanishing vector field as follows: Exhibit
( )
Xm

S 2m−1
= (x1 , y1 ), . . . , (xm , ym ) (xi + yi ) = 1 ⊆ R2m .
2 2

i=1
Now, given the rotation
 
cos t sin t
r(t) = ∈ SO(2)
− sin t cos t
define R(t) := r(t) ⊕ · · · ⊕ r(t) acting on R2m . Then R(t)(S 2m−1 ) ⊆ S 2m−1 and we
may set
d
X(z) := {t 7→ R(t)z}
dt t=0
for all z ∈ S 2m−1 . One verifies that hX(z), zi = 0 and kX(z)k = 1. Therefore, X is
a unit vector field on S 2m−1 .
To show necessity we argue by contradiction. Assume that there is a smooth
map X : S n → Rn+1 such that hX(x), xi = 0 for all x ∈ S n . Multiplying with an
appropriate smooth function we may assume that kX(x)k = 1 for all x ∈ S n . Now,
let ε > 0 and define fε : S n → Rn+1 by fε (x) := x + εX(x). Then fε (x) has norm

1 + ε2 for every x ∈ X since
hfε (x), fε (x)i = hx + εX(x), x + εX(x)i = hx, xi + ε2 hX(x), X(x)i = 1 + ε2 .

Overall, fε is a smooth map taking values in S n ( 1 + ε2 ). In order to be able to
talk about the degree of fε we endow S n (r) (r > 0) with the orientation it obtains
as boundary of the regular domain B n+1 (r) = {x ∈ Rn+1 | kxk ≤ r}.
In fact,√deg fε = 1 which can be seen as follows: Consider the map projection
Mε : S n ( 1 + ε2 ) → S n given by y 7→ y/kyk which is an orientation-preserving
RIEMANNIAN GEOMETRY 9


diffeomorphism. Precomposing it with fε yields Mε ◦fε (x) = (x+εX(x))/( 1 + ε2 )
which is homotopic to M0 f0 = IdS n . Hence deg Mε ◦ fε = 1 and thus deg fε = 1.
To obtain the announced contradiction, consider the differential form
n+1
X
ω= ci ∧ · · · ∧ dxn+1 ∈ Ωn (Rn+1 ) ∈ Ωn (Rn+1 ).
(−1)i xi x1 ∧ · · · ∧ dx
i=1

For instance, if n = 3, then ω = x1 dx2 ∧ dx3 − x2 dx1 ∧ dx3 + x3 dx1 ∧ dx2 . Observe
ci ∧ · · · ∧ dxn+1 ) = dxi ∧ dx1 ∧ · · · ∧ dx
d(xi dx1 ∧ · · · ∧ dx ci ∧ · · · ∧ dxn+1
= (−1)i−1 dx1 ∧ · · · ∧ dxn+1 .
Hence dω = (n + 1)dx1 ∧ · · · ∧ dxn+1 . Let ω|S n (r) denote the pullback of ω via the
injection S n (r) → Rn+1 . On the one hand, Stokes’ Theorem yields
Z Z
ω|S n (r) = dω = (n + 1) vol(B n+1 (r)) = (n + 1)cn rn+1
S n (r) B n+1 (r)

where cn is a constant depending solely on n. In particular, for r = 1 + ε2 we get
Z
n+1
ω|S n (r) = (n + 1)cn (1 + ε2 ) 2 .
S n (r)

On the other hand, we have


Z Z Z
n+1
(fε )∗ (ω|S n (r) ) = deg fε ω|S n (r) = ω|S n (r) = (n + 1)cn (1 + ε2 ) 2 .
S n (1) S n (r) S n (r)

Regarding the leftmost expression, note that fε∗ (ω|S n (r) ) is a differential form on
S n (1) which depends on ε. By definition
(fε∗ ω)x (v1 , . . . , vn ) = ωfε (x) (Dx fε v1 , . . . , Dx fε vn )
which is a sum over an index i ranging between 1 and n + 1 in which the i-th
summand is of the form
ci ∧ · · · ∧ dxn+1 )((Id +εDx X)v1 , . . . , (Id +εDx X)vn+1 )
(xi + εXi (x))(dx1 ∧ · · · dx
P
On this expression, one sees that (fε∗ ω|S n (r) ) = nj=0 εj ηj where ηj ∈ Ωn (S n ). We
obtain
Z n
X Z
∗ j
(fε )(ω|S n (r) ) = ε ηj
S n (1) j=0 S n (1)

which is a polynomial in ε. On the other hand, if n is even then (1 + ε2 )(n+1)/2 is


not a polynomial. 

There is a shorter proof of Theorem 1.12, see e.g. [Lee10, Thm. 13.32], using a
nowhere-vanishing vector field to build a homotopy between the identity and the
antipodal map.

1.3.1. Vector fields and derivations. Given a manifold M and a point p ∈ M ,


recall that a derivation at p is a linear map λ : C ∞ (M ) → R which satisfies
λ(f g) = f (p)λ(g) + g(p)λ(f ). Let Derp (C ∞ (M )) denote the space of all derivations
at p ∈ M . We have shown the following.
Proposition 1.13. Retain the above notation. The map
Tp M → Derp (C ∞ (M )), v 7→ {f 7→ (Dp f )(v)}
is a vector space isomorphism.
10 MARC BURGER STEPHAN TORNIER

In what is to follow we will often abuse notation by identifying a tangent vector


with a derivation and vice-versa.
In order to express vector fields in local coordinates, let (U, ϕ) be a chart of M
and p ∈ U . For f ∈ C ∞ (M ), define
∂(f ◦ ϕ−1 )
∂j (p)(f ) := (ϕ(p)).
∂xj
Then ∂j (p) ∈ Derp C ∞ (M ) and (∂1 (p), . . . , ∂m (p)) is a basis of Tp M ∼
= Derp C ∞ (M ).
Therefore, any vector field X : M → TM can be expressed as
X m
X(p) = Xi (p)∂i (p)
i=1
on U with well-defined functions Xi : U → R. In this context, the following lemma
is an important exercise.
Lemma 1.14. Retain the above notation. The vector field X is smooth if and only
if its local expression in every chart is given by smooth functions.
We now exhibit a hidden algebra structure on Γ(TM ) to which end we recall the
following definition: A derivation of C ∞ (M ) is a linear map δ : C ∞ (M ) → C ∞ (M )
such that δ(f g) = δ(f )g + f δ(g) for all f, g ∈ C ∞ (M ). Derivations can be defined
for any associative algebra over any field, they are purely algebraic objects.
Now, given a smooth vector field X ∈ Γ(TM ) on M and f ∈ C ∞ (M ) we define
LX (f )(p) = X(p)f . Then the map LX : C ∞ (M ) → C ∞ (M ) is a well-defined linear
map. Furthermore, we have the following.
Proposition 1.15. Retain the above notation. The map LX : C ∞ (M ) → C ∞ (M ) is
a derivation and the map Γ(TM ) → Der(C ∞ (M )), X 7→ LX is an isomorphism.
Proof. The fact that LX is a derivation follows from the fact that f 7→ (LX f )(p)
is a derivation at p ∈ M .
Injectivity of the map Γ(TM ) → Der(C ∞ (M )), X 7→ LX follows from the fact
that Tp M and Derp (C ∞ (M )) are isomorphic.
We now turn to surjectivity: Let δ ∈ Der(C ∞ (M )). Then for every p ∈ M ,
the map f 7→ δ(f )(p) is a derivation at p. Hence there is a well-defined vector
X(p) ∈ Tp M such that δ(f )(p) = X(p)(f ) for all f ∈ C ∞ (M ). Together, these
vectors form a vector field X on M . We need to verify that X is smooth. To this
end, we argue in local coordinates: Let (U, ϕ) be a chart of M and let X(p) =
P m
i=1 Xi (p) ∂i (p) (p ∈ U ) be the associated representation of X. Smoothness of X
amounts to smoothness of the coefficient functions Xi (i ∈ {1, . . . , m}). For every
f ∈P C ∞ (M ), the map p 7→ δ(f )(p) = X(p)f is smooth. In particular, the map
m
p 7→ i=1 Xi (p)∂i (p)(f ) is smooth on U . Now recall that
∂(f ◦ ϕ−1 )
∂i (p)(f ) = (ϕ(p)).
∂xi
Hence, if f ◦ ϕ−1 : ϕ(U ) → R was given by x 7→ xl we would obtain ∂i (p)(fl ) = δil
and conclude that Xl is smooth. However, there may not be such a function defined
on the whole of M . Nevertheless, we can remedy the argument by multiplying with
a smooth bump function. 
The interpretation of smooth vector fields on a manifold M as derivations of
C ∞ (M ) is important because it uncovers an important structure on Γ(TM ), namely
a Lie algebra structure.
Definition 1.16. Let W be a vector space and A, B ∈ End(W ). The bracket of A
and B is given by [A, B] := AB − BA ∈ End(W ).
RIEMANNIAN GEOMETRY 11

The relevance of this notion in our context is due to the following.


Lemma 1.17. Let M be a manifold and let δ, δ ′ ∈ Der(C ∞ (M )). Then [δ, δ ′ ] ∈
Der(C ∞ (M )).
Note that one would not expect a composition of derivations to be a derivation
in general as second-order derivatives might be introduced.

Proof. (Lemma 1.17). We compute


δδ ′ (f g) = δ(δ ′ (f )g + f δ ′ (g)) = δδ ′ (f )g + δ ′ (f )δ(g) + δ(f )δ ′ (g) + δδ ′ (g)f
as well as
δ ′ δ(f g) = δ ′ (δ(f )g + f δ(g)) = δ ′ δ(f )g + δ(f )δ ′ (g) + δ ′ (f )δ(g) + δ ′ δ(g)f
and therefore
[δ, δ ′ ](f g) = [δ, δ ′ ](f )g + [δ, δ ′ ](g)f.


Definition 1.18. Let M be a smooth manifold and let X, Y be smooth vector fields
on M . The bracket [X, Y ] of smooth vector fields X and Y is the smooth vector
field corresponding to [LX , LY ] ∈ Der(C ∞ (M )) via the isomorphism Γ(TM ) →
Der(C ∞ (M )).
Symbolically, we have L[X,Y ] = [LX , LY ]. We now compute [X, Y ] in local coor-
dinates. Let (U, ϕ) be a chart of M and write
m
X m
X
X(p) = Xi (p)∂i (p) and Y (q) = Yj (q)∂j (q)
i=1 j=1

for p, q ∈ U . Then for every f ∈ C ∞ (M ) we have


m
X
LX (LY (f ))(p) = Xi (p)∂i (p)(LY (f ))
i=1

where
m
X
∂i (p)(LY (f )) = ∂i (p)(Yj · ∂j f )
j=1
Xm m
X ∂ 2 (f ◦ ϕ−1 )
= ∂i (p)Yj · ∂j (p)(f ) + Yj (p) (ϕ(p))
j=1 j=1
∂xj ∂xi

We therefore have
X X ∂ 2 (f ◦ ϕ−1 )
LX (LY (f ))(p) = Xi (p)∂i (p)(Yj )∂j (p)(f ) + Xi (p)Yj (p) (ϕ(p))
i,j i,j
∂xj ∂xi

and
X X ∂ 2 (f ◦ ϕ−1 )
LY (LX (f ))(p) = Yi (p)∂i (p)(Xj )∂j (p)(f ) + Yi (p)Xj (p) (ϕ(p))
i,j i,j
∂xj ∂xi
| {z }
P 2 (f ◦ϕ−1 )
i,j Yj (p)Xi (p) ∂ ∂xi ∂xj (ϕ(p))

By Schwarz’s theorem, we thereby conclude


m m
!
X X
(L[X,Y ] f )(p) = Xi (p)∂i (p)(Yj ) − Yi (p)∂i (p)(Xj ) ∂j (p)(f )
j=1 i=1
12 MARC BURGER STEPHAN TORNIER

That is m
X
[X, Y ](p) = Zj (p)∂j (p)
j=1
where m
X
Zj (p) = (Xi (p)∂i (p)(Yj ) − Yi (p)∂i (p)(Xj ))
i=1
Note that the value of [X, Y ] at p ∈ M does not only depend on the values of X
and Y at p, which would make the bracket a linear algebra story, but on the values
of X and Y on a neighbourhood of p.
Also observe that the bracket [X, Y ] is bilinear, i.e. for all smooth vector fields
X1 , X2 , Y on M and scalars λ1 , λ2 ∈ R we have
[λ1 X1 + λ2 X2 , Y ] = λ1 [X1 , Y ] + λ2 [X2 , Y ]
and similarly for the Y -slot. It seems like this turns [−, −] into a “product”. However,
it is not associative. Instead it satisfies Jacobi’s identity which is the content of the
following Proposition.
Proposition 1.19. Let M be a smooth vector field and let X, Y and Z be smooth
vector fields on M . Then
[X, [Y, Z]] + [Y, [X, Z]] + [Z, [Y, X]] = 0
A way to remember this identity is to note that in the second and third term,
the entries X, Y and Z are cyclically permuted. It holds true more generally: Given
a vector space W and endomorphisms A, B, C ∈ End W it is an easy computation.
Another interpretation of the identity is the following: Given a smooth vector
field X on M , define
ad(X) : Γ(TM ) → Γ(TM ), Y 7→ [X, Y ].
Then ad(X) ∈ End(Γ(TM )) and the Jacobi identity amounts to ad(X) preserving
brackets: For all X1 , X2 ∈ Γ(TM ) we have
ad([X1 , X2 ]) = [ad X1 , ad X2 ].
There is also a geometric interpretation of the Jacobi identity in terms of flows of
the occurring vector fields which we exhibit later.
1.3.2. Vector Fields on Rn . Let Ω ⊆ Rn be an open set. A smooth vector field on
Ω is a smooth map X : Ω → Rn . Note however, that the mental picture of a vector
field in which the vectors are viewed as being attached to the point. We recall the
following existence and uniqueness theorem for integral curves of vector fields in
Rn or rather solutions of ordinary differential equations, see e.g. [Kön13, 4.2 II].
Theorem 1.20. Let Ω ⊆ Rn be open and let Y : Ω → Rn be a smooth vector field.
(i) For every x0 ∈ Ω there is an open interval Ix0 ⊆ R containing 0 ∈ R and
an open set Vx0 ⊆ Ω containing x0 ∈ Ω such that for every x ∈ Vx0 there
exists a smooth curve cx : Ix0 → Ω such that
(
cx (0) = x
.
Y (cx (t)) = c′x (t) ∀t ∈ Ix0
(ii) For every x ∈ Vx0 , any smooth curve γ : I → Ω satisfying
(
γx (0) = x
Y (γx (t)) = γx′ (t) ∀t ∈ Ix0
coincides with cx on some neighbourhood of 0 ∈ R.
(iii) The map Vx0 × Ix0 → Ω, (x, t) 7→ cx (t) is smooth.
RIEMANNIAN GEOMETRY 13

Recall that the key to the proof of Theorem 1.20 consists in transforming an
ordinary differential equation into an integral equation, identifying contractivity
and applying Banach’s fixed point theorem. An analogous statement holds in the
case of manifolds.
Corollary 1.21. Let M be a smooth manifold and let X : M → TM be a smooth
vector field.
(i) For every x0 ∈ M there is an open interval Ix0 ⊆ R containing 0 ∈ R and
an open set Vx0 ⊆ M containing x0 ∈ M such that for every x ∈ Vx0 there
exists a smooth curve cx : Ix0 → M such that
(
cx (0) = x
.
X(cx (t)) = c′x (t) ∀t ∈ Ix0

(ii) For every x ∈ Vx0 , any smooth curve γ : I → M satisfying


(
γx (0) = x
X(γx (t)) = γx′ (t) ∀t ∈ Ix0

coincides with cx on some neighbourhood of 0 ∈ R.


(iii) The map Vx0 × Ix0 → M , (x, t) 7→ cx (t) is smooth.
In Corollary 1.21, note that c′x (t) := Dt (cx )(1). Its proof consists of expressing
everything in local coordinates, applying Theorem 1.20 and transforming back.
Later on, we will globalize this local existence and uniqueness statement in the case
of compact manifolds.
Retain the notation of Corollary 1.21. For every t ∈ Ix0 we may define the map
θt : Vx0 → M , x 7→ cx (t). The following interesting local group property of the maps
θt is due to the uniqueness part of Corollary 1.21. Taking the necessary precautions,
its proof is mostly formal.
Corollary 1.22. Retain the above notation. Let t1 , t2 and t1 + t2 be in Ix0 . Further,
let W ⊆ Vx0 be such that θt1 (W ) ⊆ Vx0 . Then θt2 ◦ θt1 and θt1 +t2 are defined and
agree on W .
Proof. Without loss of generality, we assume t1 , t2 ≥ 0. Given x ∈ W we consider
the curve γ(s) := cx (t1 + s) for s ∈ [0, t2 ]. Setting T : R → R to denote the
translation s 7→ t1 + s we can rewrite γ(s) = cx ◦ T (s). We now compute
γ ′ (s) = (Ds γ)(1) = Ds (cx ◦ T )(1) = DT (s) cx ◦ Ds T (1)
= DT (s) cx (1) = X(cx (T (s))) = X(γ(s)).

Now observe that γ(0) = cx (t1 ). Hence, by uniqueness, γ(s) = ccx (t1 ) (s). In partic-
ular, for s = t2 we obtain for all x ∈ W :
θt1 +t2 (x) = cx (t1 + t2 ) = γ(t2 ) = ccx (t2 ) = θt2 (cx (t1 )) = θt2 (θt1 (x)).
Hence the assertion. 

Later on, we will revisit this result in the case of Lie groups where it takes a
more global form.
Pn
Example 1.23. Let M = Rn and consider the radial vector field X(x) := i=1 xi ∂i (x).
To determine its integral curves c, note that the condition X(c(t)) = c′ (t) translates
to c′i (t) = ci (t) where c(t) = (c1 (t), . . . , cn (t)). Consequently c(t) = ki et . Taking into
account an initial condition cx (0) = x we get cx (t) = et x and therefore θt (x) = et x.
14 MARC BURGER STEPHAN TORNIER

One checks that indeed θt1 +t2 = θt2 ◦ θt1 in this case.

Example 1.24. To illustrate that the precautions taken in Corollary 1.22 are indeed
necessary, consider the case where M = R and X(x) := x2 ∂x (x). The associated
initial value problem is c′x (t) = c2x (t) with cx (0) = x. Taking the physicist’s approach
to solving this, we obtain
 ′
−c′ 1 1
= = −1 ⇒ = −t + K
c2 c c
and hence c(t) = 1/(−t + K). Therefore, cx (t) = 1/(−t + 1/x). Assuming x > 0
the maximum interval of definition of cx is (−∞, 1/x). In particular, there is no
uniform interval (−ε, ε) of definition that works for all initial values.
A possible condition to ensure that the maps θt are defined on the whole manifold
for all times is compact support of the underlying vector field as in the following
proposition.
Proposition 1.25. Let M be a smooth manifold and let X ∈ Γ(TM ) be a smooth
vector field with compact support. Then the map, R → Diff(M ), t 7→ θt is a
homomorphism.
Homomorphisms as in Proposition 1.25 are termed one-parameter subgroups of
diffeomorphisms for obvious reasons.
Proof. Let K = supp(X) = {x ∈ M | X(x) 6= 0}. Using Corollary 1.21 we may
cover K with open sets V1 , . . . , Vl such that there are intervals I1 , . . . , Il to the extent
that for every x ∈ Vi there is a solution cx : Ii → M of the initial value problem
Tl
X(cx (t)) = c′x (t), cx (0) = x. Choose ε > 0 such that (−ε, ε) ⊆ i=1 Ii and set
Ω := V1 ∪· · ·∪Vl . Then for every t ∈ (−ε, ε) the map θt : Ω → M is defined. Outside
K, nothing happens: If x ∈ M \K then cx (t) = x is a solution to X(cx (t)) = c′x (t),
cx (0) = x for all t ∈ R. In this way, we obtain a map θt : M → M defined on the
whole of M for all t ∈ (−ε, ε). We now use Corollary 1.22 to extend θt to all times:
First of all, Id = θt−t = θ−t ◦ θt and hence θt ∈ Diff(M ). Secondly, given t ∈ R, let
t = k · (ε/2) + r where k ∈ Z and r ∈ (−ε, ε); then set θt := (θε/2 )k ◦ θr ∈ Diff(M ).
It remains to verify that θt1 +t2 = θ2 ◦ θ1 which can be done using additivity for
small values of t1 , t2 and t1 + t2 . 
The flows associated to vector fields are very interesting. Although they distort
the area started with they sometimes act volume-preservingly. In this case the fol-
lowing may hold: Given a subset A ⊆ M , consider the set {0 ≤ t ≤ T | θt (x) ∈ A} ⊆
R. Taking its volume, normalizing by T and letting T tend to infinity, Birkhoff’s
ergodic theorem states that
1
lim L({0 ≤ t ≤ T | θt (x) ∈ A}) = vol(A)
T →∞ T

Chaotic behaviour and return to the starting point is typical in particular in the
case of compact manifolds. For instance, given an irrationally oriented vector field
on the two-torus, the flow always returns to any set with positive measure.
RIEMANNIAN GEOMETRY 15

Corollary 1.26. Let M be a smooth compact manifold and let X ∈ Γ(TM ) be a


smooth vector field on M . Then there is a one-parameter group of diffeomorphisms
R → Diff(M ), t 7→ θt such that for every x ∈ M , the map R → M , t 7→ θt (x) is the
integral curve of X passing through x at time t = 0.
We end the discussion of vector fields by introducing the push-forward of a
vector field by a diffeomorphism, relating it the various other operations we have
introduced and giving a geometric interpretation of the Lie bracket.
Whereas differential forms can be pulled back their dual objects, vector fields,
are pushed forward. However, not every smooth map can be used to do this. For
instance, a smooth map might collapse a curve to a point, preventing a natural
choice for the push-forward vector at said point. We therefore consider the following
situation.
Definition 1.27. Let M and N be smooth manifolds and let X ∈ Γ(TM ) be a
smooth vector field. Further, let φ : M → N be a diffeomorphism. The push-forward
vector field Y on N is given by Y (φ(y)) := (Dx φ)(X(x)).
Lemma 1.28. Retain the notation of Definition 1.27. Then Y is a smooth vector
field. Furthermore, if LX and LY denote the derivatives associated to X and Y
respectively then LY f = LX (f ◦ φ) ◦ φ−1 for all f ∈ C ∞ (N ).
Given a smooth map φ : M → N of smooth manifolds, recall that the the
associated algebra homomorphism φ∗ : C ∞ (N ) → C ∞ (M ) is given by φ∗ (f ) = f ◦φ.
In terms of φ∗ , Lemma 1.28 states that the following diagram is commutative.
φ∗
C ∞ (N ) / C ∞ (M )

LY LX
 
C ∞ (N ) / C ∞ (M )
φ∗

With this in mind, the proof of Lemma 1.28 amounts to a computation.


Proof. (Lemma 1.28). Compute for all f ∈ C ∞ (N ):
(LY f )(y) = Y (y)f = (Dy f )(Y (y)) = (Dy f )(Dx φ(X(x))) = (Dφ(x) f )(Dx φ(X(x)))
= Dx (f ◦ φ)(X(x)) = LX (f ◦ φ)(x) = LX (f ◦ φ)(φ−1 (y))

The relation between push-forward vector fields and the bracket operation is the
following.
Lemma 1.29. Let M and N be smooth manifolds and let φ : M → N be a diffeo-
morphism. Further, let X and Z be smooth vector fields on M . Then
[φ∗ (X), φ∗ (Z)] = φ∗ ([X, Z]).
Proof. Here, it is beneficial to think in terms of the associated derivations: We have
Lφ∗ X = (φ∗ )−1 LX φ∗ and Lφ∗ (Z) = (φ∗ )−1 LZ φ∗ .
Computing brackets yields
[Lφ∗ X , Lφ∗ Z ] = (φ∗ )−1 LX LZ φ∗ − (φ∗ )−1 LZ LX φ∗ = (φ∗ )−1 [LX , LZ ]φ∗ .
Hence the assertion. 
Next up is the relation between push-forward vector fields and flows which relies
on a uniqueness rather than computation argument.
16 MARC BURGER STEPHAN TORNIER

Lemma 1.30. Let M and N be smooth manifolds and let φ : M → N be a diffeo-


morphism. Further, let X be a smooth vector field on M . Set Y := φ∗ X and let ψt
denote the flow of Y . Then ψt = φθt φ−1 for all t ∈ R where θt is the flow of X.
Proof. Set ψet (y) := φθt φ−1 (y). Then

d
ψet (y) = Dφ−1 (y) φ(X(φ−1 (y))) = Dx φ(X(x)) = Y (y)
dt t=0
which implies the assertion by uniqueness. 
We end this chapter with the following geometric interpretation of the bracket.
Proposition 1.31. Let M be a smooth manifold and let X and Y be smooth vector
fields on M . Assume that the local group θt of Y is defined. Then

d
((θt )∗ X) = [X, Y ].
dt t=0
Note that the expression ((θt )∗ X) (x) implicit in Proposition 1.31 is a tangent
vector at x ∈ M , namely the derivative of X with respect to Y at x ∈ M .
Proof. As an exercise, recall that if f : (−ε, ε) × M → R is smooth with f (0, p) = 0
for all p ∈ M then there is a smooth map g : (−ε, ε)×M → R with f (t, p) = tg(t, p).
Given f ∈ C ∞ (M ), we apply this fact to f (t, p) := f (θt (p)) − f (p) so that
g(0, p) = Y (p)(f ). We have
((θt )∗ X)(p)(f ) = LX (f ◦ θt ) ◦ θt−1 (p)
= X(θt−1 (p))(f ◦ θt )
= X(θt−1 (p))(f + tg(t, −)).
We therefore get
(θt )∗ X(p)(f ) − X(p)(f ) X(θt−1 (p))(f ) − X(p)(f )
= + X(θt−1 (p))(g(t, −))
t t
X(θ−t (p))(f ) − X(p)(f )
= + X(θ−t (p))(g(t, −))
t
If t tends to zero, we obtain
X(θ−t (p))(f ) − X(p)(f )
→ −LY LX (f )(p)
t
and
X(θ−t (p))(g(t, −)) = LX (g(t, −))(θ−t (p)) → LX LY (f )(p)
which implies the assertion. 
1.4. Lie Groups: A very short introduction. Lie groups are particularly in-
teresting manifolds and are central to both building large classes of examples of
manifolds and actions on manifolds. They are also crucial to seemingly unrelated
mathematics such as Fermat’s last theorem.
Definition 1.32. A Lie group is a smooth manifold G endowed with a group struc-
ture such that the product map G × G → G, (x, y) 7→ xy and the inverse map
G → G, x 7→ x−1 are smooth.
We have already seen many example of Lie groups.
Example 1.33.
(i) The manifold Rn is a Lie group with respect to addition.
(ii) The manifold GL(n, R) is a Lie group with respect to matrix multiplication.
RIEMANNIAN GEOMETRY 17

(iii) Given a quadratic form q on Rn , the manifold O(q) is a Lie group with
respect to matrix multiplication.
(iv) The symplectic groups of example 1.3 are Lie groups as well.
(v) It is a highly non-trivial result of E. Cartan that every closed subgroup
G ≤ GL(n, R) is a regular submanifold and hence a Lie group. Note the
sharp contrast to the fact that both closed subsets and non-closed subgroups
of GL(n, R) can behave very badly: Consider for instance GL(n, Q) which
can only be turned into a Lie group by making the topology discrete.

We now illustrate that all the notions we have developed work together nicely
in the case of Lie groups and thereby underline the importance of the latter. First
of all, note that if G is a Lie group and g ∈ G, the left multiplication Lg : G → G,
x 7→ gx is a preferred diffeomorphism of G sending the identity element e ∈ G to
g ∈ G. Indeed, it is smooth by the axioms of a Lie group and admits the smooth
inverse Lg−1 . Furthermore, we have the following.

Proposition 1.34. Let G be a Lie group. Then


(i) G is orientable, and
(ii) G is parallelizable.

Proof. To show that G is orientable, we construct a nowhere vanishing top form


on G: Let ω0 ∈ Λn ((Te G)∗ ) be a non-zero, alternating n-form where n = dim G
- recall that if V is an n-dimensional vector space then dim Λk (V ∗ ) = nk , in par-
ticular dim Λn (V ∗ ) = 1. We now propagate this form to the whole of G using left
multiplication: For g ∈ G, set

ωg (v1 , . . . , vn ) = ωe (Dg Lg−1 v1 , . . . , Dg Lg−1 vn ).

One checks that ω ∈ Ωn (G) so defined is smooth. By construction it is nowhere van-


ishing, i.e. volume form. Hence G is orientable. Let us endow G with the orientation
for which a basis e1 , . . . , en is positively oriented if and only if ωg (e1 , . . . , en ) > 0.
Parallelizabiliy of G is proven as in the case of the torus, see 1.11. 

Corollary 1.35. Let G be a Lie group and let ω denote the volume form on G
constructed in the proof of Proposition 1.34. Then the linear map I : C00 (G) → R,
defined Ron the space of continuous, compactly supported functions on G is by
I(f ) := G f ω yields a left-invariant positive Radon measure on G via Riesz rep-
resentation.

Let µ denote the measure obtained on a Lie group via Corollary 1.35. It is called
Haar measure and satisfies µ(gE) = µ(E) for every Borel set E ⊆ G and every
element g ∈ G. Haar measures exist more generally for locally compact Hausdorff
groups. However, in the general case there is a no smooth structure to work with
and hence the proof is based on different ideas.

Proof. (Corollary 1.35). We prove invariance of the functional I from which invari-
ance of the associated measure follows. Let f ∈ C00 (G) and h ∈ G. Observe that
(f ◦ Lh )ω = (L∗h )(f · (L∗h−1 ω) and hence
Z Z Z
I(f ◦ Lh ) = (f ◦ Lh )ω = L∗h η = η
G G G
18 MARC BURGER STEPHAN TORNIER

where η := f · (L∗h−1 ω. Now compute


((Lh−1 )∗ ω)g (v1 , . . . , vn ) = ωLh−1 (g) (Dg Lh−1 (v1 ), . . . , Dg Lh−1 (vn ))
= ωh−1 g (Dg Lh−1 (v1 ), . . . , Dg Lh−1 (vn ))
= ω0 (Dh−1 g Lg−1 h Dg Lh−1 (v1 ), . . . , Dh−1 g Lg−1 h Dg Lh−1 (vn ))
= ω0 (Dg Lg−1 v1 , . . . , Dg Lg−1 vn )
= ωg (v1 , . . . , vn ).
whence η = f ω and therefore I(f ◦ Lh ) = I(f ). 
Observe that if G is a compact Lie group then the functional I of Corollary 1.35
is defined for all f ∈ C(G).
Corollary 1.36. Every compact Lie subgroup of GL(n, R) is conjugate into O(n).
Proof. Let K ≤ GL(n, R) be a compact Lie subgroup and let µ denoteR its Haar mea-
sure. Choose any scalar product h−, −i on Rn and define B(v, w) := G hgv, gwi dµ(g)
for all v, w ∈ Rn . Then it is an easy verification that B is a K-invariant scalar
product. Now remember from linear algebra that if (−, −) denotes the standard
scalar product then there exists A ∈ GL(n, R) such that B(u, v) = (Au, Av) for
all u, v ∈ Rn . Since B is K-invariant we obtain (Aku, Akv) = (Au, Av). Setting
u′ := Au and v ′ := Av we conclude (AkA−1 u′ , AkA−1 v ′ ) = (u′ , v ′ ) and hence
AKA−1 ⊆ O(n). 
We now apply the theory that we have developed for manifolds to the special
case of Lie groups in which things behave particularly nice.
Definition 1.37. A Lie algebra is a vector space g endowed with a product g × g → g,
denoted (A, B) 7→ [A, B] which is
(i) bilinear,
(ii) satisfies [A, B] = −[B, A] for all A, B ∈ g, and
(iii) satisfies [A, [B, C]] + [B, [C, A]] + [C, [A, B]] = 0 for all A, B, C ∈ g.
Example 1.38. Let M be a smooth manifold. Then g = Γ(TM ), endowed with the
bracket of vector fields, is a Lie algebra. However, it is infinite-dimensional unless
M is a point.
Definition 1.39. Let G be a Lie group. A smooth vector field X ∈ Γ(TG) is left-
invariant if (Lg )∗ X = X for all g ∈ G.
Let Γinv (TG) be the subspace of Γ(TG) consisting of left-invariant vector fields.
Proposition 1.40. Let G be a Lie group. Then Γinv (TG) is a Lie algebra and the
map Γinv (TG) → Te G given by X 7→ X(e) is a vector space isomorphism.
In particular, the Lie algebra Γinv (TG) is finite-dimensional. In some way, it
encodes the group structure of G. However, it is unclear yet in what way it does so.
Proof. In Lemma 1.28 we have shown that (Lg )∗ [X, Y ] = [(Lg )∗ X, (Lg )∗ Y ] for all
X, Y ∈ Γ(TG) and g ∈ G. Hence, if X and Y are invariant smooth vector fields on
G then so is [X, Y ].
Injectivity of the evaluation map is due to the fact that left-invariance of X
yields X(g) = (De Lg )(X(e)). Surjectivity follows from defining Y ∈ Γinv (TG) by
setting Y (g) := (De Lg )(v) for a given v ∈ Te G. 
The isomorphism of Proposition 1.40 can be used to turn Te G into a Lie algebra
as follows: Given u, v ∈ Te G, define [u, v] ∈ Te G to be the evaluation of left-
invariant vector field [Xu , Xv ] at e where Xu and Xv are the left-invariant vector
RIEMANNIAN GEOMETRY 19

fields corresponding to u and v respectively. Note that in order to know [u, v] it


does not suffice to know u and v, it requires knowledge of the vector fields Xu and
Xv on a small neighbourhood of the identity in G.
One checks that the bracket on the Lie algebra of Rn is trivial. Generally, it
measures the degree of non-commutativity and is best used for highly non-abelian
groups, such as semisimple ones.
Definition 1.41. Let G be a Lie group. The Lie algebra of G is the vector spcae
g := Te G endowed with the product [x, y] := [X, Y ](e) where X and Y are the
left-invariant vector fields associated to x and y in Te G.
In the context of general manifolds we have studied flows of vector fields. There
is a lot to say about these in the context of Lie groups.
Proposition 1.42. Let G be a Lie group and let X ∈ Γinv (TG). Then the local group
θt is defined for all t ∈ R. In addition,
(i) θt (g) = gθt (e) for all t ∈ R and g ∈ G, and
(ii) the map R → G, t 7→ θt (e) is a smooth homomorphism whose tangent
vector at e ∈ G is X(e).
Proof. The main part of Proposition 1.42 is to show the existence of integral curves
for all times. The remaining statements are then easy consequences. For every g ∈ G,
let Ig ⊆ R be the largest open interval of definition of cg which contains 0 ∈ R.
We are going to show that Ig is independent of g which suffices as we shall see. For
once, we claim that cg (t) = Lg (ce (t)) which can be verified using uniqueness: Let
γ(t) := Lg (ce (t)) for t ∈ Ie . Then
γ ′ (t) = (Dce (t) Lg )(c′e (t)) = (Dce (t) Lg )(X(ce (t))) = X(gce (t)) = X(γ(t)).
Furthermore, γ(0) = gce (0) = ge = g. hence γ(t) = cg (t) by uniqueness. This
proves the claim. In particular, we conclude that Ig = Ie =: I for all g ∈ G. Let
ε > 0 be such that (−ε, ε) ⊆ I and choose t ∈ I. For small enough s we have
cg (t + s) = ccg (t) (s) by uniqueness. Now, the right hand side is defined at least for
all s ∈ (−ε, ε) which implies t + (−ε, ε) ⊆ Ig = Ie = I. Hence I = R.
The claim can be rewritten as θt (g) = gθt (e). Setting g = θs (e) we obtain
θt ◦ θs (e) = θs (e)θt (e).
On the other hand, we know that θt ◦ θs = θt+s and hence θt+s (e) = θs (e) · θt (e).
In other words, the map R → G, t 7→ θt (e) is a homomorphism. 

The proof of Proposition 1.42 leads us to the following, fundamental object.


Definition 1.43. Let G be a Lie group and let g be its Lie algebra. The exponential
map expG : g → G is given by x 7→ θ1 (e) where θt denotes the one-parameter group
of diffeomorphism associated to the left-invariant vector field X with X(e) = x.
We observe that as a consequence of Definition 1.43, we have θt (e) = exp(tx)
which amounts to saying the vector fields can be scaled: Indeed, the map R →
Diff(G), s 7→ θst is the one-parameter group associated to tX; its value at s = 1
being exp(tx) coincides with θt (e).
We shall see shortly that the exponential map of a Lie group G with Lie algebra
g is a local diffeomorphism at 0 ∈ g. That is, Lie groups come equipped with
canonical charts that are very well adapted to the group structure. For (Rn , +)
the exponential map merely associates the position vector to a tangent vector. Its
name, however, stems from the fact that in the case of GL(n, R), the exponential
map is the usual matrix exponential.
20 MARC BURGER STEPHAN TORNIER

Example 1.44. As an example, we determine the exponential map in the case of


GL(n, R). The manifolds structure of GL(n, R) stems from the fact that it is an
open subset of Mn,n (R). We may hence identify TId GL(n, R) with Mn,n (R). Let
x ∈ Mn,n (R). For appropriate ε > 0, the curve t 7→ Id +tx through Id ∈ Mn,n (R)
is contained in GL(n, R) for all t ∈ (−ε, ε). Let X be the left-invariant vector field
on GL(n, R) associated to x ∈ Mn,n (R). Then

d
X(g) = (De Lg )(x) = g(Id +tx) = gx.
dt t=0

From this we deduce that (LX f )(g) = X(g)(f ) = (Dg f )(gx) for all f ∈ C ∞ (G)
and g ∈ GL(n, R). A computation now implies that for all x, y ∈ Mn,n (R) =
TId GL(n, R) we have [x, y] = xy −yx. That is, in this case, the Lie bracket coincides
with the usual commutator bracket operation.
Now, the differential equation given by an invariant vector field X is X(c(t)) =
c′ (t), that is c(t)xP
= c′ (t). Its solution with initial condition c(0) = Id is given by

c(t) = Exp(tx) = k=0 (tk xk )/k!. Thus expGL(n,R) (x) = Exp(x) is the usual matrix
exponential, hence the name of the general Lie group exponential.

Proposition 1.45. Let G be a Lie group with Lie algebra g. The exponential map
expG : g → G is smooth and D0 expG : g → g is the identity map.
d

Proof. Identifying T0 g with g we have D0 expG (x) = dt t=0
expG (tx) = x. 

The importance of Proposition 1.45 is due to the fact that it implies that the
exponential map is a diffeomorphism on a neighbourhood of 0 ∈ g. Hence its inverse
can be employed as a chart.

Lie subgroups. It is also essential that the exponential map of a Lie group G de-
termines the exponential maps of all its Lie subgroups H. The difficulty of this
statement lies in the definition of Lie subgroup. Whereas this chapter does not aim
at being a comprehensive treatment of the foundations of Lie theory, we elaborate a
bit on this important point: Let G be a Lie group and let H ≤ G be a subgroup. If
H is also a regular submanifold and the map H × H → H, (x, y) 7→ xy −1 is smooth
then H is a Lie group. In this case, one can verify that h := Te H ≤ Te G = g is a
subalgebra. In general, however, we do not get every subalgebra h of g in this fash-
ion which is unacceptable from a categorical point of view. For instance, consider
the torus T2 = Z2 \ R2 whose Lie algebra is Lie(T2 ) = R2 has trivial bracket. Con-
sequently,
√ Tall subspaces of Lie(T2 ) are subalgebras. But the subalgebra generated
by (1, 2) ∈ R2 is not the Lie algebra of a regularly embedded subgroup of T2 .
In fact, one has to relax the requirement of the subgroup to be regularly embed-
ded: In general, a Lie subgroup of a Lie group G is a pair (H, i) consisting of a Lie
group H and an injective immersion i : H → G, allowing for the topologies on H
and i(H) ⊆ G to differ as in the above case.

Proposition 1.46. Let G be a Lie group with Lie algebra g and let H ≤ G be a
regular submanifold. Then the Lie algebra of H is given by

h = Te H = {v ∈ Te G = g | expG (tv) ∈ H ∀t ∈ R}.

Example 1.47. Proposition 1.46 allows us to compute the Lie algebras of most our
examples, which are subgroups of GL(n, R). In this case, the general recipe is to
write down the defining equation of the group, substitute exp(tx) for the variable,
and take the derivative at t = 0 to obtain the defining equation of the Lie group.
RIEMANNIAN GEOMETRY 21

(i) In the case of O(n) we compute


Lie(O(n)) = {x ∈ Mn,n (R) | Exp(tx) ∈ O(n) ∀t ∈ R}
= {x ∈ Mn,n (R) | Exp(tx)T Exp(tx) = Id ∀t ∈ R}
= {x ∈ Mn,n (R) | xT + x = 0}

as (d/dt)|t=0 (Exp(tx)T Exp(tx)) = xT + x.


(ii) Similar to the above, we determine in the case of Sp(2n, R):
Lie(Sp(2n, R)) = {x ∈ M2n,2n (R) | xT J + Jx = 0}.
It is an exercise to think about what happens in cases such as U(n) and SU(n) as
subgroups of GL(n, C).
In a sense linear groups such as the examples above already cover almost all Lie
groups. This is made precise using representations of Lie groups: Every Lie group G
comes equipped with a natural action on a finite-dimensional vector space: Recall
that for h ∈ G, we have the right-multiplication diffeomorphism Rh : G → G given
by g 7→ gh. Any Rh commutes with any Lg (g, h ∈ G) which is due to the easily
overlooked associativity of the group multiplication. Therefore, given X ∈ Γinv (TG)
and h ∈ G, we have (Rh )∗ X ∈ Γinv (TG). Accordingly, we may define for h ∈ G
and x ∈ g = Te G: Ad(h)x := ((Rh−1 )∗ X)(e). In this way, we get a homomorphism
Ad : G → GL(g), called the adjoint representation. The kernel of the adjoint
representation is related to the center of G and is abelian, which is not considered
a drama because abelian Lie groups are well-understood.
We end this chapter on Lie groups with the two following results on the adjoint
representation of a Lie group.
Proposition 1.48. Let G be a Lie group with Lie algebra g. Then for all h ∈ G and
x ∈ g we have

d
Ad(h)(x) = h exp(tx)h−1 .
dt t=0
The proof of Proposition 1.48 is left as an exercise. Now, the adjoint represen-
tation of a Lie group G is a map from G to GL(g). As such it can be differentiated
at t = 0 to produce a homomorphism of the associated Lie algebras g and End(g).
This produces the following interpretation of the Lie bracket.
Proposition 1.49. Let G be a Lie group with Lie algebra g. For all x, y ∈ g we have

d
Ad(exp ty)(x) = [y, x].
dt t=0
The proof of Proposition 1.49 is a consequence of Proposition 1.31.

1.5. Coverings and Fibrations. In this section we discuss the important covering
and fibration mechanisms. In the following definition, we adopt the topological
notion of covering maps to our smooth setting.
Definition 1.50. Let M and M be smooth manifolds. A map p : M → M is a
covering map if
(i) p is smooth and surjective, and
(ii) for every mS∈ M there is an open neighbourhood U of m ∈ M such that
p−1 (U ) = i∈I Ui is a disjoint union of open subsets Ui ⊆ M such that
p|Ui : Ui → U is a diffeomorphism.
22 MARC BURGER STEPHAN TORNIER

b b b b

M M

Given a covering map p : M → M the map M → N0 ∪{∞} which to m ∈ M


associates the cardinality of its preimage under p is locally constant.
Example 1.51. Here are three of the most important examples and non-examples
of covering spaces.
(i) Let M := M := S 1 and consider p : M → M , z 7→ z d for some d ∈ Z where
S 1 is considered as {z ∈ C | |z| = 1}. Then p is a d-sheeted covering map
and given z ∈ M we have p−1 (z) = {exp(2πik/d) | 0 ≤ k < d}.
(ii) The map f : C → C, z 7→ z 2 is not a covering: There is no neighbourhood
of 0 ∈ C on which card(p−1 (z)) is constant. However, one could remove 0
from both M and M to turn f into a covering.
(iii) Set M := R and M = S 1 . Then the map t 7→ exp(2πit) is a covering.
One way in which covering maps arise is the following: Let M be a locally
compact Hausdorff space on which a group Γ acts by homeomorphisms. That is,
for each γ ∈ Γ the map M → M , m 7→ γm is a homeomorphism.
Definition 1.52. Let M be a locally compact Hausdorff space on which a group Γ
acts by homeomorphisms. That is, for each γ ∈ Γ the map M → M , m 7→ γm is a
homeomorphism.
(i) The action is properly discontinuous if for every compact K ⊆ M the set
{γ ∈ Γ | γ(K) ∩ K 6= ∅} is finite.
(ii) The action is free if for every x ∈ M and γ ∈ G\{e} we have γ(x) 6= x.
Proposition 1.53. Let M be a locally compact Hausdorff space on which a group
Γ acts by homeomorphisms. If the action is properly discontinuous then Γ\M is
Hausdorff.
In the case of a group Γ acting on a set Γ\M , we denote by Γ\M the set of
equivalence classes for the relation x ∼ y ⇔ Γx = Γy. We remark that in this
situation the Smap p : M → Γ\M is open: Indeed, if V ⊆ M is any set, then
p−1 (p(V )) = γ∈Γ γ(V ). Hence, if V is open, then p−1 (p(V )) is open as a union of
open sets.
Proof. (Proposition 1.53). Since p is open it suffices to show that R = {(x, y) ∈
M × M | x ∼ y} = {(x, γx) | x ∈ M, γ ∈ Γ} is a closed subset of M × M . To this
end, let (x0 , y0 ) ∈ R. For every compact neighbourhood of (x0 , y0 ) ∈ M × M of the
form V × W , define F (V, W ) := {γ ∈ Γ | γV ∩ W 6= ∅}. Observe that F (V, W ) is
always finite: Indeed,
F (V, W ) ⊆ {γ ∈ Γ | γ(V ∪ W ) ∩ (V ∪ W ) 6= ∅}
and the latter set is finite by the definition of proper discontinuity. Now, fix a
particular neighbourhood V0 × W0 and consider
F := {F (V, W ) | V × W ∋ (x0 , y0 ) compact neighbourhood and V × W ⊆ V0 × W0 }
Now make the following observations:
RIEMANNIAN GEOMETRY 23

(i) F (V, W ) ⊆ F (V0 , W0 ) for all F (V, W ) ∈ F .


(ii) F (V, W ) 6= ∅ for all F (V, W ) ∈ F : Indeed, since (x0 , y0 ) ∈ R we have
(V × W ) ∩ R 6= ∅ and a point (x, γx) ∈ (V × W ) gives rise to γ ∈ F (V, W ).
(iii) Given F (Vi , Wi ) ∈ F (i ∈ {1, . . . , n}), we have
n n n
!
\ \ \
F (Vi , Wi ) ⊇ F Vi , Wi 6= ∅.
i=1 i=1 i=1
T
As a consequence, we conclude F (V,W )∈F F (V, W ) 6= ∅. An element γ ∈ Γ that is
contained in the latter intersection has the property that for any compact neigh-
bourhoods V of x0 ∈ M and W of y0 ∈ M we have γV ∩ W 6= ∅. Since M is
Hausdorff this implies γ(x0 ) = y0 . 

Corollary 1.54. Let Γ be a group acting freely and properly discontinuously on a


manifold M by diffeomorphisms. Then there is a unique smooth manifold structure
on Γ\M such that p : M → Γ\M is a smooth covering.
Proof. (Sketch). Proper discontinuity and freeness of the action imply that for all
x ∈ M there is an open neighbourhood Vx of x such that for all γ ∈ Γ\{e} we
have γVx ∩ Vx = ∅. This is based on the following argument: Given γ ∈ Γ\{e},
pick disjoint neighbourhoods V of x and W of γx with V ∩ W = ∅ using that M is
Hausdorff. Then V ∩γ −1 W is a neighbourhood of x and (V ∩γ −1 W )∩(γV ∩W ) = ∅.
Adjust this argument to taking into account the finiteness of our situation.
For a neighbourhood Vx of x as above, we observe that p|Vx : Vx → p(Vx ) is
continuous, open, surjective and injective, hence a homeomorphism.
Finally, choose an atlas A on M consisting of charts (Vx , ϕx ) as above. A smooth
atlas on Γ\M is then given by
A′ = {(pVx , φx ) | (Vx , φx ) ∈ A, φx = ϕx ◦ (p|Vx )−1 }.
The smoothness of the transition maps comes from the assumption that Γ acts on
M by diffeomorphisms. 

Example 1.55. We collect several examples of this efficient construction.


(i) Let Γ = {± Id} and M = S n ⊆ Rn+1 . Then Γ\S n is diffeomorphic to real
projective n-space Pn (R).
(ii) Consider M := H + := {z ∈ C | Im(z) > 0}. One verifies that the map
  
a b az + b
SL(2, R) × M → M, , z 7→
c d cz + d
defines an action of SL(2, R) on H + . The reader is encouraged to show
that Γ := SL(2, Z), consisting of all the matrices in SL(2, R) with integer
entries, acts properly discontinuously on M via the above action (in contrast
to having dense orbits on the real line).
(iii) Let G be a Lie group and let Γ ≤ G be a discrete subgroup. Then the left
action Γ×G → G, (γ, g) 7→ γg of Γ on G is free and properly discontinuous.
Hence Γ\G admits a smooth manifold structure for which p : G → Γ\G is
a covering.
For the upcoming discussion of fibrations, we record the following: Given compact
subsets K1 and K2 of a Lie group G, the set K2 K1−1 = {k2 k1−1 | k2 ∈ K2 , k1 ∈ K1 }
is compact: Indeed, it is the image of the compact set K1 × K2 ⊆ G × G under the
smooth map G × G → G given by (x, y) 7→ yx−1 .
Fibrations are a common generalization of the concepts of vector bundles and
coverings.
24 MARC BURGER STEPHAN TORNIER

Definition 1.56. Let E, B and F be smooth manifolds and let π : E → B be a


smooth map. The triple (π, E, B) is a fiber bundle with base B, total space E and
fiber F if
(i) the map π : E → B is surjective, and
(ii) there is an open cover {Ui | i ∈ I} of B and diffeomorphisms
hi : π −1 (Ui ) → Ui × F with hi (π −1 (x)) = {x} × F.

Example 1.57. b

(i) Coverings yield fiber bundles with discrete fiber.


(ii) Vector bundles are fibrations with fiber a vector space.
(iii) Consider the following interpretation of S 3 ⊆ R4 : F E
E = {(z1 , z2 ) ∈ C2 | z12 + z22 = 1}
The group S 1 = {̺ ∈ C | |̺| = 1} acts on E via b

̺(z1 , z2 ) = (̺z1 , ̺z2 ). This action is free and the quo-


tient identifies with P1 (C). The latter is incidentally π
diffeomorphic to S 2 . This is a fibration of S 3 over S 2
with fiber S 1 , the so called Hopf fibration. b
B

Fibrations are particularly fruitful in the setting of Lie groups acting on man-
ifolds: An action of a Lie group G on a manifold M is smooth if the action map
G × M → M is smooth. In particular, for every g ∈ G the map M → M , m 7→ gm
is a diffeomorphism: It is smooth as a restriction of the smooth action map and has
an inverse arising in the same fashion. The action of G on M is proper if for every
compact set K ⊆ M the set {g ∈ G | gK ∩ K 6= ∅} has compact closure in G.
Theorem 1.58. Let G be a Lie group, M a manifold and G × M → M a smooth
action of G on M . If the action is free and proper then there is a unique smooth
manifold structure on G\M such that M → G\M is a smooth fibration.
A good reference for this Theorem is [vdB06], in which Theorem is deduced from
the following, more general equivalence relation version which mostly depends on
the inverse function theorem.
Theorem 1.59. Let M be a smooth manifold and let R ⊆ M × M be an equivalence
relation on M . If R is a closed submanifold of M × M and pr1 : R → M is a
submersion then the quotient R\M has a unique structure of smooth manifold
such that π : M → R\M is a submersion.
In particular, we can apply Theorem 1.58 in the case where M = G is a Lie
group and H ≤ G is a closed subgroup of G acting on G on the right: G × H → G,
(g, h) 7→ gh.
Lemma 1.60. Let G be a Lie group and let H ≤ G be closed. Then the right action
of H on G is proper and free.
Proof. Freeness is immediate. As to properness, let K ⊆ G be compact. Then
{h ∈ H | Kh ∩ K 6= ∅} = H ∩ K −1 K.
Since H is closed and K −1 K is compact, so is the above intersection. 

As a corollary, we obtain the following powerful mechanism of producing mani-


folds and actions on them.
RIEMANNIAN GEOMETRY 25

Corollary 1.61. Let G be a Lie group and let H be a closed subgroup of G. Then
there is a unique smooth manifold structure on G/H such that the quotient map
π : G → G/H is a smooth fibration with base G/H and fiber H. With this manifold
structure, the action G × G/H → G/H is smooth.
Corollary 1.61 produces one of the most enormous classes of manifolds with many
interesting subclasses depending on properties of G and H.
Example 1.62. To illustrate the usefulness of Corollary 1.61 we consider the follow-
ing: Equip Rn with the usual scalar product h−, −i. For 1 ≤ p ≤ n, let
Sp,n := {(x1 , . . . , xp ) ∈ (Rn )p | hxi , xj i = δij ∀i, j ∈ {1, . . . , p}}.
Observe that O(n) acts transitively on Sp,n as it acts transitively on orthonormal
bases. Let (e1 , . . . , en ) be the standard orthonormal basis of Rn and fix the basepoint
p := (e1 , . . . , ep ) ∈ Sp,n . The stabilizer of p in O(n) is given by
  
Idp 0
Hp = B ∈ O(n − p)
0 B
which is a closed, regularly embedded submanifold of O(n). Hence the bijection
O(n)/Hp ∼
= O(n)/O(n − p)
can be used to equip Sp,n with a smooth manifold structure. The manifolds Sp,n are
called Stiefel manifolds and play a fundamental role when it comes to characteristic
classes and vector bundles. A similar reasoning as above, lets us treat Grassmannian
manifolds as quotients of Lie groups and hence equip them with a manifold structure
in a very convenient way.
Combining results from above we record the following.
Corollary 1.63. Let M be a manifold and let b ∈ M . Furthermore, let G be a Lie
group acting smootly and transitively on M . Let H denote the stabilizer in G of
b ∈ M . Then the bijection G/H → M , gH 7→ gb is a diffeomorphism.
Corollary 1.63 is pleasant compatibility result and in particular states that one
cannot find any exotic smooth structures on a manifold by exhibiting it as a homo-
geneous space of a Lie group as above.
Example 1.64. We now have a look at further examples of this mechanism.
(i) The Lie group GL(n, R) acts smoothly and transitively on Rn \{0}. The
stabilizer of e1 = (1, 0, . . . , 0)T ∈ Rn in GL(n, R) is given by
  
1 v n−1
P0 = v ∈ R , A ∈ GL(n − 1, R)
0 A
which can be identified as a semi-direct product. Anyway, we conclude that
GL(n, R)/P0 is diffeomorphic to Rn \{0}.
(ii) The Lie group GL(n, R) also acts on the set Pn (R) of all lines passing
through 0 ∈ Rn+1 . Recall that Pn (R) can also be defined as the quotient
of Rn+1 \{0} by the multiplication action of R∗ , and the quotient of S n by
the antipodal map. Whereas the action of GL(n, R) is not visisble in the
second definition, the first one clearly shows its transitivity. The stabilizer
of [e1 ] ∈ Pn (R) is given by
  
a v n
P := a ∈ R \{0}, v ∈ R , B ∈ GL(n, R) .
0 A
As a consequence, GL(n + 1, R)/P is diffeomorphic to Pn (R).
26 MARC BURGER STEPHAN TORNIER

In the constext of (ii), the following exercises are worthwhile: For g ∈ GL(n, R),
study the qualitative behaviour of the action of {g n | n ∈ Z} on Pn (R). Given
that linear algebra classifies elements of GL(n, R) up to conjugacy, three particular
examples to look at are
     
λ1 cos t sin t 1 x
(λ1 6= λ2 ), and .
λ2 − sin t cos t 1
What do the behaviours say towards a classification of homeomorphisms of S 1 up
to conjugacy? Finally, one can consider the same question for the case of GL(3, R)
where one discovers new phenomena, semi-hyperbolicity.
Also, does there exist an SL(2, R)-invariant probability measure on P1 (R)? The
answer to this question is “no”. However, is this a general phenomenon, in the sense
that no compact homogeneous space SL(2, R)/H of SL(2, R) supports an invariant
probability measure? Again, the answer is “no”.
(iii) Returning to the examples, consider the following construction which is of
fundamental importance for the remainder of the course. Let q : Rn+1 → R
be the quadratic form given by q(x) = x21 + · · · + x2n − x2n+1 . This quadratic
form has three kinds of level sets:
(a) For a < 0, the equation x21 + · · · + x2n = x2n+1 + a has no solutions
√ √ √
with − −a < xn+1 < −a. For |xn+1 | > a, there are (rotationally
invariant) solution. In fact, we obtain a two-sheeted hyperboloid.
(b) The level set q −1 (0) is a cone with singular point 0 ∈ Rn+1 .
(c) For a > 0, the level set q −1 (a) is a one-sheeted hyperboloid.

Recall that the symmetry group of q is


O(n, 1) = {g ∈ GL(n + 1, R) | q(gx) = q(x) ∀x ∈ Rn+1 }.
A general theorem of Witt states that O(n, 1) acts transitively on all three
kinds of level sets, except for the obvious exclusion of 0 ∈ Rn+1 in the case
a = 0. We shall now restrict our attention to the upper sheet of q −1 (−1)
and denote it by Hn := {x ∈ q −1 (−1) | xn+1 > 0}. To this end we introduce
SO0 (n, 1) = {g ∈ O(n, 1) | g(Hn ) = Hn and det g = 1}.
RIEMANNIAN GEOMETRY 27

The homomorphism O(n, 1) → Sym(π0 (q −1 (−1))) × {−1, 1}, given the per-
mutation an element of O(n, 1) induces on the two sheets of q −1 (−1) and its
determinant, is surjective and continuous for the discrete topology on the
right hand side. Therefore, SO0 (n, 1) is an open subgroup of index four in
O(n, 1). We now show that SO0 (n, 1) acts transitively on Hn . The stabilizer
in SO0 (n, 1) of en+1 is given by
  
A 0
K := A ∈ SO(n) .
0 1
It acts as SO(n) in planes parallel to the plane given by xn+1 = 0: Given
 
A 0
k= ∈K
0 1
and x = (x1 , . . . , xn+1 )T ∈ Hn we have
  
x1
  .. 
A  . 
(x) =  .
 xn 
xn+1
In particular, there is A ∈ SO(n) such that
 
  0
0  ..

   ..   .
..  .   
   
A x1 . xn =
0  and hence k(x) = q 0 
q   
2  x2n+1 − 1
xn+1 − 1
1
where k ∈ SO0 (n, 1) is defined by A. In order to translate vertically, we
introduce
   
 Idn−1  0 

A := at :=  cosh(t) sinh(t)  t ∈ R
 0 
sinh(t) cosh(t)
which is a one-parameter subgroup of SO0 (n, 1). Indeed, it preserves the up-
per sheet of q −1 (1), all its elements have determinant one and one computes
q(at x) = q(x) thanks to the identity cosh2 (t) − sinh2 (t) ≡ 1. Now,
 
0
 .. 
 . 
 
at (en+1 ) =  0 
 
 sinh t 
cosh t
which combined with the K-action shows transitivity of SO0 (n, 1) on Hn .
Corollary 1.65. The map SO0 (n, 1)/K ∼
= K, gK 7→ g(en+1 ) is a diffeomorphism
and SO0 (n, 1) is connected.
Proof. The connectedness of SO0 (n, 1) is due to the exercise that a total space of
a fiber bundle with connected base and connected fiber is itself connected. 

Let us now compute the Ptangent space of Hn at x ∈ Hn . To this end, let b :


n+1 n+1 n
R ×R → R, (x, y) 7→ i=1 xi yi − xn+1 yn+1 be the symmetric bilinear form
associated to q. Then Dx q(v) = 2b(x, v) and hence Tx Hn = (R x)⊥ where the
28 MARC BURGER STEPHAN TORNIER

orthogonal complement is taken with respect to b. Observe that since q(x) = −1


for all x ∈ Hn the space R x with form q is non-degenerate, hence the splitting
Rn+1 = R x ⊕ Tx Hn = R x ⊕ (R x)⊥ .
Silvester’s Inertia Theorem now implies that b restricted to Tx Hn is positive def-
inite. This constitutes an instance of a Riemannian metric as we shall see shortly.
The space Hn is called real hyperbolic space of dimension n, and SO0 (n, 1) will act
on it by isometries once we have properly defined the metric.

2. Riemannian Metrics, Covariant Derivative and Geodesics


2.1. Definitions and Examples. First of all, we define Riemannian metrics as
encountered in the last section and prove their existence.
Definition 2.1. Let M be a smooth manifold. A Riemannian metric on M is a map
g which to every point p ∈ M associates a scalar product gp : Tp M × Tp M →
R, satisfying the following smoothness condition: For any chart (U, ϕ) of M , the
functions U → R, p 7→ gp (∂i p, ∂j p) are smooth for all 1 ≤ i, j ≤ m.
In the context of Definition 2.1 recall that
∂(f ◦ ϕ−1 )
∂i (p)(f ) = (ϕ(p))
∂xi
for every f ∈ C ∞ (M ) and that (∂1 (p), . . . , ∂n (p)) is a basis of Tp M for all p ∈ M .
Remark 2.2. It will be useful to have the bundle definition of Riemannian metrics as
well: For a real vector space V , let S 2 (V ∗ ) denote the vector space of all symmetric
bilinear maps V ×V → R. Then one can define S 2 (T∗ M ) as a smooth vector bundle
of symmetric bilinear forms. That is, one introduces a manifold structure on
[
S 2 (T∗ M ) = {p} × S 2 ((Tp M )∗ )
p∈M

as in the case of the tangent and cotangent bundle. A Riemannian metric g on M


is then a smooth section g : M → S 2 (T∗ M ) such that gp := g(p) ≫ 0 for all p ∈ M .
Notationally, given a PRiemannian metric g on PmM , a chart (U, ϕ) and u, v ∈ Tp M
m
with coordinates u = i=1 ui ∂i (p) and v = j=1 vj ∂j (p) respectively, we have
X
gp (u, v) = ui vj gp (∂i (p), ∂j (p))
i,j

where we abbreviate gij (p) := gp (∂i (p), ∂j (p)). Recall that ui = (dxi )p (u) and
similarly for v ∈ Tp M . Then in tensor notation, the bilinear form which maps
(u, v) to ui vj is (dxi )p ⊗ (dxj )p . Thus, from the S 2 (T∗ M ) viewpoint on Riemannian
metrics, we have
X
gp = gij (p)(dxi ⊗ dxj )p .
i,j

The traditional
P (sloppy) way to refer to a Riemannian metric in local coordinates
is g = i,j gij dxi dxj .
We now turn to the existence of Riemannian metrics which is unclear, given that
we are asking for a highly non-zero section of a certain bundle and in view of what
we have learned about e.g. vector fields on the two-dimensional sphere. Yet we have
the following remarkable result whose proof is not even overly difficult.
Proposition 2.3. Every smooth manifold admits a Riemannian metric.
RIEMANNIAN GEOMETRY 29

Proof. First, consider a single chart (U, ϕ) of M . Then we have an associated basis
(∂1 (p), . . . , ∂m (p)) of Tp M for all p ∈ U whence for u, v ∈ Tp M we may define
m
X
gpU (u, v) := ui vi ,
i=1

employing the coordinates of u and v with respect to said basis. In other words, we
set gpU (∂i (p), ∂j (p)) := δij . This defines a Riemannian metric on U .
As to the whole manifold, let (Ui , ϕi )i∈I be a locally finite covering of M by
charts and let (fi )i∈I be a subordinate smooth partition of unity. For p ∈ M now
define X
gp := fk (p)gpUi .
i∈I

Since for every p ∈ M admits a neighbourhood Vp such that {i ∈ I | Ui ∩ Vp 6= ∅}


is finite, gp is a scalar product as a convex combination of scalar products. 

An alternative proof of Proposition 2.3 is to simply refer to the general version


of Whitney’s Embedding Theorem and restrict the Euclidean scalar product to the
tangent spaces of the embedding.
In a sense, Whitney’s theorem states that in order to study smooth manifolds
one has to look no further than submanifolds of Euclidean space. An interesting
question is whether this also holds for Riemannian manifold: Can every Riemannian
manifold be embedded isometrically into some Euclidean space? The answer is yes
and due to Nash.
Example 2.4. Arguably, the most immediate example of a Riemannian manifold
is Rn : For all p ∈ Rn and u, v ∈ Tp Rn = Rn set gp (u, v) = hu, vi where h−, −i
denotes the standard scalar product on Rn .
Earlier on, we have seen that differential forms can be pulled back via general
smooth maps and that vector fields can be pushed forward under some assumptions
on the smooth maps. To the assumptions on has to put on a smooth map for it
to allow pulling back a Riemannian metric are of intermediary nature: Let M and
N be smooth manifolds and let f : M → N be a smooth immersion. If h is a
Riemannian metric on N then
gp (u, v) := hf (p) (Dp f (u), Dp f (v))
where p ∈ M and u, v ∈ Tp M defines a Riemannian metric on g, denoted f ∗ (h).
Definition 2.5. Let (M, g) and (N, h) be Riemannian manifolds and let f : M → N
be a smooth map. Then f is an isometry if f is a diffeomorphism and g = f ∗ (h).
Given a Riemannian manifold (M, g), the set Iso(M, g) of isometries of M is
a subgroup of the group of diffeomorphisms of M . Whereas the latter is infinite-
dimensional from any point of view, Iso(M, g) is a Lie group by a result of Myers
and Steenrod. Later on we will look at the particularly intriguing case in which
Iso(M, g) acts transitively on M .
One of the points of having a Riemannian metric is to be able to measure the
length of certain curves.
Definition 2.6. Let (M, g) be a Riemannian manifold and let c : [a, b] → M be a
C 1 -curve. The length l(c) of c is
Z bq
l(c) := gc(t) (c′ (t), c′ (t)) dt.
a
30 MARC BURGER STEPHAN TORNIER

It is useful to extend Definition 2.6 to piecewise C 1 -curves: If c : [a, b] → M is


1
C on [ai , ai+1 ] for all i ∈ [0, n − 1] where a0 = a < a1 < · · · < an = b then we set
Pn−1
l(c) := i=0 l(c|[ai ,ai+1 ] ).
Note that we are defining the length of a curve using its instant velocities which
is quite the other way around as in calculus where there are length and time first,
then their quotient. We will see more instances of this reverse engineering. Anyway,
one shows that the length of a piecewise C 1 -curve does not depend on the choice of
parametrization. Also, one can always reparametrize such a curve by arc length in
which case kc′ (t)k = 1. One of the first problems in Riemannian geometry is to find
curves of shortest length between given points. We will later on characterize these
as having constant speed and “zero acceleration”. The term “acceleration” however
requires clarification as the notion of “second derivative of a curve” is just not there.
This will be remedied by introducing connections which provide a preferred way of
transporting tangent vectors along a curve. Whereas there are plenty connections in
general, Riemannian manifolds come with a natural one, the Levi-Civita connection.
Now, given a connected Riemannian manifold M and x, y ∈ M we define the
distance between a and b by
d(x, y) := inf{l(c) | c : [a, b] → M piecewise C 1 , c(a) = x, c(b) = y}.
Note that in a connected Riemannian manifold as above, there always is a piece-
wise C 1 -curve between any two given points: Find a sequence of coordinate charts
(Ui , ϕi )ni=0 with x ∈ U0 , y ∈ Un and Ui ∩ Ui+1 6= ∅.

···
U0 x b b y Un

Proposition 2.7. Let (M, g) be a connected Riemannian manifold and define the
map d : M × M → R by
d(x, y) := inf{l(c) | c : [a, b] → M piecewise C 1 , c(a) = x, c(b) = y}.
Then d is a metric on M inducing the present topology.
Although this does not present much difficulties we skip the proof and move on
to examples of Riemannian manifolds and construction methods.
Example 2.8. This example is to illustrate that for a sub- b

manifold M of Rn , the distance function induced from the


restricted scalar product is generally quite different from b
α b

the ambient distance function: For instance, consider the


sphere S n = {x ∈ Rn+1 | kxk = 1}. Then the distance
on Sn between two points x and y on the sphere at an
angle α is α whereas their distance in Euclidean space b b b

is 2(1 − cos α). Also, erasing a neighbourhood of 0 ∈ R2


naturally results in an altered metric.
In particular, note that given a metric space one can perfectly restrict the metric
to a subset and thereby produce a new metric space, this is not what happens when
one restricts a Riemannian metric to a submanifold.
Example 2.9. The hyperbolic spaces Hn introduced in Example 1.64 provide more
key Riemannian manifolds. Retain the notation of said example. Given x ∈ Hn ,
set hx (u, v) := b(u, v) for all u, v ∈ Tx Hn . Then h is a Riemannian metric on Hn .
Observe that for every g ∈ SO0 (n, 1) and x ∈ Hn , the map Dx g sends hx to hg(x) ,
i.e. SO0 (n, 1) acts by isometries. This yields an injective group homomorphism
RIEMANNIAN GEOMETRY 31

SO0 (n, 1) → Iso(Hn , h). Later on, we shall see that this homomorphism is almost
surjective. Another model of hyperbolic space Hn via the diffeomorphism
 
n n n 2yi 1 + kyk2
f : D := {y ∈ R | kyk < 1} → H , y 7→ ,
1 − kyk2 1 − kyk2
One verifies that f indeed ranges in Hn by computing q(f (y)) and that it admits a
smooth inverse. Furthermore, one obtains
Xn
∗ (dyi )2 4hu, vi
g := f (h) = 4 2
, i.e. gy (u, v) = .
i=1
(1 − kyk (1 − kyk2 )2
In this model the boundary S n−1 of Dn is infinitely far away from the origin: Indeed,
for 0 ≤ r < 1 one computes
Z r  
dt 1+r
lg ([0, re1 ]) = 4 2
= ln .
0 1−t 1−r
Another interesting observation is that whereas a random walk starting at the origin
in E2 will return to a neighbourhood of the origin infinitely many times. In the disk
model of H2 , however, random walks do not come back, due to the expanding space:
Whereas the area of a ball of radius R in E2 behaves like R2 , its behaviour with
respect to the hyperbolic metric is exponential in R. Also, whereas the area of a
Euclidean annulus of width R behaves like R, a hyperbolic annulus with larger
radius equal to one already contains a large fraction of the entire area.
For later use, we now record two formal definitions that allow us to construct
new Riemannian manifolds out of old.
Definition 2.10. Let (M, g) and (N, h) be a Riemannian manifolds. Then (N, h) is
a Riemannian submanifold of (M, g) if
(i) N is a submanifold of M , and
(ii) i∗ (g) = h where i : N → M is the canonical injection.
Definition 2.11. Let (M, g) and (N, h) be Riemannian manifolds. Then M × N is
a Riemannian manifold with metric g × h defined by
(g × h)(x,y) (u, v) = gx (u1 , v1 ) + hy (u2 , v2 )
where u = (u1 , u2 ) and v = (v1 , v2 ) with respect to T(x,y) (M × N ) = Tx M × Ty N .
It is also important to know how Riemannian metric behave with respect to
covering maps.
Definition 2.12. Let (M, g) and (N, h) be Riemannian manifolds and let p : N → M
be a smooth map. Then p is a Riemannian covering if
(i) p is a smooth covering, and
(ii) h = p∗ g.
In particular, a Riemannian covering is a local isometry. It is plain that if p :
N → M is a smooth covering of smooth manifolds and g is a Riemannian metric
on M then p is a Riemannian covering when N is equipped with the Riemannian
metric h = p∗ (g).
We now turn to the question under which circumstances a Riemannian metric
on N can be pushed down to M via p.
Proposition 2.13. Let (N, h) be a Riemannian manifold and let Γ be a group. Fur-
ther, let Γ × N → N be an action of Γ on N which is free, properly discontinuously,
and by isometries. Then there is a unique Riemannian metric g on the quotient
manifold M := Γ\N such that p : N → M is a Riemannian covering.
32 MARC BURGER STEPHAN TORNIER

Proof. Let x ∈ M and suppose y and y ′ are elements of p−1 (x). Then there is an
isometry γ ∈ Γ with γy = y ′ and the diagram
γ
N❇ /N
❇❇ ⑥⑥
❇❇ ⑥⑥

p ❇❇ ⑥⑥p
~ ⑥⑥
M
commutes. So does the diagram obtained by taking derivates.
Dy γ
Tx N❋ / Ty ′ N
❋❋ ①①
❋❋ ①①
❋ ①
Dy p ❋❋
{ ① Dy′ p

" ①
xM
Since p is a covering map, both Dy p and Dy′ p are vector space isomorphisms. We
use Dy p to introduce a scalar product on Tx M by gx := ((Dy p)−1 )∗ (hy ). Observe
that since γ is an isometry we have (Dy γ)∗ (hy′ ) = hy . Therefore
gx = ((Dy p)−1 )∗ (hy ) = ((Dy p)−1 )∗ (Dy γ)∗ (hy′ )
= (Dy γ ◦ (Dy p)−1 )∗ (hy′ ) = ((Dy′ p)−1 )∗ (hy′ )
which shows that gx is well-defined. The smoothness of the resulting Riemannian
metric is due to the fact that p is a local diffeomorphism. This way, p becomes a
Riemannian covering by construction. 

Example 2.14. We equip RnPwith the Riemannian metric coming from the standard
n
scalar product hx, yip := i=1 xi yi , denoted can. One can then show that the
Riemannian distance of x, y ∈ Rn given by
d(x, y) := inf{l(c) | c : [a, b] → M piecewise C 1 , c(a) = x, c(b) = y}.
p
coincides with the Euclidean distance of x and y given by kx−yk = hx − y, x − yi.
As a hint, consider the smooth, distance non-increasing projection onto the straight
line connecting two points or go back to the definition of the Riemann integral and
the mean value theorem.
Proposition 2.15. Every Riemannian isometry g of (Rn , can) is given by g : v 7→
Rv + a for some R ∈ O(n) and a ∈ Rn .
Sketch of Proof. By Example 2.14, an isometry g ∈ Iso(Rn ) is a bijection which
satisfies kg(x) − g(y)k = kx − yk for all x, y ∈ Rn . Now one can invoke the Mazur-
Ulam theorem.
Theorem 2.16 (Mazur-Ulam). Let V be a normed vector space over R and let
T : V → V be a bijection satisfying kg(x) − g(y)k = kx − yk for all x, y ∈ V . Then
T is of the form T (x) = Ax + b where a : V → V is linear.
By this theorem, we have g(v) = R(v) + a for all v ∈ Rn where R is linear
and kR(x) − R(y)k = kx − yk. By the parallelogram law, this implies R ∈ O(n):
hx, yi = (kx + yk2 − kx − yk2 )/4.
We now examine the group law on Iso(Rn , can): Every g ∈ Iso(Rn ) is represented
by a pair (R, a) in the set-theoretic cartesian product O(n) × Rn . Given isometries
g1 = (R1 , a1 ) and g2 = (R2 , a2 ) we compute
g1 g2 (v) = g1 (R2 v + a2 ) = R1 R2 v + R1 a2 + a1
RIEMANNIAN GEOMETRY 33

and thus g1 g2 = (R1 R2 , R1 a2 + a1 ). Thus, as a group, Iso(Rn , can) is isomorphic


to the semidirect product O(n) ⋉ Rn . The map r : O(n) ⋉ Rn → O(n) is a homo-
morphism whose kernel is the group of pure translations. Given a ∈ Rn we write
Ta (v) = v + a.
Riemannian coverings p : Rn → M are given by subgroups ΓIso(Rn ) which act
freely and properly discontinuously on Rn . In this context, we have the following.
Definition 2.17. A subgroup Γ ≤ Iso(Rn , can) is crystallographic if it acts properly
discontinuously and Γ\ Rn is compact. It is Bieberbach if it is crystallographic and
acts freely.
In particular, a Bieberbach group Γ ≤ Iso(Rn , can) defines a Riemannian cover-
ing pΓ : Rn → Γ\ Rn where Γ\ Rn is a compact manifold.
Example 2.18. Let P denote the subgroup of O(n) consisting of permutation matri-
ces and set Γ := {(R, γ) | R ∈ P, γ ∈ Zn } is crystallographic but not Bieberbach.
Bieberbach groups are more difficult to construct. The trivial ones are the content
of the following proposition.
Proposition 2.19. Let (a1 , . . . , an ) be a basis of Rn and Γ = {Tγ | γ ∈ Z a1 +
· · · Z an }. Then Γ is Bieberbach and Γ\ Rn is diffeomorphic to Tn = S 1 × · · · × S 1 .
Furthermore, the Riemannian manifolds Γ\ Rn and Γ′ \ Rn are isomorphic if and
only if there is R ∈ O(n) with R(Z a1 + · · · + Z an ) = Z a′1 + · · · + Z a′n where
Γ = {Tγ | γ ∈ Z a1 + · · · Z an } and Γ′ = {Tγ ′ | γ ′ ∈ Z a′1 + · · · Z a′n }.

Proof. Consider the map e : Rn → Tn given by x 7→ e2πix1 , . . . , e2πixn where
x = (x1 , . . . , xn )T ∈ Rn . The map e is a smooth covering and we have e(x) = e(y)
if and only if x−y ∈ Z a1 +· · ·+Z an , that is if and only if Tγ (x) = y for some Tγ ∈ Γ.
This implies that e : Γ\ Rn → Tn , being a local diffeomorphism and bijective, is a
diffeomorphism.
Assume now that f : Γ\ Rn → Γ′ \ Rn is an isometry. Since pΓ and pΓ′ are cover-
ing maps and Rn is simply connected, there exists a diffeomorphism fe : Rn → Rn
such that the following diagram commutes:
f
Γ\ Rn / Γ ′ \ Rn
O O
pΓ pΓ′

Rn / Rn .
fe

Now observe that fe is an isometry: Let gΓ and gΓ′ denote the Riemannian metrics
on Γ\ Rn and Γ′ \ Rn respectively for which pΓ and pΓ′ are Riemannian coverings.
Furthermore, let gcan denote the canonical Riemannian metric on Rn . Then
fe∗ (gcan ) = fe∗ p∗ ′ (gΓ′ ) = (pΓ′ ◦ fe)∗ (gΓ′ )
Γ
= (f ◦ pΓ )∗ (gΓ′ ) = p∗Γ f ∗ (gΓ′ ) = p∗Γ (gΓ ) = gcan .
By the above we conclude that fe = (R, a) for some R ∈ O(n) and a ∈ Rn . From
this one deduces that Γ and Γ′ are conjugate as subgroups of Iso(Rn ): feΓfe−1 = Γ′ .
This implies R(Z a1 + · · · + Z an ) = Z a′1 + · · · + Z a′n . 
Remark 2.20. The Riemannian manifolds described in Proposition 2.19 are called
flat tori because they are diffeomorphic to a torus and flat from a curvature point
of view as we shall see later. We now ask whether there is a "reasonable space"
parametrizing the set of flat tori up to isometry: Let R be the set of all subgroups of
Rn of the form Z a1 +· · ·+Z an where (a1 , . . . , an ) is a basis of Rn . Clearly, GL(n, R)
34 MARC BURGER STEPHAN TORNIER

acts on R: Given Λ = Z a1 + · · · + Z an and g ∈ GL(n, R we have g(Λ) = Z ga1 +


· · · + Z gan where (ga1 , . . . , gan ) is again a basis of Rn . This action is transitive: Fix
Λ0 = Z e1 + · · · + Z en . Given Λ = Z a1 + · · · + Z an set g = (a1 , . . . , an ) ∈ GL(n, R).
Then g(Λ0 ) = Λ. The stabilizer in GL(n, R) of Λ0 is given by
{g ∈ GL(n, R) | g(Λ0 ) = Λ0 }
−1
= {g ∈ GL(n, R) | ∀i, j ∈ {1, . . . , n} : gij ∈ Z, gij ∈ Z}
= {g ∈ GL(n, R) | ∀i, j ∈ {1, . . . , n} : gij ∈ Z and det g ∈ {±1}}.
Hence R ∼ = GL(n, R)/ GL(n, Z) as sets. Observe that GL(n, Z) is a discrete sub-
group of the Lie group GL(n, R) whence GL(n, R)/ GL(n, Z) has a canonical mani-
fold structure. Taking into account isometry classes one sees that at the end of the
day, flat tori are parametrized by the double quotient O(n)\ GL(n, R)/ GL(n, Z)
where GL(n, Z) acts on O(n)\ GL(n, R) on the right because O(n) is compact.
Whereas the above double quotient has a nice structure, equally natural questions
lead to double quotients where the left hand subgroup is non-compact, resulting in
the fact that the double quotient is not even a standard Borel space.
A less obvious Bieberbach group is Γ := hγ1 , γ2 i ≤ Iso(R2 ) where
           
x x 1 x x 0
γ1 = + and γ2 = = + .
y −y 0 y y 1
We claim that Γ is a Bieberbach group, i.e. it acts properly discontinuously, freely
and with compact quotient on R2 . Geometrically, we have
   
x x
γ2
y y

   
x x
γ1
y y

To get a geometric idea of the quotient, consider the domain

1
  
2
x
2 1
D D= ∈ R 0 ≤ x ≤ 1, |y| ≤
1 y 2

Any v ∈ R2 is Γ-quivalent to a point in D: Given v = (x, y)T , applying an


appropriate power k of γ2 yields γ2k (x, y)T = (x′ , y ′ ) with |y ′ | ≤ 21 . Then, consider

γ1l γ2k (x, y)T = γ1l (x′ , y ′ )T = (x′ + l, (−1)l y ′ )


so that 0 ≤ x′ + l ≤ 1 for appropriate l while |(−1)l y ′ | = |y ′ | ≤ 1/2. One easily
determines the identifications that Γ induces on the boundary of D, namely

1
2
D
1

The quotient Γ\ R2 is termed Klein bottle. Note that Γ contains the purely trans-
lational subgroup Γ′ := {Tγ | γ ∈ 2 Z e1 + Z e2 } and that the torus Γ′ \ R2 projects
with degree two onto Γ\ R2 . It is a good exercise to visualize this projection. Later
on, we will study the Riemannian geometries on these examples.
RIEMANNIAN GEOMETRY 35

An example of a Bieberbach group in three dimensions is the Hantsche-Wendt


group Γ := hγ1 , γ2 i ≤ Iso(R3 ) where
           
x −x 1/2 x x 1/2
γ1 y  = −y  + 1/2 and γ2 y  = −y +  0  .
z z 1/2 z −z 0
Interestingly, Γ\ R3 and S 3 have the same Betti numbers but are not diffeomor-
phic: They can be distinguished with integral homology for instance. The following
theorem of Bieberbach shows that it is always a good idea to look for a purely trans-
lational subgroup in order to prove cocompactness. Recall that r : Iso(Rn ) → O(n)
denotes the map which to an isometry associates its linear part.
Theorem 2.21 (Bieberbach). Let Γ\Iso(Rn ) be a crystallographic grup. Then
(i) r(Γ) is a finite subgroup of O(n),
(ii) Γ ∩ {Ta | a ∈ Rn } is Bieberbach, and
(iii) Γ acts freely on Rn if and only if it is torsion-free.
The proof of Theorem 2.21 requires several good ideas. We refer the reader
to [Cha12] and the classic [Wol11].
Riemannian Submersions. We have seen how Riemannian metrics behave with re-
spect to coverings. Now, we look at the more general case of submersions: Let
p : M → N be a submersion, i.e. Dx p : Tx M → Tp(x) N is surjective for all x ∈ M ,
and let g be a Riemannian metric on M . Then we can define the horizontal subspace
Hx := (ker Dx p)⊥ where the orthogonal complement is taken with respect to the
Riemannian metric. Clearly, Dx p|Hx : Hx → Tp(x) N is an isomorphism.
Definition 2.22. Retain the above notation. The submersion p is Riemannian if the
map Dx p|Hx : (Hx , g|Hx ) → (Tp(x) N, hp(x) ) is an isometry.
Proposition 2.23. Let (Mf, e
g) be a Riemannian manifold and let G × M f→M f be
f
a free, proper Lie group action by isometries of G on M . Then there is a unique
Riemannian metric g on the quotient M := G\M f such that p : M
f → M is a Rie-
mannian submersion.
Proof. The proof is quite analogous to the proof of Proposition 2.13: Let x ∈ M
and y, y ′ ∈ p−1 (x). Also, let g ∈ G be such that g(y) = y ′ . Then the following
diagram commutes:
Dy g
f
Ty M f
/ Ty ′ M
❋❋ ✇
❋❋ ✇
❋❋ ✇✇
Dy p ❋❋❋ ✇✇✇Dy′ p
# {✇✇
Tx M
gy′ ) we get Dy g(Hy ) = Hy′
In particular, Dy g(ker Dy p) = ker Dy′ p. Since (Dy g)∗ (e
where Hy = (ker Dy p)⊥ and Hy′ = (ker Dy p)⊥ . Now define
gx = ((Dy p|Hy )−1 )∗ (e
gy ).
Then the above shows that gx is independent of the choice of y ∈ p−1 (x). 
Example 2.24. Recall that Pn (C) = C∗ \ Cn+1 −{0} has a manifold structure with
charts arising from affine hyperplanes in Cn+1 . In fact, it is even a complex analytic
manifold. Whereas in the real case, the projection S n → (Z /2 Z)\S n = Pn (R) is a
covering, we Pobtain a non-discrete fibration in the complex case: Consider S 2n+1 =
{z ∈ C n+1
| n+1 2
i=1 |zi | = 1} and the inclusion S
2n+1
→ Cn+1 \{0}. The Lie group
1 2n+1
S = {z ∈ C | |z| = 1} acts on S by ζ(z1 , . . . , zn+1 ) = (ζz1 , . . . , ζzn+1 ). This
action is free and proper; in fact any smooth action of a compact Lie group is
36 MARC BURGER STEPHAN TORNIER

proper. Now, the inclusion S 2n+1 → Cn+1 \{0} induces a surjective submersion
p : S 2n+1 → Pn (C) and one can show that it induces a diffeomorphism
S 1 \S 2n+1 → Pn (C).
P
Let (z, w) = n+1
i=1 zi w i be the unitary scalar product on C
n+1
. Then the Euclidean
n+1
scalar product on the real vector space C is given by hu, wi = Re(u, w). We
now equip Cn+1 with the standard Riemannian metric coming from h−, −i and
S 2n+1 ⊆ Cn+1 with the induced Riemannian metric. Then for z ∈ S 2n+1 we have
Tz S 2n+1 = {u ∈ Cn+1 | hz, ui = 0} = {u ∈ Cn+1 | Re(z, u) = 0}.
and the orthogonal decomposition Cn+1 = R z ⊕ Tz S 2n+1 where both constituents
of the right hand side are real vector spaces. The S 1 -orbit through z ∈ S 2n+1 can
be parametrized by c : R → S 2n+1 , t 7→ eit · z. Hence ker Dz p = R ·i · z where
p : S 2n+1 → S 1 \S 2n+1 . Using Hz ⊆ Tz S 2n+1 we have the following orthogonal
decomposition by definition:
Tz S 2n+1 = R iz ⊕ Hz .
On Pn C we put the Riemannian metric coming from S 2n+1 or rather the restriction
to Hz . Thus, putting everything together, we have an orthogonal decomposition
Cn+1 = R z + R iz + Hz
| {z }
=C z

where Hz is of real codimension one in Tz S 2n+1 . Hence Hz is in fact orthogonal to


Cz for the unitary scalar product and hence a complex vector space. This is a good
example to keep in mind for a submersion with interesting additional structure.
In the context of the above example it is worth mentioning that one can do
differential geometry over the complex numbers but that things become more rigid.
For instance, there is no analogue of Whitney’s embedding theorem or embeddings
into projective space. In fact, complex submanifolds of complex projective space
are very special.
Lie Groups. As usual, things work particularly well in the context of Lie groups.
Example 2.25. Let G be a Lie group and let h−, −ie be a scalar product on Te G.
Then we can define hu, vig = hDg Lg−1 (u), Dg Lg−1 (v)ie for all u, v ∈ Tg G, produc-
ing a left-invariant Riemannian metric e
g on G. In particular, G can be viewed as a
subgroup of Iso(G, ge).
One might as well construct a right-invariant Riemannian metric on a Lie group.
In general, however, there need not be one which is bi-invariant.
Proposition 2.26. Let G be a Lie group. Then the following are equivalent.
(i) The Lie group G admits a bi-invariant Riemannian metric.
(ii) The image Ad(G) ⊆ GL(g) is contained in a compact subgroup.
Corollary 2.27. Let G be a compact Lie group. Then G admits a bi-invariant Rie-
mannian metric.
2.2. Covariant Derivative. This section truly kicks off the theory of Riemannian
geometry by introducing the crucial notion of connection, which was only under-
stood in the 1920’s by very few people, including É. Cartan. Nevertheless, it leads
to the Riemannian curvature tensor which was introduced much earlier, namely in
1853, by Riemann.
The problem we study is the following: Let X and Y be smooth vector fields on
Rn ; later on we shall be interested in general manifolds, of course. Is there a natural
notion of taking the derivative of Y with respect to X, that is the “variation of Y
RIEMANNIAN GEOMETRY 37

from the point of view of X”? One possible way to deal with this is to consider the
integral curve cp : (−ε, ε) → Rn of X passing through p ∈ M at time t = 0,
(
c′p (t) = X(cp (t))
,
cp (0) = p

and to look at the vectors Y (cp (t)) ∈ Tcp (t) Rn . We would like to define the first
order variation of the map t 7→ Y (cp (t)) near t = 0. If we had started out with
a general manifold we would be stuck at this point because the different tangent
spaces don’t know about each other, that is, we have no means to compare tan-
gent vectors at different points. In Rn , however, we can utilize the translations to
establish a privileged identification Tcp (t) Rn → Tp Rn . Using this identification we
consider Y (cp (t)) to be an element of Tp Rn and define
Y (cp (t)) − Y (p)
(∇X Y )(p) := lim
t→0 t
Writing Y = (Y1 , . . . , Yn ) and X = (X1 , . . . , Xn ) one verifies that
n
!
X ∂Yj (p)
(∇X Y )(p) = . . . , Xi (p), . . . ,
i=1
∂xi

its merely a directional derivative. The map ∇ : Γ(T Rn ) × Γ(T Rn ) → Γ(T Rn )


satisfies the following formal properties.
(i) ∇X Y − ∇Y X = [X, Y ] for all X, Y ∈ Γ(T Rn ).
(ii) ∇f X Y = f ∇X Y for all X ∈ Γ(T Rn ) and all f ∈ C ∞ (Rn ).
(iii) ∇X (f Y ) = (Xf )Y + f ∇X Y for all X, Y ∈ Γ(T Rn ) and f ∈ C ∞ (Rn ).
Using these properties, we now define connections in general. This will lead to
parallel transport which allows to identify tangent spaces at distinct points of a
manifold depending on a path connecting them, taking curvature into account. The
fact that our identification in Rn does not depend on a path connecting given points
resembles the vanishing of curvature.

Definition 2.28. Let M be a smooth manifold. A connection on M is an R-linear


map ∇ : Γ(TM ) × Γ(TM ) → Γ(TM ) which satisfies the following properties for all
X, Y ∈ Γ(TM ) and f ∈ C ∞ (M ):
(i) ∇f X Y = f ∇X Y ,
(ii) ∇X (f Y ) = (Xf )Y + f ∇X Y , and
(iii) ∇X Y − ∇Y X = [X, Y ].

Note that defining ∇ to be the zero map does not yield a connection due to the
second property. Hence existence of connections is not completely obvious. However,
if ∇1 and ∇2 are connections on a manifold M then so is g∇1 + (1 − g)∇2 for every
g ∈ C ∞ (M ). Therefore, using a partition of unity to patch together connections
defined on charts, one shows that every smooth manifolds admits many different
connections and thus many different notions of acceleration which mathematicians
had a hard time to accept.

Remark 2.29. Let ∇ be a connection on a smooth manifold M . Then the first


condition implies that for given Y ∈ Γ(TM ) the vector (∇X Y )(p) (p ∈ M ) only
depends on X(p) rather than the values of X in a neighbourhood of p ∈ M : In-
deed, let (U, ϕ) be a chart of M at p and recall that for every q ∈ U we ob-
tain a basis (∂1 (q), . . . , ∂m (q)) of Tq M . Consider ∂i ∈ Γ(TU ) (i ∈ {1, . . . , m})
as a smooth vector field. Then the local expression of X on U takes the form
38 MARC BURGER STEPHAN TORNIER

Pm Pm
X(q) = i=1 Xi (q)∂i (q). In vector field notation: X = i=1 Xi ∂i . Therefore
m
X
∇X Y = ∇Pm
i=1 Xi ∂i
Y = Xi ∇∂i Y
i=1
Pm
and hence (∇X Y )(p) = i=1 Xi (p)(∇∂i Y )(p). It follows that for X, X ′ ∈ Γ(TM )
with X(p) = X ′ (p) we have Xi (p) = Xi′ (p) and hence (∇X Y )(p) = (∇X ′ Y )(p).
The next theorem probably deserves to be called the fundamental theorem of
Riemannian geometry.
Theorem 2.30. Let (M, g) be a Riemannian manifold. Then there exists a unique
connection ∇ on M which for all X, Y, Z ∈ Γ(TM ) satisfies
Xg(Y, Z) = g(∇X Y, Z) + g(Y, ∇X Z)
where g(Y, Z) denotes the function M → R, p 7→ gp (Y (p), Z(p)).
Proof. First, we show uniqueness which suggests a formula to prove existence. Con-
sider the three equations that arise from the assumption through cyclic permutation
of the variables:


Xg(Y, Z) = g(∇X Y, Z) + g(Y, ∇X Z)
Y g(Z, X) = g(∇Y Z, X) + g(Z, ∇Y X)


Zg(X, Y ) = g(∇Z X, Y ) + g(X, ∇Z Y ).
Computing the sum of the first two right hand sides and subtracting the third yields
g(∇X Y, Z)+g(∇Y X, Z) + g(∇X Z, Y ) − g(∇Z X, Y ) + g(∇Y Z, X) − g(∇Z Y, X)
= 2g(∇X Y, Z) − g([X, Y ], Z) + g([X, Z], Y ) + g([Y, Z], X)
Therefore, we may record that 2g(∇X Y, Z) equals
Xg(Y, Z) + Y g(Z, X) − Zg(X, Y ) + g([X, Y ], Z) − g([X, Z], Y ) − g([Y, Z], X)
which shows uniqueness since g is non-degenerate. As for existence, let T (X, Y, Z)
denote the right hand side of the above equation which we write as
(g([X, Y ], Z) − Zg(X, Y )) + (Y g(Z, X) − g([Y, Z], X)) + (Xg(Z, Y ) − g([X, Z], Y ))
| {z } | {z } | {z }
A(X,Y,Z) B(X,Y,Z) B(Y,X,Z)

Observe that 2g(∇X Y, Z) = 2g(∇X Y, Z). Now, it remains to verify that T (X, Y, Z)(p)
only depends on Z(p) in which case the expression T (X, Y, Z)(p) is a linear form
in Z(p) on Tp M for fixed X and Y , to that there exists a unique vector ∇X Y (p) ∈
Tp M such that 2gp (∇X Y (p), Z(p)) = T (X, Y, Z)(p).
In order to verify this, we check what happens if we replace Z by f Z for some
f ∈ C ∞ (M ). Compute
A(X, Y, f Z) = g([X, Y ], f Z) + f Zg(X, Y )
= f (g([X, Y ], Z) + Zg(X, Y ))
= f A(X, Y, Z)
as well as
B(X, Y, f Z) = Y g(f Z, X) − g([Y, f Z], X)
= Y (f g(Z, X)) − g(Y (f )Z + f [Y, Z], X)
= Y (f )g(Z, X) + f Y g(Z, X) − Y (f )g(Z, X) − f g([Y, Z], X)
= f (Yg (Z, X) − g([Y, Z], X))
= f B(X, Y, Z).
RIEMANNIAN GEOMETRY 39

This shows T (X, Y, f Z) = f T (X, Y, Z) and hence as in Remark 2.29 implies that
T (X, Y, Z)(p) only depends on Z(p) for given X and Y . As explained above we
may hence define ∇X Y (p) using the equation that showed uniqueness. It is left as
an exercise to verify that the map ∇ so defined is indeed a connection. 

Definition 2.31. Let (M, g) be a Riemannian manifold. The connection of Theorem


2.30 is the Levi-Civita connection of (M, g).

We now aim to get a hold on the Levi-Civita connection by establishing a formula


in local coordinates. This dates back to Einstein and his work with Grossmann at
ETH Zurich around 1910. Recall that given a chart (U, ϕ) of M we have for every
p ∈ M a basis (∂1 (p), . . . , ∂m (p)) of Tp M given by

∂(f ◦ ϕ−1 )
∂i (p)(f ) = (ϕ(p)) for all f ∈ C ∞ (M ).
∂xi
Now consider the vector fields ∂i ∈ Γ(TU ) and let ∇ denote the Levi-Civita con-
nection of M . Then
m
X
∇∂j ∂k (p) = Γljk ∂l (p)
l=1

where the smooth functionsPΓljk : U → R are called Christoffel symbols and which
we now determine: Let g = m i,j=1 gij dxi ⊗ dxj be the expression of the Riemannian
metric in the chart (U, ϕ). Recall in particular that gij (p) = gp (∂i (p), ∂j (p)). We
are going to use the formula established in the proof of Theorem 2.30. To this end,
observe that [∂i , ∂j ] = 0 on U because of the definition of the ∂i and the fact that

∂2g ∂2g
= for all g ∈ C ∞ (ϕ(U )).
∂xi ∂xj ∂xj ∂xi
Employing said formula we obtain

2g(∇∂j ∂k , ∂i ) = ∂j g(∂k , ∂i ) + ∂k g(∂i , ∂j ) − ∂i g(∂j , ∂k )


= ∂j gki + ∂k gij − ∂i gjk

Substituting the definition of the Christoffel symbols yields


m
X
2 Γljk gli = ∂j gki + ∂k gij − ∂i gjk .
l=1

Let g −1 denote the matrix inverse of (gij ). Multiplying


Pm the equation by (g −1 )ir and
summing over i ∈ {1, . . . , m} we obtain using i=1 gli (g −1 )ir = δlr :
m
X
2Γrjk = (g −1 )ir (∂j gki + ∂k gij − ∂i gjk ) .
i=1

Proposition 2.32. Retain the above notation. Then


 
m
X X X
∇X Y =  Xj ∂j Yi + Γijk Xj Yk  ∂i
i=1 j j,k

where
m
X
2Γljk = (g −1 )ir (∂j gki + ∂k gij − ∂i gjk ) .
i=1
40 MARC BURGER STEPHAN TORNIER

Proof. We compute
!
X
∇X Y = ∇P i Xi ∂i Yk ∂k
k
X
= Xj ∇∂j (Yk ∂k )
j,k
X X
= Xj ∂j Yk · ∂k + Xj Yk ∇∂j ∂k
j,k j,k
 
X X X X
=  Xj ∂j Yi  ∂i + Xj Yk Γijk ∂i
i j j,k i
 
X X X
=  Xj ∂j Yi + Γijk Xj Yk  ∂i .
i j j,k


Example 2.33. Consider the following two examples.
(p)
(i) Consider (Rn , can). Then gij = δij for all p ∈ Rn whence Γijk = 0 for all
i, j, k ∈ {1, . . . , n}. We therefore have
 
X X
∇X Y =  Xj ∂j Yi  ∂i .
i j

In fact, ∇∂j ∂k = 0 for all j, k ∈ {1, . . . , n}.


(ii) Consider the upper-half-plane H 2 = {(x, y) ∈ R2 | y > 0} with the Rie-
mannian metric ((dx)2 + (dy)2 )/y 2 . For the computations, it is useful to
think of x and y in terms of x1 and x2 , and accordingly replace ∂x and ∂y
with ∂1 and ∂2 respectively. We then get
2g(∇∂j ∂k , ∂i ) = ∂j (δki x−2 −2 −2
2 ) + ∂k (δij x2 ) − ∂i (δjk x2 ).

For instance, let us compute ∇∂1 ∂1 : We have 2g(∇∂1 ∂1 , ∂1 ) = 0 and


2g(∇∂1 ∂1 , ∂2 ) = 2x−3 −1
2 = 2x2 g(∂2 , ∂2 ),

and therefore obtain


∇∂1 ∂1 = x−1
2 ∂2

and similarly
∇∂1 ∂2 = ∇∂2 ∂1 = −x−1
2 ∂1 and ∇∂2 ∂2 = −x−1
2 ∂2 .

Canonical Connection on a Riemannian Submanifold. A nice consequence


of the formula for the Levi-Civita connection is that it is simple to compute the
Levi-Civita connection of a Riemannian submanifold: Let (M, g) be a Riemannian
submanifold of a Riemannian manifold (N, h), that is M ⊆ N is embedded and for
all p ∈ M we have gp = hp |Tp M×Tp M .
Proposition 2.34. Retain the above notation and let ∇M and ∇N denote the Levi-
Civita connections of (M, g) and (N, h) respectively. Further, let X, Y ∈ Γ(TM )
and X ′ , Y ′ ∈ Γ(TN ) be smooth vector fields with X ′ |M = X and Y ′ |M = Y . Then
⊥p
(∇M N ′
X Y )(p) = ∇X ′ Y (p)

for all p ∈ M where ⊥p denotes the orthogonal projection from Tp N to Tp M .


RIEMANNIAN GEOMETRY 41

Proof. Let (U, ϕ) be a chart of N at p ∈ M ⊆ N such that ϕ(U ) = Cεn (0) and
ϕ(U ∩ M ) = Cεm (0) × (0)n−m . Now consider X, Y, Z ∈ Γ(TM ) and let X ′ , Y ′
and ZP ′
be extensions of X, Y and Z to open neighbourhoods of M in N . Then
n
X = i=1 Xi′ ∂i with Xi′ |U∩M = 0 for all i ∈ {m + 1, . . . , n} and similarly for Y ′

and Z . Now recall the formular for the bracket of vector fields in local coordinates:

X 
[X ′ , Y ′ ] = Xi′ ∂i Yj′ − Yi′ ∂i Xj′ ∂j
i,j

which, taking into account the vanishing of Xi′ and Yi′ on U ∩ M for i ≥ m + 1,
implies: [X ′ , Y ′ ]U∩M = [X, Y ]. The formula in the proof of Theorem 2.30 now
implies for p ∈ U ∩ M :

2hp (∇N ′ ′ M
X ′ Y (p), Z (p)) = 2gp (∇X Y (p), Z(p))

which implies the proposition since Z ′ (p) = Z(p). 

Covariant Derivative Along a Curve. In this section we discuss the covariant


derivative along curves which ultimately leads to the important notion of geodesics.

Definition 2.35. Let c : I → M be a smooth curve defined on some open interval


I ⊆ R. A vector field along c is a smooth map X : I → TM with X(t) ∈ Tc(t) M .

Retaining the above notation, our aim is to define "∇c′ (t) X" for a vector field
X along c. Let Γ(c∗ TM ) denote the vector space of such vector fields. It is also a
C ∞ (I)-module.

Remark 2.36. The notation c∗ TM stands for the pullback of the tangent bundle
TM by c. It is defined as

c∗ TM = {(t, v) ∈ I × TM | v ∈ Tc(t) M }.

Then c∗ TM is a smooth vector bundle with base I and a smooth vector field along
c is merely a smooth section of c∗ TM .

Theorem 2.37. Let (M, g) be a Riemannian manifold with Levi-Civita connection


∇ and let c : I → M be a smooth curve defined on an open interval I ⊆ R. Then
there is a linear map
D
: Γ(c∗ TM ) → Γ(c∗ TM )
dt
such that for all f ∈ C ∞ (I) and Y ∈ Γ(c∗ TM ):

D DY
(f Y ) = f ′ Y + f
dt dt
If furthermore X ∈ Γ(TM ) satisfies Y (t) = X(c(t)) for all t ∈ I we have

DY
(t) = (∇c′ (t) X)(c(t)).
dt
Proof. Let Y ∈ Γ(c∗ (TM )) be such that Y (t) = X(c(t)) for some X ∈ Γ(TM ). We
compute (∇c′ (t) X)(c(t)) in a chart (U, ϕ): Let ϕ(c(t)) = (x1 (t), . . . , xm (t)). Then
m
X
c′ (t) = (Dt c)(1) = x′i (t)∂i (c(t)).
i=1
42 MARC BURGER STEPHAN TORNIER

Indeed, we have

c′ (t)(f ) = (Dc(t) f )(c′ (t))


= Dc(t) ((f ◦ ϕ−1 ) ◦ ϕ)(c′ (t))
= Dϕ(c(t)) (f ◦ ϕ−1 )(Dc(t) ϕ)(c′ (t))
= Dx(t) (f ◦ ϕ−1 )Dt (ϕ ◦ c)(1)
Xm
∂(f ◦ ϕ−1 )
= (x(t))x′i (t)
i=1
∂x i
m
X
= x′i (t)∂i (c(t))(f ).
i=1

Using Proposition 2.32 we therefore have


 
m
X X X
(∇c′ (t) X)(c(t)) =  x′j (t)∂j Xi (c(t)) + Γijk (c(t))x′j (t)Xk (c(t)) ∂i .
i=1 j j,k
P
Now observe that j x′j (t)∂j Xi (c(t)) = Yi′ (t) whence
 
X X
(∇c′ (t) X)(c(t)) = Yi′ (t) + Γijk (c(t))x′j (t)Yk (t) ∂i
i j,k

Given the above computation, we now define for any Y ∈ Γ(c∗ TM ) and local
coordinates (U, ϕ):
 
Xm X
DY Yi′ (t) +
(t) := Γijk (c(t))x′j (t)Yk (t) ∂i
dt i=1 j,k

and leave the verifications to the reader. 

Theorem 2.37 allows us to make the following, central definition.

Definition 2.38. Let (M, g) be a Riemannian manifold with Levi-Civita connection


∇ and let c : I → M be a smooth curve defined on an open interval I ⊆ R. Further,
let D/dt : c∗ (TM ) → c∗ (TM ) be the associated linear map and X ∈ Γ(c∗ (TM )).
Then X is parallel if D/dt(X) = 0.

Example 2.39. Consider the Riemannian manifold (Rn , can). Let c : R → Rn be


smooth and write c(t) = (x1 (t), . . . , xn (t)). Then the map R → T Rn given by
t 7→ c′ (t) is a smooth vector field along c. We have
X n
Dc′
(t) = x′′i (t)∂i (c(t)) = c′′ (t).
dt i=1

Thus, Dc′ /dt is the acceleration.


If now M ⊆ Rn is a regular submanifold equipped with the induced Riemannian
metric g and if c : R → M is a smooth curve then Dc′ /dt = (c′′ (t))⊥ ; that is, the
covariant derivative of c′ (t) is the orthogonal projection of c′′ (t) onto Tc(t) M .

We now turn to the notion of parallel transport, which in Euclidean space looks
as follows.
RIEMANNIAN GEOMETRY 43

Proposition 2.40. Let c : I → M be a C 1 -curve defined on an open interval I ⊆ R,


and let t0 ∈ I. Then for every v ∈ Tc(t0 ) M there exists a unique parallel vector
field X : I → TM along c such that X(t0 ) = 0.
Proof. It suffices to prove the Proposition for every compact subinterval [t0 , T ] ⊆ R.
Let t0 < t1 < · · · < tn = T be such that c([ti , ti+1 ]) ⊆ Ui where (Ui , ϕi ) is a chart.
We now look at the condition for X to be parallel in the chart (U, ϕ): The condition
DX/dt = 0 is equivalent to the system
X
Xi′ (t) + Γijk (c(t))x′j (t)Xk (t) = 0 (i ∈ {1, . . . , m})
j,k

for all t in an open interval J ⊆ R with c(J) ⊆ U . For t′ ∈ J and (v1 , . . . , vm ) ∈ Rn


there is a unique solution of the above system on the whole of J with Xk (t′ ) = vk .
Applying this argument to each subinterval [ti , ti+1 ] yields the conclusion. 
Definition 2.41. Let c : [a, b] → M be a C 1 -map. The parallel transport Pc,a,b is the
linear map Tc(a) M → Tc(b) M which to every v ∈ Tc(a) M associates X(b) ∈ Tc(b)
where X : I → TM is the parallel vector field with X(a) = v and c is the restriction
of a C 1 -map defined on an open interval I ⊆ R containing [a, b].
Proposition 2.42. Retain the notation of Definition 2.41. Then the parallel transport
Pc,a,b : Tc(a) M → Tc(b) M is an isometry. More generally, if X, Y ∈ Γ(c∗ TM ) then
   
d D D
gc(t) (X(t), Y (t)) = gc(t) X(t), Y (t) + gc(t) X(t), Y (t) .
dt dt dt
Proof. The general formula is left as an exercise. To deduce that parallel transport
is an isometry, note that if X and Y are parallel we obtain (d/dt)g(X(t), Y (t)) = 0
and hence g(X(a), Y (a)) = g(X(b), Y (b)). 

One can recover the Levi-Civita connection from the parallel transport.
Proposition 2.43. Let (M, g) be a Riemannian manifold, X, Y ∈ Γ(TM ), p ∈ M
and c : (−ε, ε) → M an integral curve of X with c(0) = p. Then

d
(∇X Y )(p) = P −1 (Y (c(t)))
dt t=0 c,0,t
Proof. (Sketch). Define an operator D : Γ(TM ) × Γ(TM ) → Γ(TM ) by

d
DX Y (p) = P −1 (Y (c(t)))
dt t=0 c,0,t
and argue that D = ∇ using the uniqueness of the Levi-Civita connection: Consider
another Z ∈ Γ(TM ). Since Pc,0,t is an isometry we have
−1 −1
g(Pc,0,t (Y (c(t))), Pc,0,t (Z(c(t)))) = g(Y (c(t)), Z(c(t))).
Taking the derivative at t = 0 yields:
Xg(Y, Z) = g(DX Y, Z) + g(Y, DX Z).
at p. Also, one verifies that Df X Y = f DX Y and DX f Y = X(f )Y + f DX Y . 
44 MARC BURGER STEPHAN TORNIER

Example 2.44. Consider the sphere S 2 = {x ∈ R3 | x21 + x22 + x23 = 1} with its
induced Riemannian metric. In this example, we compute the parallel transport
along the circle
p p
c : [0, 2π] → S 2 , t 7→ ( 1 − r2 cos t, 1 − r2 sin t, r)
for r ∈ [0, 1). Specifically, we determine the map Tc,0,2π : Tc(0) S 2 → Tc(0) S 2 .

c
r

To this end, let x(t) = (x1 (t), x2 (t), x3 (t)) ∈ Tc(t) S 2 . That is, for all t ∈ [0, 2π]:
p p
(1) x1 (t) 1 − r2 cos t + x2 (t) 1 − r2 sin t + x3 (t)r = 0.
According to Example 2.39, the parallelity condition Dx/dt = 0 translates to x′ (t)
being orthogonal to Tc(t) S 2 in this context. That is
x′ (t) x′ (t) x′ (t)
(2) √ 1 =√ 2 = 3
1 − r2 cos t 1 − r2 sin t r
Taking the derivative of (1) with respect to t and substituting expressions for the
x′1 (t) and x′2 (t) summands in terms of x′3 resulting from (2) yields
p p x′ (t)
(3) − x1 (t) 1 − r2 sin t + x2 (t) 1 − r2 cos t + 3 = 0.
r
Taking the derivative of (3) and taking into account (1) as well as (2) yields
x′′3 + rx3 = 0
Hence x3 (t) = A cos(rt) + B sin(rt). From this one obtains using (2) formulas for
x′1 and x′2 which can be integrated, leading to
−A
x1 (t) = √ (sin(t) sin(rt) + r cos(t) cos(rt))
1 + r2
B
+√ (sin(t) cos(rt) − r cos(t) sin(rt)) + c1
1 − r2
and
−A
x2 (t) = √ (− cos(t) sin(rt) + r sin(t) cos(rt))
1 + r2
B
−√ (cos(t) cos(rt) + r sin(t) sin(rt)) + c2 .
1 − r2
The orthogonality condition hx(t), c(t)i = 0 leads to c1 = 0 = c2 . Finally, we get
 
−Ar −Br
x(0) = √ ,√ ,A ,
1 − r2 1−r

−r
x(2π) = √ (A cos(2πr) + B sin(2πr)) ,
1 − r2
r
√ (A sin(2πr) − B cos(2πr)) ,
1 − r2
A cos(2πr) + B sin(2πr)) ,
RIEMANNIAN GEOMETRY 45


Thus, if we take v1 := (r, 0, − 1 − r2 ) and v2 := (0, 1, 0) as a basis of Tc(0) S 2 , the
matrix of Tc,0,2π is given by
 
cos(2πr) sin(2πr)
.
− sin(2πr) cos(2πr)
2.3. Geodesics. In this section, we finally discuss the important notion of geodesics.
Let (M, g) be a Riemannian manifold with Levi-Civita connection ∇. Furthermore,
let D/dt denote the covariant derivative along a smooth curve c.
Definition 2.45. Retain the above notation. Let c : I → M , defined on an open
interval I ⊆ R, be a smooth curve. Then c is a geodesic if Dc′ /dt = 0 where
c′ : I → TM is the tangent velocity vector field along c.
In other words, a geodesic is a curve whose tangent vector field is invariant
with respect to parallel transport along itself. Intuitively, the position vector does
not experience any acceleration. We will shed light on the exact relation between
geodesics and distance-minimising curves later on but for the moment, convince
yourself that just as in driving, accelerating corresponds to making a detour.
Geodesics can also be looked at from a variational point of view: Given points
x, y ∈ M , consider all curves c connecting x to y and the functional L which to
a given curve associates its length. If one sets up things correctly then the Euler-
Lagrange equations arise from the condition that a curve c be a critical point of
the functional L, see e.g. [Spi79].
In local coordinates (U, ϕ) with ϕ(c(t)) = (x1 (t), . . . , xm (t)), the functions xi
(i ∈ {1, . . . , m}) associated to the geodesic c are the solution of the differential
equation
X
(GLC) x′′i (t) + Γijk (x(t))x′k (t)x′j (t) = 0.
j,k

Example 2.46. Before going into the theory of geodesics, we collect some examples.
(i) Let M ⊆ (Rn , can) be a regular submanifold with the induced Riemannian
metric. Then a smooth curve c : I → M ⊆ Rn , defined on an open interval
I ⊆ R is a geodesic in M if and only if the acceleration vector c′′ (t) is
orthogonal to Tc(t) M for all t ∈ I: Indeed, if ∇ is the Levi-Civita connection
on M and ∇ is the one on Rn we know that c′′ (t) = Dc′ /dt and Dc′ /dt =
(Dc′ /dx)⊥c(x) .
(ii) Geodesics in (Rn , can) are of the form c(t) = tv + w for some v, w ∈ Rn as
in this case (GLC) simply says x′′i (t) = 0.
(iii) Consider S n ⊆ Rn+1 with the induced Riemannian metric. A great circle
is determined by a pair u, v ∈ S n of orthogonal vectors via
c(θ) = cos(θ)u + sin(θ)v.
Then c′ (θ) = − sin(θ)u + cos(θ)v and c′′ (θ) = − cos(θ)u − sin(θ)v = −c(θ).
As a result, Dc′ /dt is orthgonal to Tc(θ) S n and hence c is a geodesic of S n .
The local existence and uniqueness of geodesics follows from the following well-
known theorem.
Theorem 2.47. Let Ω ⊆ Rn be open and let F : Ω × Rn → Rn be a smooth map.
Then for every (s0 , v0 ) ∈ Ω × Rn there is a neighbourhood E × V ⊆ Ω × Rn of
(s0 , v0 ) and δ > 0 such that for every (s, v) ∈ E × V there is a unique smooth curve
xs,v : (−δ, δ) → Ω satisfying
(i) x′′ (t) = F (x(t), x′ (t)),
(ii) x(0) = s and x′ (0) = v.
Furthermore, the map E × V × (−δ, δ) → Ω, (s, v, t) 7→ xs,v (t) is smooth.
46 MARC BURGER STEPHAN TORNIER

We now apply Theorem 2.47 to our situation.


Corollary 2.48. Let (M, g) be a Riemannian manifold with Levi-Civita connection
∇. Then every p0 ∈ M admits a neighbourhood U ⊆ M , δ > 0 and ε > 0 such
that for every p ∈ U and v ∈ Tp M with |v| < ε there is a unique geodesic c(p,v) :
(−δ, δ) → M such that c(p,v) (0) = p and c′(p,v) (0) = v. Moreover, the map

C : {(p, v) ∈ TU | |v| < ε} × (−δ, δ) → M, ((p, v), t) → c(p,v) (t)


is smooth.
The following homogeneity lemma often occurs underrated in the literature. It
critically depends on the form of (GLC).
Lemma 2.49. Retain the notation of Corollary 2.48. Assume that the geodesic
c(p,v) is defined on (−δ, δ) and let a > 0. Then c(p,av) is defined on (−δ/a, δ/a) and
c(p,av) (t) = c(p,v) (at).
Proof. To simplify notation, we put η(t) = c(p,v) (at) and γ(t) = c(p,v) (t). Observe
that η(0) = p and η ′ (0) = av. We show that η is a geodesic, i.e. Dη ′ /dt = 0 where
the covariant derivative is taken along η. Taking a local chart (U, ϕ) and setting
ϕ(γ(t)) + (x1 (t), . . . , xm (t)) and ϕ(η(t)) = (y1 (t), . . . , ym (t))
We know that yi (t) = xi (at). Therefore:
X X
yi′′ (t) = Γijk (η(t))yk′ (t)yj′ (t) = a2 x′′i (at) + Γijk (γ(at))ax′k (at)ax′j (at)
j,k j,k
 
X
= a2 x′′i (at) + Γijk (γ(at))x′k (at)x′j (at) = 0
j,k

which proves the assertion. 

Corollary 2.50. Retain the above notation. Then every p0 ∈ M admits a neigh-
bourhood U ⊆ M and ε > 0 such that for all p ∈ U and v ∈ Tp U with |v| < ε there
is a unique geodesic c(p,v) : (−2, 2) → M with c(p,v) (0) = p and c′(p,v) (0) = v.

Proof. This follows from the homogeneity Lemma 2.49: If c(p,v) is defined on (−δ, δ)
then ε(p,δv/2) is defined on (−2, 2). 

In the context of Corollary 2.50, let


Ω := {(p, v) ∈ TM | c(p,v) is defined on (−2, 2)}.
We have seen that Ω is a neighbourhood of the trivial section of TM .
Definition 2.51. Let (M, g) be a Riemannian manifold with Levi-Civita connection
∇ and retain the above notation. The exponential map is
exp : Ω → M, exp((p, v)) = c(p,v) (1).
We denote its restriction to Ω ∩ Tp M by expp .
As a consequence of Corollary 2.50, the exponential map is smooth.
Proposition 2.52. Retain the above and set φ : Ω → M × M , (p, v) 7→ (p, expp v).
Then φ is smooth and for every (p0 , 0) ∈ Ω the map φ is a diffeomorphism from a
neighbourhood of (p0 , 0) to its image in M × M .
RIEMANNIAN GEOMETRY 47

Proof. The map φ is smooth because its components are. For the second state-
ment we employ the inverse function theorem for which end we have to compute
D(p0 ,0) φ: Let (U, ϕ) be a chart of M at p0 ∈ M and recall that it yields a basis
(∂1 (p), . . . , ∂m (p)) of Tp U for all p ∈ U . We will also use the diffeomorphism
m
!
X
U × Rm → TU, (p, y) 7→ p, yi ∂i (p)
i=1
Pm
Slightly pedantic, we put ψ(p, y) := φ(p, v) where v = i=1 yi ∂i (p). Finally, we
identify T(p,0) (U × Rm ) with Tp U ⊕ Rm where we take
(∂1 (p), 0), . . . , (∂m (p), 0), (0, ∂1 y), . . . , (0, ∂m y)
with ∂i y = ∂/∂yi |y=0 as basis of the right hand space. Now, let’s consider the
variation of ψ in the direction of U , keeping the other fixed: We have
ψ(c(p,v) (t), 0) = (c(p,v) (t), c(p,v) (t))
and therefore D(p,0) ψ(v, 0) = (v, v) ∈ Tp M ⊕Tp M . For the other variation, consider
the smooth path t 7→ (p, ty) in U × Rm with tangent vector (0, y) ∈ Tp M ⊕ Rm .
Then
ψ(p, ty) = (p, c(p,tv) (1)) = (p, c(p,v) (t))
Pm
where v = i=1 yi ∂i (p). Thus D(p,0) ψ(0, y) = (0, v). Identifying Tp M × Tp M with
Rm ⊕ Rm using the basis
((∂1 (p), 0), . . . , (∂m (p), 0), (0, ∂1 (p)), . . . , (0, ∂m (p)))
the matrix of D(p,0) ψ is
 
Id 0
D(p,0) ψ = .
Id Id

Corollary 2.53. Retain the above notation. For every p0 ∈ M there is an open
neighbourhood U of p0 in M and ε > 0 such that the following hold.
(i) For every (x, y) ∈ U × U there is a unique v ∈ Tx M with |v| < ε and
expx (v) = y.
In this case, let c(x, y, t) := expx (tv).
(ii) The map U × U × (−2, 2) → M, (x, y, t) 7→ c(x, y, t) is defined and smooth.
(iii) For every x ∈ U the map expx : B(0, ε) → M is a diffeomorphism onto its
image.
Proof. Retaining the above notation, let W ⊆ Ω be an open neighbourhood of
(p0 , 0) ∈ Ω such that φ : W → φ(W ) is a diffeomorphism. It is easy to see from
local triviality of TM that there exists an open neighbourhood V of p0 ∈ M and
ε > 0 such that T<ε V := {(p, v) ∈ TM | p ∈ V, |v| < ε} ⊆ W . In particular,
φ(T<ε V ) is an open neighbourhood of (p0 , p0 ) = φ(p0 , 0) ⊆ M × M . Hence there is
an open neighbourhood U of p0 ∈ M with U × U ⊆ φ(T<ε V ). In particular U ⊆ V .
This pair (U, ε) satisfies the conclusions. 
The infinitesimal definition of geodesics has the disadvantage that its relation to
distance-minimising curves between points remains unclear at the moment. On the
plus side, it enables one to prove the following which certainly would not hold if
geodesics had been defined as curves of shortest lengths between points: Think of
the projection of a straight line onto the torus.
Corollary 2.54. Let (N, h) and (M, g) be Riemannian manifolds. Furthermore, let
p : (N, h) → (M, g) be a Riemannian covering. Given a smooth curve c : I → M ,
let e
c : I → N be any lift of c, i.e. p ◦ e
c = c. Then c is a geodesic if and only if e
c is.
48 MARC BURGER STEPHAN TORNIER

For instance, Corollary 2.54 can be used to understand the geodesics of Bieber-
bach manifolds.

2.4. Bi-invariant Metrics on Lie Groups. In this section, we illustrate many


of the notions introduced so far for Lie groups G that admit a bi-invariant metric.
Recall that this means that all left- and right-translations of G are isometries. In
turn, this implies that h−, −ie is invariant under Ad(G). Conversely, recall that an
Ad(G)-invariant scalar product on Te G gives rise to a bi-invariant metric via left-
and right-translation.
Example 2.55. We recall that any compact Lie group admits a bi-invariant metric.
Thus all orthogonal groups do. For instance, consider G = O(n) ⊆ GL(n, R) with
Lie(O(n)) = TId G = {X ∈ Mn,n (R) | X + X T = 0}.
As for all matrix Lie groups, we have Ad(g)X = gXg −1 . Given X, Y ∈ Te G =
Lie(G), set hX, Y i = − tr(XY ). This is a bilinear, symmetric form. In addition, we
have
X
hX, iX = − tr(XX) = tr((−X)X) = tr(X T X) = x2ij ≥ 0
i,j

which proves positive-definiteness. Furthermore,


hAd(g)X, Ad(g)Y i = − tr(gXg −1 gY g −1 ) = − tr(gXY g −1 ) = − tr(XY ) = hX, Y i.
Hence h−, −ie determines a bi-invariant metric on O(n).
We now show that for the class of Lie groups under consideration, the Rie-
mannian and the Lie group exponential maps coincide which yields a very explicit
understanding of all geodesics, particularly for matrix Lie groups.
Theorem 2.56. Let G be a Lie group admitting a bi-invariant Riemannian metric.
Then the Riemannian exponential is defined on the whole of TG and its restriction
expe : Te G → G coincides with the Lie group exponential.
Our proof of Theorem 2.56 relies on the following two lemmas.
Lemma 2.57. Let G be a Lie group admitting a bi-invariant Riemannian metric.
Then i : G → G, g 7→ g −1 is an isometry.
Proof. We write the relation gi(g) = e in the following way: The map m ◦ ∆ is
constant and equal to e ∈ G where ∆ : G → G × G, g 7→ (g, i(g)). In particular,
we have for all h ∈ G and v ∈ ThG: Dh (m ◦ ∆)(v) = 0. Note that Dh (m ◦ ∆) =
D∆(h) m ◦ Dh ∆ and
D(h1 ,h2 ) m(v1 , v2 ) = Dh1 Rh2 (v1 ) + Dh2 Lh1 (v2 ).
Indeed, consider for instance m(cv1 (t), h2 ) = cv1 (t)h2 . Also, Dh ∆(v) = (v, ∆h i(v)).
Hence 0 = D(h,h−1 ) m(v, Dh i(v)) = Dh Rh−1 (v) + Dh−1 Lh (Dh i(v)) whence
Dh−1 Lh (Dh i(v)) = −Dh Rh−1 (v)
which proves the assertion. More neatly, we record Dh i = −De Lh−1 ◦ Dh Rh−1 . In
particular, De i = −Id. 

Lemma 2.58. Retain the above notation. For every v ∈ Te G the maximal in-
terval of definition of the geodesic c(e,v) is the whole of R, and c(e,v) (t1 + t2 ) =
c(e,v) (t1 )c(e,v) (t2 ) for all t1 , t2 ∈ R.
RIEMANNIAN GEOMETRY 49

Proof. Fix v ∈ Te G and let (a, b) containing 0 ∈ R be the maximal interval of


definition of c(e,v) . Since i is an isometry with De i = −Id we have that i(c(e,v) ) is
a geodesic with velocity −v at t = 0. Hence we have i(c(e,v) ) = c(e,−v) by unique-
ness which implies that, whenever defined, c(e,v) (t)−1 = i(c(e,v) (t)) = c(e,−v) (t) =
c(e,v) (t). Thus a = −b. Now let ε > 0 be such that expe : B(0, ε) → G is a diffeo-
morphism onto its image and consider for 0 < t0 < ε: Γ(s) = c(t0 )c(s) = Lc(t0 ) c(s).
By left-invariance of the metric, Γ(s) is a geodesic which is defined on (−b, b). Note
that Γ(−t0 ) = c(t0 )c(−t0 ) = e and Γ(0) = c(t0 ) and observe that s 7→ c(t0 + s) is a
geodesic which at s = −t0 takes the value e ∈ G and at s = 0 the value c(t0 ). By
Corollary 2.53 we deduce that Γ(s) = c(s + t0 ). Since Γ is defined on (−b, b), this
would extend the domain of definition of c beyond b if b was finite since t0 + b > b.
This implies b = ∞, i.e. c(e,v) is defined on the whole of R and
c(e,v) (t0 )c(e,v) (s) = c(e,v) (t0 + s) for all s ∈ R, |t0 | < ε.
Now let t be arbitrary and take n ∈ N \{0} such that |t|/n < ε. Then
     
t (n − 1)t t (n − 1)t
c(e,v) (t + s) = c(e,v) + + s = c(e,v) c(e,v) +s
n n n n
 n
t
= c(e,v) c(s).
n
In particular, for s = 0 we obtain c(e,v) (t) = c(e,v) (t/n)n which plugged into the
above implies c(e,v) (t + s) = c(t)c(s). 
The proof of Theorem 2.56 is now no longer difficult.
Proof. (Theorem 2.56). Let v ∈ Te G and let Xv be the left-invariant vector field
determined by v. Furthermore, let c(e,v) be the geodesic determined by v. We need
to show that c(e,v) is an integral curve for Xv : Write c(e,v) (s)c(e,v) (t) = c(s + t), i.e.
Lc(e,v) (s) (c(e,v) (t)) = c(e,v) (s + t). Taking the derivative with respect to t yields
Dc(e,v) Lc(e,v) (s) (c′(e,v) (t)) = c′(e,v) (s + t).
Evaluating at t = 0 proves indeed Xv (c(s)) = De Lc(e,v) (s) (c′(e,v) (0)) = c′ (s). 
Example 2.59. Returning to our example
P∞ O(n), the Riemannian exponential map
is given by Lie(O(n)) → O(n), X 7→ n=0 X n /n!. In particular,
X∞ n n
t X
c(Id,X) (t) = .
n=0
n!
Therefore, length(c(Id,X) [0, T ]) = T kXk. The matrix exponential can be computed
via diagonalization and conjugation. In the present example, one can for instance
use this to show that the exponential map is surjective and to determine periodic
geodesics and their length.
Corollary 2.60. Let G be a Lie group admitting a bi-invariant Riemannian metric.
Let ∇ denote its Levi-Civita connection and let X, Y be left-invariant vector fields
on G. Then ∇X Y = [X, Y ]/2.
Proof. Let Z be any left-invariant vector field and c : R → G any integral curve of
Z. If c(0) = e then c is a geodesic. If c(0) 6= e it is nevertheless a geodesic by left-
invariance of Z, thus Dc′ /dt = 0. Hence ∇c′ (t) Z = 0 since c′ (t) = Z(c(t)). For the
same reason, ∇Z Z = 0. In particular ∇X+Y (X +Y ) = 0 which by expanding implies
0 = ∇X Y + ∇Y X = 0. Combining this with the fact that ∇X Y − ∇Y X = [X, Y ]
yields the assertion. 
Later on, we will also examine the Riemannian curvature tensor in this setting
and use it to derive global properties, e.g. information on the fundamental group.
50 MARC BURGER STEPHAN TORNIER

2.5. Geodesics and Distance. In this section, we characterize geodesics as being


locally length minimizing. The example of the sphere and the torus show that in
general geodesics are not globally length-minimizing.

Now, let (M, g) be a Riemannian manifold and fix p0 ∈ M . An open neighbour-


hood U of p0 ∈ M for which there is ε > 0 satisfying Corollary 2.53 is totally
normal. Recall that in this case we have for all p ∈ U : The exponential map
expp : B(0, ε) ⊆ Tp M → expp (B(0, ε))
is a diffeomorphism onto its image, which contains U . In this context, we now prove
the following.
Theorem 2.61. Retain the above notation. In particular, let U be a totally normal
neighbourhood of p0 ∈ M . Then
(i) for all p, q ∈ U there is a unique geodesic c of length l(c) < ε connecting p
to q, and
(ii) for any piecewise C 1 -curve γ : [0, 1] → M with γ(0) = p and γ(1) = q we
have l(γ) ≥ l(c), with equality if and only if γ equals c up to reparametri-
sation.
The main ingredient in the proof of Theorem 2.61 is a lemma of Gauss which
expresses the Riemannian metric in polar coordinates. As preparation, consider
p ∈ M and ε > 0 such that expp : B := B(0, ε) → M is a diffeomorphism onto
its image B ′ . Further, let S n−1 := {v ∈ Tp M | kvk = 1}. We introduce polar
coordinates on B\{0} via
expp
B\{0} / B ′ \{p}
O q 8
qqqqq
q
qqq f
(0, ε) × S n−1

where f (r, v) := expp (rv).

Lemma 2.62 (Gauss). Retain the above notation. The decomposition


T(r,v) ((0, ε) × S n−1 ) = Tr (0, ∞) ⊕ Tv S n−1
is orthogonal with respect to f ∗ (g)(r,v) . In these coordinates,

f ∗ (g)(r,v) = dr2 + h(r,v)

where h(r,v) is a scalar product on Tv S n−1 depending on r.

Proof. First of all, observe that the restriction to Tr (0, ∞) of f ∗ (g)(r,v) equals dr2 .
Indeed, the map
cv : (0, ε) → B ′ , r 7→ expp (rv) = f (r, v)
is a geodesic parametrized by arc length. To show the asserted orthogonality, we
consider certain vector fields: Let X be a smooth vector field on S n−1 , considered as
RIEMANNIAN GEOMETRY 51

a map S n−1 → Tp M , v 7→ X(v) satisfying hv, X(v)i = 0 for all v ∈ S n−1 . It induces
e v) := rX(r, v) on B\{0} via the polar coordinatization.
the vector field X(r,

rX(v)
X(v)

e be the direct image of X


Furthermore, we let Y := expp,∗ (X) e via the diffeomorphism
f , that is
Y (f (r, v)) = D(r,v) f (X(r, v)) = Drv expp (rX(v)).
Now, let ∂/∂r denote the radial vector field on (0, ε × S n−1 .
Then the lemma follows if we show that f∗ (∂/∂r) and Y are orthogonal at every
point f (r, v). To this end, observe that
   
∂ ∂
f∗ (f (r, v)) = f∗ (cv (r)) = c′v (r) ∈ Tf (r,v) M.
∂r ∂r
We therefore have using Proposition 2.42 and the properties of the Levi-Civita
connection
   
d ∂ d
g Y (f (r, v)), f∗ (f (r, v)) = g(Y (cv (r)), c′v (r))
dr ∂r dr
= g(∇c′v Y, c′v ) + g(Y, ∇c′v c′v )
| {z }
=0
= g(∇Y c′v , c′v ) + g([c′v , Y ], c′v )
1
= Y g(c′v , c′v ) + g([c′v , Y ], c′v ).
2
Since cv is a geodesic, g(c′v , c′v )cv (t) is constant in t with initial value g(v, v) = 1 at
t = 0. Thus the map (r, v) 7→ g(c′v , c′v )cv (r) is constant and hence Y g(c′v , c′v ) = 0. It
therefore remains to evaluate [c′v , Y ]: We have
 

[c′v , Y ] = f∗ ,X
∂r
A simple computation yields [∂/∂r, X] = 0. Overall, we conclude
  
d ∂
g Y, f∗ = 0.
dr ∂r
Since Y (0) = 0 this concludes the lemma. 

We now turn to Theorem 2.61

Proof. (Theorem 2.61). Regarding the first assertion, we already know that there
is a unique v ∈ Tp M with kvk < ε and cv (1) = q. Clearly, l(cv ([0, 1])) = kvk <
ε. Conversely, assume there is w ∈ Tp M and t > 0 such that cw (t) = q and
l(cw ([0, t])) < ε. Since l(cw ([0, t])) = tkwk, setting u := tw yields
cu (1) = ctw (1) = cw (t) = q
by homogeneity of geodesics. And since kuk = tkwk < ε we conclude u = v by
uniqueness of geodesics.
For the second assertion, assume that γ : [0, 1] → M is a piecewise C 1 -curve with
γ(0) = p and γ(1) = q. Then either γ([0, 1]) is contained in B ′ or there is s ∈ (0, 1)
52 MARC BURGER STEPHAN TORNIER

such that γ[0, s) ⊆ B ′ and γ(s) 6∈ B ′ . For 0 ≤ t ≤ s, write γ(t) = f (r(t), v(t)). Then
Lemma 2.62 implies
Z sq
l(γ) > gγ(t) (γ ′ (t), γ ′ (t)) dt
0
Z s Z s

≥ |r′ (t)| dt ≥ r′ (t) dt = ε.
0 0
Therefore, l(γ)) > ε > l(c). Suppose now that γ([0, 1]) is contained in B ′ . Again,
using polar coordinates, we have
Z 1q
l(γ) = [r′ (t)2 + h(r(t),v(t)) (v ′ (t), v ′ (t))] dt ≥ r(1) − r(0) = l(c).
0
Moreover, equality occurs if and only if h(r(t),v(t)) (v ′ (t), v ′ (t)) = 0 for all 0 ≤ t ≤ 1,
i.e. v ′ (t) = 0 whence v(t) = v for all t ∈ [0, 1], and r is monotone. Assuming that
γ is parametrized proportional to arc length with velocity l, we get γ(t) = f (lt, v)
and γ(t/l) = cv (t). 
Corollary 2.63. Retain the above notation. A piecewise C 1 -curve γ : I → M
parametrized proportional to arc length is a geodesic if and only if for all t ∈ I
there is ε > 0 such that d(c(t − ε), c(t + ε)) = l(c([t − ε, t + ε])).
Proof. This is due to Theorem 2.61 if γ is C 1 and is left as an exercise otherwise. 
Remark 2.64. As mentioned in the introduction, the examples of the sphere and
the torus show that in general geodesics are not globally length minimizing.
Definition 2.65. Retain the above notation. A geodesic c : [a, b] → M is minimal if
l(c([a, b])) = d(c(a), c(b)).
Proposition 2.66. Retain the above notation. Let c : [a, b] → M be piecewise C 1 .
If l(c) = d(c(a), c(b)) and c is parametrized proportional to arc length then c is a
geodesic.
Proof. Observe that if a ≤ a′ ≤ b′ ≤ b then l(c([a′ , b′ ])) = d(c(a′ ), c(b′ )). Indeed,
otherwise l(c([a′ , b′ ])) > d(c(a′ ), c(b′ )) and there is by definition a piecewise C 1 -
curve γ : [a′ , b′ ] → M with l(γ) < l(c([a′ , b′ ])), γ(a′ ) = c(a′ ) and γ(b′ ) = c(b′ ). But
then we have for the concatenation η of c|[a,a′ ] , γ and c|[b′ ,b] :
l(η) = l(c([a, a′ ])) + l(γ) + l(c([b′ , b]))
= l(c) − l(c([a′ , b′ ])) + l(γ)
< l(c) = d(c(a), c(b)).
Combining this observation with Corollary 2.63 implies the proposition. 
Normal Coordinates. As before, let (M, g) be a Riemannian manifold with Levi-
Civita connection ∇. Recall that in a local coordinate system (U, ϕ) of M we have
defined the Christoffel symbols Γijk by
X
∇∂j ∂k = Γij,k ∂i .
i
P P
Since ∇∂j ∂k − ∇∂k ∂j = [∂j , ∂k ] = 0 we get ∇∂k ∂j = i Γikj ∂i = i Γijk ∂i and
hence Γikj = Γijk .
Now let p ∈ M and ε > 0 such that expp : B(0, ε) → U ⊆ M is a diffeomorphism
onto an open neighbourhood U of p. Choosing an orthonormal basis e1 , . . . , em of
(Tp M, gp ) and setting ϕ := exp−1 p : U → Tp M , we obtain local coordinates at
p ∈ M . In these coordinates we have
(i) gp (∂i (p), ∂j (p)) = δij , and
RIEMANNIAN GEOMETRY 53

(ii) ∇∂i ∂j (p) = 0


Indeed, for (i) note that by definition D0 expp (ei ) = ∂i (p) and since D0 expp =
Id we get gp (∂i (p), ∂j (p)) = gp (ei , ej ) = δij . For the second assertion, let v ∈
Tp M and cv : I → M , t 7→ cv (t) the corresponding geodesics. Then by definition
p (cv (t)) = tv, i.e. ϕ(cv (t)) = (x1 (t), . . . , xm (t)) with xi (t) = tvi . Since cv is a
exp−1
geodesic we have for all i ∈ {1, . . . , m}:
X
x′′i (t) + Γikj (cv (t))x′k (t)x′j (t) = 0.
j,k

Evaluating at t = 0 yields
X
Γikj (p)vk vj = 0.
k,j

Since Γikj is symmetric in (k, j) and the above holds for any vector v ∈ B(0, ε) we
conclude Γikj (p) = 0.

2.6. The Hopf-Rinow Theorem. Let (M, g) be a connected Riemannian mani-


fold. We have seen that one can define a distance on M by setting
d(p, q) := inf{l(c) | c : [0, 1] → M is piecewise C 1 , c(0) = p, c(1) = q}
for all p, q ∈ M . In addition, d induces the given topology on M .
Definition 2.67. A Riemannian manifold (M, g) is geodesically complete if for every
p ∈ M , the expoential map expp is defined on the whole of Tp M .
That is, in the case of geodesically complete Riemannian manifold M , given
(p, v) ∈ TM the geodesic c(p,v) (t) is defined for all t ∈ R. In this section, we prove
the following fundamental theorem, linking the properties of (M, d) as a metric
space with geodesic completeness.
Theorem 2.68 (Hopf-Rinow). Let (M, g) be a Riemannian manifold, p ∈ M and d
the metric induced by g. Then the following statements are equivalent.
(i) The map expp is defined on the whole of Tp M .
(ii) Closed and bounded sets in M are compact.
(iii) The metric space (M, d) is complete.
(iv) The Riemannian manifold (M, g) is geodesically complete.
In addition, any of the above statements implies the following.
(v) For any q ∈ M there is a minimal geodesic joining p to q.
Proof. The main point of the proof lies in the implication (i)⇒(v). For this, we
first need to find a candidate for the tangent vector v ∈ Tp M giving the minimal
geodesic. To this end, let δ > 0 be such that B(p, δ) is a normal ball, that is, the
exponential map expp : B(0, δ) ⊆ Tp M → M is a diffeomorphism onto its image.
Let S := exp(S(0, δ)) be the image of the boundary of B(0, δ) and let x0 ∈ S be
a point such that d(x0 , q) = min{d(x, q) | x ∈ S} which exists by compactness of
S. Now, let v ∈ Tp M be the unique vector with kvk = 1 and expp (δv) = x0 , and
let γ : R → M , s 7→ expp (sv) be the corresponding geodesic which by assumption
is defined on the whole of R. We proceed to show that γ(r) = q for r := d(p, q):
Consider the set A := {s ∈ [0, r] | d(γ(s), q) = r − s}. Then A is non-empty since
0 ∈ A and closed by continuity. Now, let s0 ∈ A with s0 < r and δ ′ > 0 such that
(i) s0 + δ ′ ≤ r, and
(ii) B(γ(s0 ), δ ′ ) is a normal ball.
54 MARC BURGER STEPHAN TORNIER

In this setting, we show that s0 + δ ′ ∈ A. Since A is closed, this will imply that
r ∈ A which concludes this part of the proof. Consider the following figure.

x′0 b q
δ′ b
b

γ(s0 )
γ

p b

Let x′0 ∈ S ′ := expγ(s0 ) (B(0, δ ′ )) be such that d(x′0 , q) = min{d(x, q) | x ∈ S ′ }.


We claim that x′0 = γ(s0 + δ ′ ): Let v ′ ∈ Tγ(s0 ) M with kv ′ k = 1 be such that
expγ(s0 ) δ ′ v ′ = x′0 . The concatenation η of the geodesic γ([0, s0 ]) with γ ′ ([0, δ ′ ])
where γ ′ (t) = expγ(s0 ) (tv ′ ) for 0 ≤ t ≤ δ ′ is a piecewise C 1 -curve joining p to x′0
and has length l(η) = s0 + δ ′ . On the other hand, we have
(1) d(p, x′0 ) ≥ d(p, q) − d(q, x′0 )
and
(2) d(γ(s0 ), q) = δ ′ + min{d(x, q) | x ∈ S ′ } = δ ′ + d(x′0 , q).
Equation (2) implies
(3) r − s0 = δ ′ + d(x′0 , q).
Substituting (3) back into (1) yields
d(p, x′0 ) ≥ r − (r − s0 − δ ′ ) = s0 + δ ′ .
Since l(η) = s0 + δ ′ we conclude that d(p, x′0 ) = s0 + δ ′ and that η is a minimal
geodesic joining p to x′0 . Given that η ′ (0) = γ ′ (0) we conclude that η(t) = γ(t) for
all t ∈ [0, s0 + δ ′ ]. In particular, γ(s0 + δ ′ ) = x′0 which proves the claim.
We proceeed by showing how the claim implies s0 + δ ′ ∈ A: Indeed, (3) now
implies
r − s0 = δ ′ + d(x′0 , q) = δ ′ + d(γ(s0 + δ ′ ), q).
Equivalently, d(γ(s0 + δ ′ ), q) = r − (s0 + δ ′ ), i.e. s0 + δ ′ ∈ A.
For the implication (i)⇒(ii), let F be a closed and bounded set. Then
sup{d(p, x) | x ∈ F } < ∞.
We may hence choose T > 0 with d(p, x) ≤ T for all x ∈ F . By (v), we have
expp (B(0, T )) ⊇ F . That is, F is a closed subset of a compact set and hence
compact itself.
The implication (ii)⇒(iii) is a general topology statement.
To prove that (iii) implies (iv), assume that (M, g) is not geodesically complete.
Then there is q ∈ M and v ∈ Tq M with kvk = 1 such that γ(s) := expq (sv) is
defined on [0, s0 ) but not for s = s0 . Pick a sequence (sn )n∈N with sn < s0 for
all n ∈ N such that limn→∞ sn = s0 . Then d(γ(sn ), γ(sm )) ≤ |sn − sm | and hence
(γ(sn ))n∈N is a Cauchy sequence in M , which by assumption converges to a point,
say p0 = limn→∞ γ(sn ). Let (W, δ) be a totally normal neighbourhood of p0 . That
is, for all p ∈ W , the map
expp : B(0, δ) → expp (B(0, δ))
is a diffeomorphism and expp (B(0, δ)) ⊇ W . For large enough n, m, the images
points γ(sn ) and γ(sm ) lie in W and |sn − sm | < δ. Now there is a unique geodesic
RIEMANNIAN GEOMETRY 55

g joining γ(sn ) to γ(sm ) of length strictly less than δ. Clearly, γ coincides with g
whenever defined. Now expγ(sn ) : B(0, δ) → M is a diffeomorphism onto its image
which contains W . Hence g extends γ beyond s0 . 
Corollary 2.69. Every compact connected Riemannian manifold ist geodesically
complete.
Corollary 2.70. Let G be a compact connected Lie group. Then the Lie group
exponential expG : Lie(G) → G is surjective.
Corollary 2.71. Every closed submanifold of a complete connected Riemannian
manifold is complete.

3. Curvature
In this section we finally discuss various locally defined notions of curvature
and their global implications. The first was introduced by Gauss for surfaces in R3
and later on generalized by Riemann to sectional curvature: Given a Riemannian
manifold (M, g) and a point p ∈ M , look at two-dimensional subspaces E of Tp M
and apply Gauss’ notion of curvature to the surface pieces that arise from looking
at small geodesics with initial velocity in E. This yields a local notion of curvature
on M depending on the choice of a tangent plane.
There is a notion of curvature which subsumes all the above and many oth-
ers, called Riemannian curvature. Although not invented by Riemann himself it is
determined by all the sectional curvatures above.

3.1. Definition and Formal Properties. Given a Riemannian manifold (M, g)


with Levi-Civita connection ∇, consider the map
R(X, Y ) : Γ(TM ) → Γ(TM ), Z 7→ ∇Y ∇X Z − ∇X ∇Y Z + ∇[X,Y ] Z
One way to at it is to study the extent to which the map Γ(TM ) → End(Γ(TM ))
which maps X to ∇X fails to be a Lie algebra homomorphism.
Proposition 3.1. Retain the above notation. The map
Γ(TM ) × Γ(TM ) × Γ(TM ) → Γ(TM ), (X, Y, Z) 7→ R(X, Y )Z
is a tri-linear map of C ∞ (M )-modules.
Proof. As prepration for the proof, we compute for f, ϕ ∈ C ∞ (M ):
[f X, Y ](ϕ) = (f X)(Y (ϕ)) − Y (f X)(ϕ)
= f X(Y (ϕ)) − Y (f )Y (ϕ) − f Y X(ϕ) = (f [X, Y ] − Y (f )X)(ϕ).
Therefore, [f X, Y ] = f [X, Y ] − Y (f )X and [X, gY ] = g[X, Y ] + X(g)Y . Now, we
treat the C ∞ (M )-linearity of the given map in its three slots individually. For the
first one, we have
R(f X, Y )Z = ∇Y ∇f X Z − ∇f X ∇Y Z + ∇[f X,Y ] Z
= ∇Y (f ∇X Z) − f ∇X ∇Y Z + f ∇[X,Y ] Z − Y (f )∇X Z
= Y (f )∇X Z + f (∇Y ∇X Z − ∇X ∇Y Z + ∇[X,Y ] Z) − Y (f )∇X Z
= f R(X, Y )Z.
A similar computation works in the case of the second variable. Concerning the
Z-variable, first compute
∇Y ∇X (f Z) = ∇Y (X(f )Z + f ∇X Z)
= Y (X(f ))Z + X(f )∇Y Z + Y (f )∇X Z + f ∇Y ∇X Z
56 MARC BURGER STEPHAN TORNIER

and similarly ∇X ∇Y (f Z). The difference of the two equals


Y (X(f ))Z + f ∇Y ∇X Z − X(Y (f ))Z − f ∇X ∇Y Z
= (Y (X(f )) − X(Y (f )))Z + f (∇Y ∇X Z − ∇X ∇Y Z).
On the other hand, we have
∇[X,Y ] f Z = [X, Y ](f )Z + f ∇[X,Y ] Z = (X(Y (f )) − Y (X(f )))Z + f ∇[X,Y ] Z.
Overall, this completes the proof. 

Now, let (U, ϕ) be a chart of M at p ∈ M and let (∂i )m


i=1 be the associated
coordinate vector fields on U . Then
X
l
R(∂i , ∂j )∂k = Rijk ∂l
l
l
P
where Rijk : U → R are smooth functions on U . Decomposing X = Xi ∂i ,
P P i
Y = j j j and Z =
Y ∂ k Zk ∂k on U we get
  !
X X X
R(X, Y )Z = R  Xi ∂i , Yj ∂j  Zk ∂k
i j k
X
= Xi Yj Zk R(∂i , ∂j )∂k
i,j,k
X
l
= Xi Yj Zk Rijk ∂l .
i,j,k,l

As a corollary we record that R(X, Y )Z ∈ Γ(TM ) is a tensor on M in the


following sense.
Corollary 3.2. Retain the above notation. The value of R(X, Y )Z(p) only depends
on X(p), Y (p) and Z(p).
Definition 3.3. Let (M, g) be a Riemannian manifold with Levi-Civita connection
∇. The Riemannian curvature tensor of (M, g) is the collection of trilinear maps
Rp : Tp M × Tp M × Tp M → Tp M, (v, w, u) 7→ R(X, Y )Z(p) (p ∈ M )
where X, Y, Z are vector fields defined on open neighbourhoods of p ∈ M with
X(p) = v), Y (p) = w and Z(p) = u.
Equivalently, one can view the Riemannian curvature as a collection Rp (v, w) ∈
End(Tp M ) of endomorphisms, v, w ∈ Tp M , p ∈ M .
All curvature notions that we discuss are descendents of the Riemannian cur-
vature tensor but may have a more explicit geometric interpretation. In order to
work with these, we establish certain symmetry properties of R, coming from the
Lie algbera structure on Γ(TM ) as well as the relation between the Levi-Civita
connection and the metric.
Proposition 3.4 (Bianchi). Retain the above notation, let X, Y, Z ∈ Γ(TM ). Then
R(X, Y )Z + R(Y, Z)X + R(Z, X)Y = 0.
Proof. Expanding the definition, the proposition amounts to proving
0 = ∇Y ∇X Z − ∇X ∇Y Z + ∇[X,Y ] Z
+ ∇Z ∇Y X − ∇Y ∇Z X + ∇[Y,Z] X
+ ∇X ∇Z Y − ∇Z ∇X Y + ∇[Z,X] Y.
RIEMANNIAN GEOMETRY 57

Pairing appropriate expressions, this retranslates to


0 = ∇Y (∇X Z − ∇Z X) + ∇X (∇Z Y − ∇Y Z) + ∇Z (∇Y X − ∇X Y )
+ ∇[X,Y ] Z + ∇[Y,Z] X + ∇[Z,Y ] Y
Using relations like ∇X Z − ∇Z X = [X, Z] the above reduces to the Jacobi identity.


For the following, consider X, Y, Z, T ∈ Γ(TM ) on M and define for p ∈ M :


(X, Y, Z, T )p := hR(X, Y )Z(p), T (p)iTp M .
This way, we hope to deduce further properties of R taking into account the fact
∇ is not just a connection but behaves nicely with respect to the metric.
Proposition 3.5. Retain the above notation. Then the following hold.
(i) (X, Y, Z, T ) + (Y, Z, X, T ) + (Z, X, Y, T ) = 0.
(ii) (X, Y, Z, T ) = −(Y, X, Z, T ).
(iii) (X, Y, Z, T ) = −(X, Y, T, Z).
(iv) (X, Y, Z, T ) = (Z, T, X, Y ).
A way to remember Proposition 3.5 is to note the simple behaviour of the form
(−, −, −, −) under the action of the subgroup of S4 which preserves the block
decomposition {{1, 2}, {3, 4}} of {1, 2, 3, 4} depicted by (ii), (iii) and (iv). If one is
concerned with oher permutations one has to use Bianchi’s equality (i).

Proof. (Proposition 3.5). As remarked above, (i) is merely a reformulation of Bianchi’s


identity. The second assertion is an immediate consequnce of the definition. For (iii),
note that by doubling variables, it suffices to show that (X, Y, Z, Z) = 0. We are
therefore looking at h∇Y ∇X Z − ∇X ∇Y Z + ∇[X,Y ] Z − Zi. Using that ∇ is the
Levi-Civita connection, we have
h∇Y ∇X Z, Zi = Y h∇X Z, Zi − h∇X Z, ∇Y Zi
and
h∇X ∇Y Z, Zi = Xh∇Y Z, Zi − h∇Y Z, ∇X Zi.
Taking the difference of the above equalities we obtain Y h∇X Z, Zi − Xh∇Y Z, Zi.
Observing that
hZ, Zi = h∇X Z, Zi + hZ, ∇X Zi = 2h∇X Z, Zi
the above difference is
1 1 1
Y XhZ, Zi − XY hZ, Zi = − [X, Y ]hZ, Zi = h∇[X,Y ] Z, Zi.
2 2 2
Hence the assertion. Finally, for (iv) compute all instances of Bianchi’s identity (i)
arising from cyclic permutation of the variables and sum the result. Due to (ii) and
(iii), one obtains
2(X, Z, T, Y ) − 2(T, Y, X, Z) = 0
which is (iv). 
l
If so inclined, one can make the functions Rijk ∂l occuring in the equality
X
l
R(∂i , ∂j )∂k = Rijk ∂l
l

explicit in terms of the metric as in the case of the Christoffel symbols: Given that
[∂i , ∂j ] = 0 for all i, j ∈ {1, . . . , m} we have R(∂i , ∂j )∂k = ∇∂j ∇∂i ∂k − ∇∂i ∇∂j ∂k .
58 MARC BURGER STEPHAN TORNIER

P
Recall that ∇∂i ∂k = l Γlik ∂l . Plugging this into the above and comparing coeffi-
cients one gets the following, seldomly used, formula
X 
l
Rijk = (∂j Γlik − ∂i Γljk ) + Γsik Γljs − Γsjk Γlis
s
Going back to the formula of the Christoffel symbols in terms of the Riemannian
l
metric, one obtains an expression of Rijk in terms of the same.
3.1.1. Sectional Curvature. We now turn to sectional curvature, introduced by Rie-
mann, which is simpler than the Riemannian curvature tensor abovep but still deter-
mines it. Given vectors x and y in a Euclidean space, let |x∧y| = |x|2 |y|2 − hx, yi2
denote the area of the parallelogram spanned by x and y.
Proposition 3.6. Let (M, g) be a Riemannian manifold, p ∈ M and E ≤ Tp M
two-dimensional. Given a basis (x, y) of E, the quantity
(x, y, x, y)
K(x, y) :=
|x ∧ y|2
does not depend on the choice of basis of E.
Definition 3.7. Retain the above notation. The sectional curvature of M at p with
respect to E is given by K(E) := K(x, y) where (x, y) is any basis of E.
Proof. We show that K is invariant under the following transformations:
(i) (x, y) 7→ (x, y),
(ii) (x, y) 7→ (λx, y), λ ∈ R \{0},
(iii) (x, y) 7→ (x, y + µx), µ ∈ R.
This suffices since the associated matrices
     
1 λ 1 µ
, and
1 1 1
generate GL(2, R) for the given ranges of λ and µ. In particular, they realize any
possible change of basis. For the first transformation, note that
(y, x, y, x) = −(x, y, y, x) = (x, y, x, y)
2 2
and |x ∧ y| = |y ∧ x| . For (ii), observe
(λx, y, λx, y) = λ2 (x, y, x, y)
and |λx ∧ y|2 = λ2 |x ∧ y|2 . Finally, for the third assertion, compute
(x, y + µx, x, y + µx) = (x, y, x, y) + (x, y, x, µx) + (x, µx, x, y) + (x, µx, x, µx)
= (x, y, x, y) + µ(x, y, x, x) + µ(x, x, x, y) + µ2 (x, x, x, x)
= (x, y, x, y)
and |x ∧ (y + µx)|2 = |x ∧ y|2 . 
Remark 3.8. The Riemannian curvature tensor can be recovered from sectional
curvature in the following way: For x, y, z, t ∈ Tp M we have
∂2
(x + αz, y + βt, x + αz, y + βt) = 6(x, y, z, t)
∂α∂β
In order to give some examples, we make the following preliminary remarks:
Let (M, g) be a Riemannian manifold and f ∈ Iso(M, g). Using the uniqueness of
the Levi-Civita connection or the formula for the same developed in the proof one
verifies that for any X, Y, Z ∈ TM we have
∇f∗ X f∗ Y = f∗ (∇X Y )
RIEMANNIAN GEOMETRY 59

from which we deduce R(f∗ X, f∗ Y )f∗ Z = f∗ (R(X, Y )Z) by definition of R. Hence


for x ∈ M and E ≤ Tx M two-dimensional we have Kx (E) = Kf (x) (Dx f (E)). In
particular, the map
Kx : Gr2 (Tx M ) → R
is invariant under stabIso(M,g) (x).
Example 3.9. The above remarks facilitate the computation of sectional curvature
enormously in the following examples.
(i) The Riemannian curvature of (Rn , can) vanishes identically: Indeed, for
X, Y ∈ T Rn we have ∇X Y = LX Y if Y is considered as a map from Rn
to Rn . Consequently,
R(X, Y )Z = ∇Y ∇X Z − ∇X ∇Y Z + ∇[X,Y ] Z
= LY LX Z − LX LY Z + L[X,Y ] Z = 0
(ii) Now consider S n ⊆ Rn+1 with the induced Riemannian metric. We show
that K : Gr2 (S n ) → R is constant: Recall that O(n + 1) acts transtively by
isometries on S n since it does so on Rn+1 . Next,
  
A 0
stabO(n) (en+1 ) = A ∈ O(n)
0 1
acts on Ten+1 S n = {(v, 0)T ∈ Rn+1 | v ∈ Rn }. As remarked above, the map
Ken+1 : Gr2 (Tn+1 Rn ) → R is invariant under stabO(n+1) (en+1 ). Observing
that the action of O(n) on Gr2 (Rn ) is transitive we deduce that Ken+1 is
constant. Hence so is K by transitivity of O(n + 1).
(iii) Finally, consider hyperbolic space
Hn = {x ∈ Rn+1 | x21 + · · · + x2n − x2n+1 = −1, xn+1 > 0}.
We know that SO0 (n, 1) = {g ∈ O(n, 1) | det g = 1, g(Hn ) = Hn } acts
transitively on Hn by Riemannian isometries. As before, we look at the
base point en+1 ∈ Hn : We have Ten+1 Hn = {(v, 0)T ∈ Rn+1 | v ∈ Rn } and
the stabilizer
  
A 0
stabSO0 (n,1) (en+1 ) = A ∈ SO(n) .
0 1
The action of this stabilizer on Ten+1 Hn is equivalent to the SO(n)-action
on Rn which is again transitive on two-dimensional subspaces of Rn+1 .
Hence, as before, the sectional curvature of Hn is constant.
As a matter of fact, the three examples above are the only complete, simply con-
nected Riemannian manifolds with constant sectional curvature.
Example 3.10. As one might expect, things turn out particularly nice in the setting
of Lie groups: Let G be a Lie group which admits a bi-invariant Riemannian metric.
We have seen that for X, Y ∈ Γinv (TG) we have ∇X Y = 12 [X, Y ]. Hence we obtain
for X, Y, Z ∈ Γinv (TG):
R(X, Y )Z = ∇Y ∇X Z − ∇X ∇Y Z + ∇[X,Y ] Z
1 1 1
= [Y, [X, Z]] − [X, [Y, Z]] + [[X, Y ], Z]
4 4 2
1
= [[X, Y ], Z]
4
by Jacobi’s identity. Notice that since R is a tensor we do not have to worry about
only knowing its values on invariant vector fields: For x, y, z ∈ g = Te G we have
Re (x, y)z = 14 [[x, y], z].
60 MARC BURGER STEPHAN TORNIER

Now suppose that x, y ∈ g are orthonormal. Then


 
1 1
(x, y, x, y) = [[x, y], x], y = − h[x, [x, y]], y]ig .
4 g 4
Applying the invariance of h−, −i under the adjoint representation, i.e. for all
v, w, u ∈ g we have

d
0= hAd(exp(tv))w, Ad(exp(tv))ui
dt t=0
   
d d
= 0= Ad(exp(tv))w, v + w, Ad(exp(tv))u
dt t=0 dt t=0
= h[v, w], ui + hw, [v, u]i,
we obtain
1 1
h[x, y], [x, y]i = k[x, y]k2 .
(x, y, x, y) =
4 4
In particular, the sectional curvature of any two-dimensional subspace E ⊆ g is
greater or equal to zero. Equality holds if and only if E is an abelian subalgebra.
These considerations lead to root systems of Lie algebras and flat subspaces of
manifolds.
3.1.2. Ricci Curvature and Scalar Curvature. Although sectional curvature is easier
to deal with than the Riemannian curvature tensor, it is still fairly complicated given
that it determines the latter. Two weaker and simpler curvature notions are Ricci
curvature and scalar curvature which we define in this section.
Let (M, g) be a Riemannian manifold, p ∈ M and x ∈ Tp M a unit tangent
vector. Pick an orthonormal basis (z1 , . . . , zm ) of Tp M such that zm = x.
Definition 3.11. Retain the above notation. Then the Ricci curvature of M at p
with respect to x is
m−1
1 X
Ricp (x) := K(x, zi )
m − 1 i=1
and the scalar curvature of M at p is
m
1 X
K(p) := Ricp (zi )
m i=1

Notice that whereas sectional curvature depends on a point and a two-dimensional


subspace of that point’s tangent space, Ricci curvature only requires a point and a
tangent vector, and scalar curvature only depends on a point. We need to show that
the above definitions do not depend on the choice of orthonormal basis involving
the given vector x ∈ Tp M . To this end, consider the bilinear map
Q : Tp M × Tp M → R, (x, y) 7→ tr(z 7→ Rp (x, z)y).
Then Q is symmetric since
m
X m
X m
X
Q(x, y) = hRp (x, zi )y, zi i = (x, zi , y, zi ) = (y, zi , x, zi ) = Q(y, x).
i=1 i=1 i=1

Thus Q is a symmetric bilinear form and


m
X m−1
X m−1
X
Q(x, x) = (x, zi , x, zi ) = (x, zi , x, zi ) = K(x, zi ) = (m − 1)Ricp (x).
i=1 i=1 i=1
RIEMANNIAN GEOMETRY 61

For scalar curvature, let K ∈ End(Tp M ) be the symmetric endomorphism with


Q(x, y) = hK(x), yi. Then
m
X m
X m
X
tr(K) = hK(zi ), zi i = Q(zi , zi ) = (m − 1) Ricp (zi ) = m(m − 1)K(p).
i=1 i=1 i=1

3.2. The first and second variation formula. We now aim to deduce global
properties of a manifold from local curvature properties. To this end, we first need
to generalize the notions "vector field along a curve" and "covariant derivative" to
the setting of "parametrized submanifolds". Let N and M be a manifolds and let
h : N → M be a smooth map. In practice, N is a domain in R2 . Recall that
h∗ (TM ) := {(x, v) ∈ N × TM | v ∈ Th(x) M }
is the pullback bundle of TM under h. Its sections are vector fields along h. The
following two Lie algebra homomorphisms come with the above definition:
Γ(TN ) → Γ(h∗ TM ), X 7→ X, X(x) := Dx h(X(x)), and
Γ(TM ) → Γ(h∗ TM ), Y 7→ h∗ (Y ), h∗ (Y )(x) = Y (h(x)).
The following generalizes Theorem 2.37.
Proposition 3.12. Retain the above notation. There is a bilinear map
∇h : Γ(TN ) × Γ(h∗ (TM )) → Γ(h∗ (TM ))
such that for all X ∈ Γ(TN ), Y ∈ Γ(h∗ (TM )) and f ∈ C ∞ (N ):
(i) ∇hfX Y = f ∇hX Y , and
(ii) ∇hX (f Y ) = X(f )Y + f ∇hX Y .
Furthermore, we have for X ∈ Γ(TN ) and Y ∈ Γ(TM ):
(iii) ∇hX h∗ (Y ) = h∗ (∇X Y ).
Proof. We proceed as in the case of the covariant derivative along curves. If (U, ϕ)
is a chart of M and V := h−1 (U ) then for Y ∈ Γ(h∗ (TM )) and y ∈ V we have
m
X
Y (y) = Yi (y)h∗ (∂i )(y).
i=1
h
Assuming that ∇ with the asserted properties exists we thus have
m
X m
X 
∇hX Y (y) = X(Yi )(y)∂i (h(y)) + Yi (y) ∇hX ∂i (h(y))
i=1 i=1
Using 
(∇hX ∂i )(h(y)) = h∗ (∇X ∂i ) (y) = ∇Dy h(X(y)) ∂i (h(y))
we thus have
m
X m
X 
∇hX Y (y) = X(Yi )(y)∂i (h(y)) + Yi (y) ∇Dy h(X(y)) ∂i (h(y))
i=1 i=1

which shows uniqueness. As before, the above can be used to define ∇hX Y locally
from which one deduces the asserted properties. 
One can now pretend that ∇h produces a curavture notion and obtain the fol-
lowing.
Proposition 3.13. Let M and N be Riemannian manifolds and let h : N → M be
a smooth map. Further, let X, Y ∈ Γ(TN ) and U, V ∈ Γ(h∗ TM ). Then
(i) ∇hX Y − ∇hY X = [X, Y ].
(ii) XhU, V i = h∇hX U, V i + hU, ∇hX V i.
62 MARC BURGER STEPHAN TORNIER

(iii) ∇hY ∇hX U − ∇hX ∇hY U + ∇h[X,Y ] U = R(X, Y )U



where R(X, Y )U (y) := Rh(y) X(y), Y (y) (U (y)).

To prove e.g. part (i) of Proposition 3.13 one shows that ∇hX Y − ∇hY X − [X, Y ] is
C (N )-linear in X and Y and conludes by evaluating on coordinate vector fields.

The same works for part (ii) and (iii).


3.2.1. First Variation Formula. Let (M, g) be a Riemannian manifold. Recall that
Rb
for a piecewise C 1 -curce c : [a, b] → M , we have defined l(c) := a kc′ (t)k dt. For
many reasons, the energy of c, defined by
Z
1 b ′
E(c) := kc (t)k2 dt,
2 a
is a better object to work with due the smoothness and strict convexity of the
square of the absolute value, although it depends on the parametrization of c. Note
that by Cauchy-Schwarz we have
Z b Z b !1/2 Z !1/2
′ 2
b
′ 2
√ p
l(c) = 1 · kc (t)k dt ≤ 1 dt kc (t)k dt = b − a E(c).
a a a

In other words, l(c)2 ≤ (b − a)E(c) with equality if and only if kc′ (t)k is constant,
i.e. c being parametrized proportional to arc length.
We are interested in the minimal and critical points of thi energy functional.
Lemma 3.14. Let γ : [a, b] → M be a minimizing geodesic connecting p to q. Then
for any piecewise C 1 -curve c : [a, b] → M joining p to q we have E(γ) ≤ E(c) with
equality if and only if c is a minimizing geodesic.
Proof. By the above, (b − a)E(γ) = l(γ)2 ≤ l(c)2 ≤ (b − a)E(c). 
Formalizing the above question, consider the map
E : Ωp,q := {c : [a, b] → M | x is smooth, c(a) = p, c(b) = q} → R .
Then by the above the absolute minima of E arise through the minimal geodesics.
We now examime critical points which is made precise by the following.
Definition 3.15. Let M be a manifold and let c : [a, b] → M be a smooth curve. A
variation of c is a smooth map
h : [a, b] × (−ε, ε) → M
such that h(s, 0) = c(s). For t ∈ (−ε, ε), set ct : [a, b] → M , s 7→ h(s, t). The
variation h has fixed endpoints if h(a, t) = c(a) = p and h(b, t) = c(b) = q for all
t ∈ (−ε, ε).

ε c(b)
h
a b s

c(a)
RIEMANNIAN GEOMETRY 63

In the context of Definition 3.15, smoothness of h means that it is the restriction


of a smooth map defined on an open neighbourhood of [a, b] × (−ε, ε). We let ∂/∂t
and ∂/∂s denote the coordinate vector fields on R2 .
Lemma 3.16. Let M be a manifold and c : [a, b] → M be a smooth curve. Further,
let h : [a, b] × (−ε, ε) → M a variation of c. Then Y (s) := D(s,0) h(∂/∂t) is a vector
field along c. Conversely, given Y ∈ Γ(c∗ TM ) there is a variation h of c satisfying
D(s,0) h(∂/∂t).
Proof. The first assertion is immediate. Conversely, given Y ∈ Γ(c∗ TM ) set
h(s, t) := expc(s) tY (s)
which by a compactness argument is defined for t ∈ (−ε, ε) for some ε > 0. 

Theorem 3.17. Let (M, g) be a connected Riemannian manifold.


(i) Let c : [a, b] → M a smooth curve and let h be a variation of c. Define
Y ∈ Γ(c∗ TM ) by Y (s) := D(s,0) h(∂/∂t). Then
Z b
d ′ b
E(c ) = g(Y (s), c (s))| − g(Y (s), ∇c′ (s) c′ (s)) ds.
dt t=0
t a
a

(ii) The critical points of E : Ωp,q → R, i.e. curves c for which



d
E(ct ) = 0
dt t=0
for all variations of c with fixed endpoints, are exactly the geodesics con-
necting p to q.
Proof. Given the definition of the energy functional, we compute g(c′t (s), c′t (s)) and
integrate the result over s ∈ [a, b]. Let N be an open neighbourhood of [a, b]×(−ε, ε)
in R2 so that h : N → M is smooth. Then
   
∂ ∂
= D(s0 ,t0 ) h = c′t0 (s0 ).
∂s (s0 ,t0 ) ∂s
We therefore obtain using Proposition 3.12 and 3.13:
   
d ′ ′ d ∂ ∂ h ∂ ∂
g(c (s), ct (s)) = g , = 2g ∇ ∂ , .
dt t dt ∂s ∂s ∂t ∂s ∂s

Also, by Proposition 3.13 we have


 
h ∂ h ∂ ∂ ∂
∇∂ −∇∂ = , = 0.
∂t ∂s ∂s ∂t ∂t ∂s
Hence the above computation continues as
      
h ∂ ∂ d ∂ ∂ ∂ h ∂
= 2g ∇ ∂ , =2 g , −g ,∇ ∂
∂s ∂t ∂s ds ∂t ∂s ∂t ∂s ∂s

Finally, note that


 
∂ ∂ ∂ ∂
= D(s,0) h = Y (s), = c′ (s), and ∇h∂ = ∇c′ (s) c′ (s).
∂t t=0 ∂t ∂s t=0 ∂s ∂s

Overall, we therefore obtain


 
d ′ ′ d ′ ′
g(c (s), c (s)) = 2 g(Y (s), c (s)) − g(Y (s), ∇ c (s))
dt t t c (s)

t=0 ds
which is the assertion after integrating with respect to s.
64 MARC BURGER STEPHAN TORNIER

For the second assertion, note that by (i) we have for every variation of c with
fixed endpoints:
Z b
d
E(c ) = − g(Y (s), ∇c′ c′ (s)) ds
dt t=0
t
a
d

If c is a geodesic, then ∇c′ c′ = 0 and hence dt t=0
E(ct ) = 0. Conversely, assume
that the above integral vanishes for all variations of c with fixed endpoints. Let
f : [a, b] → R be a smooth function with f (a) = 0 = f (b) and f (s) > 0 for all
s ∈ (a, b). Set Y (s) := f (s)∇c′ c′ (s). Then
Z b
d
E(ct ) = − f (s)k∇c′ c′ (s)k2 ds = 0
dt t=0 a

Then ∇c′ c′ vanishes on (a, b) and hence also at a and b by continuity which shows
that c is a geodesic. 

3.2.2. Second Variation Formula. The second variation formula concerns the sec-
ond derivative of the energy functional. Retain the above notation and define in
addition: Y (s, t) := D(s,t) h(∂/∂t). Then Y is a smooth vector field along h.
Theorem 3.18. Retain the above notation. Let c : [a, b] → M be a geodesic. Then
  b Z b 2
1 d2 h ′ h ′ ′
E(c ) = g ∇ ∂ Y (s, 0), c (s) + ∇ ∂ Y (s) − (Y, c , Y, c )(s, 0) ds
2 dt2
t
t=0
∂t a ∂s
a

In the context of Theorem 3.18, recall that


(Y, c′ , Y, c′ )(s, 0) = g(Rc(s) (Y (s, 0), c′ (s))Y (s, 0), c′ (s)).

Proof. We determine an expression for



d2
g(c′t (s), c′t (s))
dt2 t=0
which we integrate with respect to s in order to obtain the second variation of the
energy functional. In view of Proposition 3.13 we have
   
∂ ∂ ∂ ∂
(s, t) = D(s,t) h = c′t (s) and (s, t) = D(s,t) h = Y (s, t).
∂s ∂s ∂t ∂t
To begin with, we compute
     
d d ∂ ∂ ∂ ∂ ∂ ∂
g(c′t (s), c′t (s)) = g , = 2g ∇h∂ , = 2g ∇h∂ , .
dt dt ∂s ∂s ∂t ∂s ∂s ∂s ∂t ∂s

Therefore we obtain
 
1 d2 d h ∂ ∂
= g ∇∂ ,
2 dt2 dt ∂s ∂t ∂s
   
h h ∂ ∂ h ∂ h ∂
=g ∇∂ ∇∂ , +g ∇∂ ,∇ ∂
∂t ∂s ∂t ∂s ∂s ∂s ∂t ∂s
| {z } | {z }
2
(1) h
=
∇ ∂ Y
∂s

The second term in the above sum is in its final term. We continue with the first
one using Proposition 3.13
     
h h ∂ ∂ ∂ ∂ ∂ ∂
(1) = g ∇ ∂ ∇ ∂ , +g R , ,
∂s ∂t ∂t ∂s ∂s ∂t ∂t ∂s
| {z } | {z }
(2) =−(Y,c′ ,Y,c′ )
RIEMANNIAN GEOMETRY 65

Again, the second term is final. We continue with (2):


   
d ∂ ∂ ∂ ∂
(2) = g ∇∂ , − g ∇h∂ , ∇h∂ .
ds ∂t ∂t ∂s ∂t ∂t ∂s ∂s

Evaluating at t = 0 and using that



∇h∂ = ∇h∂ c′ (s) = ∇c′ (s) c′ (s) = 0
∂s ∂s ∂s

since c is a geodesic we obtain overall:


2
1 d2 d  
2 E(ct ) = ∇h∂ Y (s, 0) − (Y, c′ , Y, c′ )(s, 0) + g ∇h∂ Y (s, 0), c′ (s)
2 dt t=0 ∂s ds ∂t

which implies the assertion. 


3.3. Curvature and Topology. Finally, we are in a position to discuss some in-
teresting applications. In fact, a large amount of recent research in Riemannian
Geometry has focused on the relation between the curvature of a Riemannian man-
ifold and its global topological properties. Ricci curvature in particular is involved
in many interesting statements. Recall that given a Riemannian manifold (M, g),
p ∈ M and x ∈ Tp M we have defined
m−1
X
1
Ricp (s) = hR(x, zi )x, zi i
(m − 1) i=1
where z1 , . . . , zm−1 , zm = x is an orthonormal basis of Tp M . As an example, con-
sider S m (r), the sphere of radius r. In this case, Ricp (x) = 1/r2 for all p ∈ S m (r)
and x ∈ Tp S m (r). Furthermore, the diamater of S m (r) is given by diamS m (r) = πr.
This example is the extremal case in the following Theorem.
Theorem 3.19 (Bonnet-Myers). Let M be a complete connected Riemannian man-
ifold with Ricp (x) ≥ 1/r2 > 0 for all (p, x) ∈ TM . Then M is compact and in fact
diam(M ) ≤ πr.
It is a theorem of Schenk that equality in the second inequality fo Theorem 3.19
implies that M is isometriec to a sphere of radius r. This relates to eigenvalues of
the Laplacian among other things.
Proof. (Theorem 3.19). By Theorem 2.68 it suffices to show that all minimizing
geodesics have length at most πr. Let p, q ∈ M and let γ : [0, 1] → M be
a minimizing geodesic with γ(0) = p and γ(1) = q. Let l = d(p, q). Then we
may extend γ ′ (0)/l to an orthonormal basis e1 , . . . , em−1 , em = γ ′ (0)/l of Tp M .
Let e1 (s), . . . , em−1 (s) be the parallel vector fields along γ with initial conditions
e1 , . . . , em−1 and define Yj (s) = sin(πs)ej (s). Then Yj (0) = 0 and Yj (1) = 0. Fi-
nally, let hj : [0, 1] × (−ε, ε) → M be a variation of c with fixed end points and

D(s,0) h ∂t = Yj (s). We now compute the second variation of the energy of the
map s 7→ hj (s, t), denoted by Ej′′ (0). Since
∇h∂ Yj (s, 0) = ∇h∂ (sin(πs)ej (s)) = π cos(πs)ej (s) + sin(πs)∇h∂ ej (s)
∂s ∂s
| ∂s{z }
=0

we have k∇h∂/∂s Yj (s, 0)k2 = π 2 cos(πs)2 . Furthermore,


(Yj , γ ′ , Yj , γ ′ )(s,0) = (sin πs)2 l2 (ej (s), em (s), ej (s), em (s)) (s, 0)
and hence
Z 1
1 ′′
E (0) = π 2 cos(πs)2 − l2 sin(πs)2 K(ej (s), em (s)) ds
2 j 0
66 MARC BURGER STEPHAN TORNIER

Partial integration of the π 2 cos(πs)2 term yields


Z 1
1 ′′ 
Ej (0) = sin(πs)2 π 2 − l2 K(ej (s), em (s)) ds
2 0
and therefore
1 X Z 1
m−1

Z 1

2 2 2
= sin(πs) π − l Ricc(s) (em (s)) ds ≤ sin(πs)2 π 2 − l2 /r2
2(m − 1) j=0 0 0
Pm−1
Now, if l πr then i=1 Ej′′ (0) 0. Hence Ej′′ (0) 0 for some j ∈ {1, . . . , m − 1}
which contradicts γ being minimizing. 
Corollary 3.20. Let M be a complete connected Riemannian manifold with Ricp (x) ≥
f of M is compact and
δ > 0 for all (p, x) ∈ T1 M . Then the universal covering M
π1 (M ) is finite.
f satisfies the same curvature
Proof. Since the covering map is a local isometry, M
f
bounds as M . Hence M is compact by Theorem 3.19 and π1 (M ) is finite. 
As usual, Lie groups with bi-invariant metrics constitute a particularly nice class
of examples.
Corollary 3.21. Let G be a connected Lie group which admits a bi-invariant metric
and let g := Lie(G). Assume that Z(g) = 0. Then both G and G e are compact and
π1 (G) is finite.
The assumption of Corollary 3.21 that Z(g) be trivial is equivalent to the center
of G being zero-dimensional.
Proof. (Corollary 3.21). Since G acts on itself by isometric left-translations, we
have Ricg (De Lg (v)) = Rice (v) for all v ∈ g with |v| = 1. Hence it suffices to bound
Rice (v) from below by a positive quantity. Recall that for orthonormal X, Y ∈ g
we have K(X, Y ) = 41 k[X, Y ]k2 . Now, let Y1 , . . . , Ym−1 , Ym = X be an orthonormal
basis of g. Then
m−1 m−1
1 X 1 X
Rice (X) = K(X, Yi ) = k[X, Y ]k2 ≥ 0
m − 1 i=1 m − 1 i=1
with equality if and only if [X, Yi ] = 0 for all i ∈ {1, . . . , m − 1} which in turn is
equivalent to X ∈ Z(g), i.e. X = 0 which contridicts kXk = 1. Therefore, the Ricci
curvature being a positive function on a compact sphere, it is bounded below by
some δ which implies the assertions by Theorem 3.19. 
As a remark in the context of the proof of Corollary 3.21 we state that
m
1 X1
Rice (X) = k[X, Yi ]k2
m − 1 i=1 4
Xm
1
= had(X)Yi , ad(X)Yi i
4(m − 1) i=1
X m
1
=− had(X)2 Yi , Yi i
4(m − 1) i=1
1
=− tr(ad(X)2 )
4(m − 1)
1
=− Kg (X, X)
4(m − 1)
RIEMANNIAN GEOMETRY 67

where Kg denotes the Killing form of g defined by


Kg : g × g → R, (X, Y ) 7→ tr(ad(X) ◦ ad(Y ))
Using this, one obtains more information about the diameter of such Lie groups in
terms of root systems.
Corollary 3.22. Let G be a connected compact Lie group with zero-dimensional
e is compact and π1 (G) is finite.
center. Then G
Proof. A compact Lie group always admits a bi-invariant metric. 
Another application of the second variation formula is the following.
Theorem 3.23 (Weinstein). Let M be a compact connected oriented Riemannian
manifold which has everywhere strictly positive sectional curvature. Further let
f ∈ Iso(M ) be orientation-preserving if dim M is even and orientation-reversing if
dim M is odd. Then f has a fixed point.
The example of the two-dimensional sphere and the antipodal map shows that
the assumptions on f are necessary.
Before turning to the proof of Theorem 3.23, consider the following linear version.
Lemma 3.24. Let A ∈ O(m − 1) and suppose that det A = (−1)m . Then A fixes a
non-trivial vector.
Proof. We show that 1 ∈ R is an eigenvalue of A: Let λ1 , . . . , λr , µ1 , µ1 , . . . , µk , µk
be the eigenvalues of A, listed with multiplicity, with λi ∈ R for all i ∈ {1, . . . , r} and
µi 6= µi for all i ∈ {1, . . . , n}. Since AQis orthgonal, λi ∈ {±1} for all i ∈ {1, . . . , r}
r
and all µi have unit length. Therefore i=1 λi = (−1)m . Also we have r+2k = m−1.
Now, assume that m is even. Then m − 1 − 2k is odd and hence so is r. Hence there
is i ∈ {1, . . . , r} such that λi = 1. Conversely, if m is odd, then −1 = λ1 · · · λr
implies r ≥ 1. Also, r is even since r = m − 1 − 2k. Consequently there is again
i ∈ {1, . . . , r} such that λi = 1. 
Proof. (Theorem 3.23). We argue by contradiction: Choose p ∈ M with
d(p, f (p)) = min{d(q, f (q)) | q ∈ M }
and assume that 0 < d(p, f (p)) =: l. In this case, let γ : [0, l] → M be a minimizing
geodesic connecting p to f (p). First, we claim that Dp f (γ ′ (0)) = γ ′ (l): To this end,
consider 0 < t′ < l and p′ = γ(t′ ). Then
l ≤ d(p′ , f (p′ )) ≤ l([p′ , f (p)] ∪ [f (p), f (p′ )]) = l.
Since the concatenation [p′ , f (p)] ∪ [f (p), f (p′ )] is a minimizing geodesic and hence
C 1 we conclude that γ ′ (l) = Dp f (γ ′ (0)).

f ◦γ b

Dp f (γ ′ (0)) f 2 (p)

γ b
γ ′ (l)
γ ′ (0) f (p)

Now consider A := Pγ,f (p),p ◦ Dp f ∈ GL(Tp M ). We have A(γ ′ (0)) = γ ′ (0) as well
as det A = (−1)m . Applying Lemma 3.24 to A ∈ O(γ ′ (0)⊥ ), there is e1 ∈ Tp M ,
orthogonal to γ ′ (0) which is fixed by A. Let e1 (s) ∈ Tγ(s) M be the vector field
parallel to γ with e1 (0) = e1 . Then Dp f (e1 (0)) = e1 (l).
68 MARC BURGER STEPHAN TORNIER

Now look at the variation h(s, t) := expγ(s) (te1 (s)). The above implies that
f (h(0, t)) = h(l, t), i.e. f (γt (0)) = γt (l). Hence
l(γt ) ≥ d(γt (0), γt (l)) = d(γt (0), f (γt (0))) ≥ l.
This implies that γ = γ0 is a local minimum of t 7→ l(γt ). Hence d2 /dt2 |t=0 l(γt ) ≥ 0.
On the other hand, the second variation formula yields
  l Z l 2
1 d2 ′ ′ ′
E(γ ) = g ∇ ∂ e1 (s, 0), γ (s) + ∇ ∂ e1 − (e1 , γ , e1 , γ )(s, 0) ds
2 dt2 t=0
t
∂t 0 ∂t
0
Z l
=− K(e1 (s), γ ′ (s)) ds < 0
0
which contradicts the above. 
We now discuss further applications of the second variation formula in both
positive and negative curavture.
Corollary 3.25 (Synge). Let M be a compact, connected Riemannian manifold
of dimension m with everywhere strictly positive sectional curvature. Then the
following hold.
(i) If m is even and M is orientable then M is simply connected.
(ii) If m i sodd then M is orientable.
As a preliminary remark towards the proof of Corollary 3.25 be note that every
connected Riemannian manifold M admits an orientable cover. Indeed, set
M := {(p, Op ) | p ∈ M, Op orientation on Tp M }.
It is obvious that one can equip M with the structure of a smooth manifold such
that p : M → M , (p, Op ) → p is a two-sheeted covering which is connected if
and only if M is not orientable. It is a Galois covering with non-trivial covering
transformation σ : (p, Op ) 7→ (p, O p ) where O p is the orientation opposite Op .
Proof. (Corollary 3.25). For the first assertion, let p : M f → M be the universal
cover of M and let e g be the Riemannian metric on M f turning p into a Riemannian
covering. In particular, if δ > 0 is a lower bound on the sectional curvature of M then
δ is also a lower bound on the sectional curvature of (M f, ge). Therefore, Theorem
f
3.19 implies that M is compact. Now, let π1 (M ) act by covering transformations on
f; they are orientation-preserving isometries since M is orientable. If π1 (M ) 6= {e}
M
then there is f ∈ π1 (M )\{Id} which by Theorem 3.23 has at least one fixed point.
This, however, contradicts the fact, that non-trivial covering transformations do
not have fixed points.
For part (ii), let m be odd and assume that M is not orientable. In this case,
consider the orientable Riemannian double cover M of M with Riemannian metric
g making the covering map Riemannian. Then the generator σ of π1 (M ) can be
viewed as an isometry of M and by Theorem 3.23 has a fixed point which yields
the same contradiction as before. 
3.3.1. Jacobi Fields. In this section we introduce another important, namely Jacobi
vector fields, which will allow us to give a description of the exponential map. To
motivate this, let’s rewrite the integral part of the second variation formula: Since
D E d D E D E
∇ ∂ Y, ∇ ∂ Y = Y, ∇ ∂ Y − Y, ∇2∂ ,
∂s ∂s ds ∂s ∂s

the integral term reads


Z b D E b Z b D E

k∇ ∂ Y k2 −(c′ , Y, c′ , Y )(s, 0) ds = Y, ∇ ∂ − Y, ∇2∂ Y + R(c′ , Y )c′ ds.
∂s ∂s a ∂s
a a
RIEMANNIAN GEOMETRY 69

Therefore,
Z bD E
d2 stuff depending on
E(ct ) = − Y, ∇2∂ Y + R(c′ , Y )c′ ds
dt2 t=0
endpoints only a ∂s

RbD E
Here, the symmetric bilinear form − a Y, ∇2∂ Y + R(c′ , Y )c′ ds can be viewed
∂s
as the second derivative of the energy functional at c and Jacobi vector fields are
vector fields in directions where “nothing happens”, reflecting degeneracy.
Definition 3.26. Let M be a Riemannian manifold and let c be a geodesic into M .
A Jacobi vector field is a vector field Y along c satisfying
∇ ∂ Y + R(c′ , Y )c′ = 0.
∂s

The following are the main statement we shall prove about Jacobi vector fields.
Theorem 3.27. Let M be a Riemannian manifold and let c be a geodesic into M .
Further, let u, v ∈ Tc(0) M . Then there is exactly one Jacobi vector field Y along c
with Y (0) = u and Y ′ (0) := ∇∂/∂s Y (0) = v.
Proposition 3.28. Let M be a Riemannian manifold, c : [a, b] → M a geodesic and
h : [a, b] × (−ε, ε) → M a variation of c such that for every t ∈ (−ε, ε), the curve
ct : [a, b] → M is a geodesic. Then Y (s) := D(s,0) h(∂/∂t) is a Jacobi vector field
along c. Coversely, every Jacobi vector field can be obtained in this way.
Proof. (Theorem 3.27). Fix an orthonormal frame X1 , . . . , Xm of parallel vector
fields along c using that parallel transport is isometric. Then any vector field Y
along c can be expressed as
Xm
Y (s) = yi (s)Xi (s)
i=1
for some smooth functions yi . Since the Xi are parallel along c we conclude
Xm
∇2∂ Y (s) = yi′′ (s)Xi (s).
∂s
i=1
Therefore,
D E
∇2∂ Y + R(c′ , Y )c′ = 0 ⇔ ∇2∂ Y + R(c′ , Y )c′ , Xi = 0 ∀1 ≤ i ≤ m.
∂s ∂s

Replacing Y by its expression in the Xi leads to the following system of ordinary


differential equations:
m
X
′′
yi (s) + yj (s)hc′ , Xj , c′ , Xi i(s) = 0.
j=1

Standard results now imply the theorem. 


We now turn to Proposition 3.28.
Proof. (Proposition 3.28). For the first part, assume that h : [a, b] × (−ε, ε) is a
variation of c such that all ct are geodesics and compute
 
2 ∂ ∂ ∂ ∂ ∂ ∂
∇∂ Y =∇∂ ∇∂ =∇∂ ∇∂ =∇∂ ∇∂ +R , .
∂s ∂s ∂s ∂t ∂s ∂t ∂s ∂t ∂s ∂s ∂t ∂s ∂s
Now note that
∂ ∂
(s, t) = D(s,t) h = c′t (s).
∂s ∂s
Because of this and the fact that all ct are geodesics we obtain evaluating at t = 0:
∇2∂ Y = R(Y, c′ )c′ = −R(c′ , Y )c′ .
∂s
70 MARC BURGER STEPHAN TORNIER

For the converse, let Y be a Jacobi vector field along c and let γ : (−ε, ε) → M be
a geodesic with γ(0) = c(0). Also, let X be a vector field along γ. Both are going to
be determined later. In any case, h(s, t) := expγ(t) (sX(t)) has the property that for
each t the map s 7→ h(s, t) is a geodesic. We have h(s, 0) = expc(0) (sX(0)). We are
going to choose X(t) = X1 (t) + tX2 (t) for some parallel vector fields X1 , X2 along
γ with X1 (0) = c′ (0). For the moment, compute the Jacobi vector field associated
to this geodesic variation.
∂ ∂
= D(s,0) h .
∂t ∂t
Unable to do this directly, we look at the initial conditions
∂ ∂
(0, 0) and ∇ ∂ (0, 0).
∂t ∂s ∂t
Regarding the first, we have

∂ ∂ d
(0, 0) = D(0,0) h = (t 7→ expγ(t) (0) = γ(t)) = γ ′ (0)
∂t ∂t dt t=0
Therefore, we set, γ ′ (0) = Y (0). Regarding the second, we have
∂ ∂
∇∂ (s = 0) = ∇ ∂ (s = 0).
∂s ∂t ∂t ∂s
Using that

∂ ∂
= D(0,t) h = X(t) = X1 (t) + tX2 (t)
∂s s=0 ∂s
we conclude

∇∂ (0, t) = X2 (t)
∂t ∂s
and hence choose X2 (0) = Y ′ (0). 

The following result finally links Jacobi vector fields to the derivative of the
expoential map.
Proposition 3.29. Retain the above notation. Let u, v ∈ Tp M and c(s) := exp(su).
Further, let Y be the Jacobi vector field along c with Y (0) = 0 and Y ′ (0) = 0.
Then Dsv (expp )(su) = Y (s).
Proof. Consider the variation of c given by h(s, t) := expp (s(v + tu)). Then we have
h(s, 0) = exp(sv) = c(s) and h(0, t) = p = c(0) = ct (0). We compute the Jacobi
vector field Z associated to h. In fact, in view of Theorem 3.27, we determine Z(0)
and Z ′ (0). Obviously, Z(0) = 0. Next,
∂ ∂
Z ′ (0) = ∇ ∂ (0, 0) = ∇ ∂ (0, 0)
∂s ∂t ∂t ∂s

where
∂ ∂
= D(s,t) h = Ds(v+tu) expp (v + tu)
∂s ∂s
At s = 0 we therefore have ∂/∂s(0, t) = v + tu. Hence Z ′ (0) = u. Therefore,
Z(s) = Y (s) but by definition


Z(s) = expp (s(v + tu)) = Dsu expp (su).
∂t t=0

RIEMANNIAN GEOMETRY 71

As an application of the theory of Jacobi fields, we present the following argu-


ments to determine the sectional curvature of spheres and hyperbolic space, by-
passing several pages of computations.
Consider the sphere S n . Let x ∈ S n and pick v ∈ S n with hx, vi = 0. Then
c(s) := (cos s)x + (sin s)v
is a geodesic. Now pick u ∈ S n with hu, xi = 0 = hu, vi and consider the variation
h(s, t) := cos s + sin s((cos t)v + (sin t)u).
Then h(s, 0) = c(s) and h(0, t) = x. Also,

Y (s) :=
= (sins)U (s)
∂t
for a vector field U along c with U (s) = u. Then
∇ ∂ Y = (cos s)U (s) and ∇2∂ Y = −(sin s)U (s)
∂s ∂s

because U is parallel along c given that parallel transport is an orientation-preserving


isometry. We thus have
0 = h∇2∂ Y + R(c′ , Y )c′ , Y i ⇔ 0 = − sin2 s + sin2 s(c′ u, c′ , u)
∂s

for all s with sin(s) 6= 0, implying (c′ , u, c′ , u) = 1 for such s and by continuity
(u, v, u, v) = 1.
The case of hyperbolic space is left as an exercise.
3.3.2. The Cartan-Hadamard Theorem. In this section, we see an example of how
negative curvature impacts on the topology of manifolds.
Theorem 3.30 (Cartan-Hadamard). Let (M, g) be a complete Riemannian manifold
with everywhere non-positive sectional curvature. Then expp : Tp M → M is a
covering map.
In the context of Theorem 3.30 note that if M is simply connected then expp is
a diffeomorphism. Its proof relies on the following lemma.
Lemma 3.31. Let (M, g) be a complete Riemannian manifold with everywhere non-
positive sectional curvature. Then expp : Tp M → M is a local diffeomorphism.
Proof. Here, we utilize our knowledge of the derivative of the exponential map as
well as the inverse function theorem. Using the notation of Proposition 3.29, let
Y (s) = Dsv expp (su) and assume u 6= 0. We know that Y (s) = 0 and Y ′ (0) = u.
Now consider the function f (s) := hY (s), Y (s)i. Then
D E
f ′ (s) = 2 ∇ ∂ Y (s), Y (s)
∂s

and D E D E
′′
(s) = 2 ∇ ∂ Y (s), ∇ ∂ Y (s) + 2 ∇2∂ Y (s), Y (s) .
∂s ∂s ∂s

where D E
∇2∂ Y, Y = −(c′ , Y, c′ , Y ) ≥ 0
∂s

which implies f ′′ (s) ≥ 0. Also, we have f (0) = 0 and f ′ (0) = 0. From the first,
we conclude that f ′ is increasing. In particular, since f ′ (0) = 0 we get f ′ (s) ≥ 0
for all s ∈ [0, ∞). Hence f is increasing and positive. Assume there is s0 > 0
with f (s0 ) = 0 which implies f (s) = 0 for all s ∈ [0, s0 ]. Then f ′′ (0) = 0 for all
s ∈ (0, s0 ), i.e.
k∇ ∂ Y (s)k = 0 ∀s ∈ (0, s0 ).
∂s

and by continuity
k∇ ∂ Y (0)k = 0.
∂s
72 MARC BURGER STEPHAN TORNIER

That is, u = 0, a contradiction. 


In order to pass from a local diffeomorphism to a covering map, we shall use the
following lemma whose proof is left as an exercise.
Lemma 3.32. Let (N1 , g1 ) and (N2 , g2 ) be connected Riemannian manifolds and
let p : N1 → N2 be a local isometry. Assume that (N1 , g1 ) is complete. Then p is
covering map.
Proof. (Theorem 3.30). Let ge be the pullback of g via expp , so that expp is a local
isometry. The metric ge is complete because straight lines in Tp M issuing from
0 ∈ Tp M are geodesics by the Lemma 2.62. 

4. What’s Beyond
First, we record the following important theorem, see [Wol11].
Theorem 4.1. Let M be a simply connected, complete Riemannian manifold of
constant sectional curvature k. Then M is, up to rescaling of the metric, isometric
to either Sm , Em or Hm .
What about classifying Riemannian manifold of constant sectional curvature in
general? To avoid pathologies like taking the complement of a closed set in one of
the above, one should require completeness at least. The positive curvature case
then amounts to determine, up to conjugacy, finite subgroups G of O(m + 1) such
that for every g ∈ G\{Id}, the number one is not an eigenvalue. This is fairly
complicated and was done accomplished by Vincent, a student of de Rham in the
1940’s. In the flat case, this is still an open question. For instance, Bieberbach
groups have only been classified for m ≤ 6, the case m = 3 being particularly
classical and used by chemist’s ever since. There is a mechanism due to Calabi by
which one can understand n-dimensional flat manifolds with β1 (M ) > 0 in terms
of lower dimensional ones N with β1 (N ) = 0.
In negative curvature, the story has a completely different flavour. For instance,
recall the classification of compact orientable manifolds.

Sphere Torus Surface with two holes

There is the sphere, the torus, and surfaces Sg (g ≥ 2) with g holes. The Gauss-
Bonnet theorem states that for a compact orientad surface S with sectional curva-
ture k : M → R we have
Z
1
kg dω = χ(S) = 2 − 2g
2π S
where χ(S) is the Euler characteristic of S. In particular, only the sphere carries
a positively curved metric, only the torus carries a flat metric and only the higher
genus surfaces potentially carry hyperbolic metric. In this case, the Gauss-Bonnet
Theorem implies that area(S) = 4π(g − 1).
In fact, there are many hyperbolic metrics on higher genus surfaces. To describe
these one can either turn to the study of discrete subgroups of the isomorphism
group of two-dimensional hyperbolic space or use the following explicit approach
due to Buser in 1978 involving right-angled hexagons in two-dimensional hyperbolic
RIEMANNIAN GEOMETRY 73

space. To describe these, we use the Poincaré disk model of the hyperbolic plane
which has the advantages that the Euclidean angles one sees are the same as the
actual (hyperbolic) angles since at every point in the disk, the hyperbolic metric is
a multiple of the Euclidean metric.
b

b
b

b
a
b

b
c

b
In a right-angled hexagons, bounded by geodesic rays, the side lengths a, b and c
are free parameters. Let H(a, b, c) denote the associated hexagon and H(a, b, c) the
one with opposite orientation. Glueing H(a, b, c) to H(a, b, c) yields a pair of pants
which closes up nicely due to the fact that the hexagons are right-angled.
2b

2a

2c
A surface of genus two now arises through gluing two pairs of pants along the
geodesic boundaries. Here, three more rotational parameters are introduced.

In total, there are six parameters. In the case of arbitrary higher genus g, there are
6g − 6 parameters.
In higher dimensions, the story yet takes a completely different flavour.
Theorem 4.2 (Mostow). Let M1 and M2 be compact Riemannian manifolds of
constant sectional curvature k = −1 and of dimension at least three. Then any
isomorphism of π1 (M1 ) and π1 (M2 ) is induced by an isometry of M1 and M2 .
In particular, such a manifold carries only one hyperbolic metric. Understanding
the isomorphism classes of such π1 (M ) inside SO(n, 1) remains a challenge though.
We also did not touch at all pinching theorems in the spirit of Berger-Klingenberg
mentioned in the introduction.
A revolutionary type of result was introduced by Gromov. The basic idea dates back
to Ulam who introduced the study of small perturbations of notions in algebra. For
instance, instead of studying the homomorphism equation ϕ(ab) − ϕ(a) − ϕ(b) = 0,
one may, in the presence of a metric, ask for the left-hand-side to be bounded and see
74 MARC BURGER STEPHAN TORNIER

what remains. In a Riemannian geometric setting, one may want to study manifolds
for which the absolute value of the sectional curvature is bounded. Or rather, since
the metric can always be scaled, ask for the product of the absolute value of the
sectional curvature and the diameter of the manifold to be bounded. If the manifold
has zero curvature and is compact then its fundamental group contains Zm and is
virtually abelian. If one only asks for a small bound, nilpotent fundamental groups
arise and it is an influential theorem of Gromov stating that this is always the case
in a certain sense.
Theorem 4.3. For every n ≥ 1 there is a constant εn > 0 such that if (M, g) is a
compact Riemannian n-manifold with
|k|diam(M )2 < εn
then there is a finite covering of M of the form Γ\N where N is a simply connected
nilpotent group and Γ is a lattice. In particular, π1 (M ) is virtually nilpotent.
For instance, the manifold
  
 1 x y 
N :=  1 z  x, y, z ∈ R
 
1
admits such an almost flat metric and the quotient N (Z)\N has nilpotent funda-
mental group N (Z).
RIEMANNIAN GEOMETRY 75

References
[BT15] M. Burger and S. Tornier, Differential geometry, 2015.
[Cha12] L. S. Charlap, Bieberbach groups and flat manifolds, Springer Science & Business Media,
2012.
[Kön13] K. Königsberger, Analysis 2, Springer, 2013.
[Lee10] J. Lee, Introduction to topological manifolds, vol. 940, Springer, 2010.
[Rie54] B. Riemann, Über die Hypothesen, welche der Geometrie zugrunde liegen, Werke 2
(1854), 272–287.
[Spi79] M. Spivak, Differential Geometry, vol. I-II, Publish or Perish, 1979.
[vdB06] E. van den Ban, Notes on quotients and group actions, 2006.
[Wol11] J. A. Wolf, Spaces of constant curvature, vol. 372, American Mathematical Society, 2011.

You might also like