0% found this document useful (0 votes)
33 views70 pages

Master DG

This document provides a transcript of lectures on differential geometry given by Marco Zambon at KU Leuven in 2019-20. The transcript covers topics including differentiable manifolds, tangent vectors, vector fields, fiber bundles, differential forms, integration on manifolds, and Lie groups. The notes were taken by Gilles Castel and reference two classical textbooks on the subject by Lee and Tu.

Uploaded by

Shinta
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
33 views70 pages

Master DG

This document provides a transcript of lectures on differential geometry given by Marco Zambon at KU Leuven in 2019-20. The transcript covers topics including differentiable manifolds, tangent vectors, vector fields, fiber bundles, differential forms, integration on manifolds, and Lie groups. The notes were taken by Gilles Castel and reference two classical textbooks on the subject by Lee and Tu.

Uploaded by

Shinta
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 70

Differential Geometry

Transcript1 of lectures by Marco Zambon at KU Leuven in 2019–20


Notes taken and typeset by Gilles Castel, proofread by the lecturer

1
Version: 19 October 2023.
This is a transcript of the Differential Geometry lectures by Marco Zam-
bon1 at the KU Leuven Master program in Mathematics, in 2019–20.
The notes were taken and typeset by Gilles Castel, whom the lecturer
gratefully thanks, and were proofread by the lecturer.

The lectures mainly drew from the following two classical references:

[Lee] Lee, John


Introduction to Smooth Manifolds
Graduate Texts in Mathematics, Volume 218, 2012 (Second edition).
https://sites.math.washington.edu/~lee/Books/ISM/

[Tu] Tu, Loring


An introduction to Manifolds
Universitext, 2010 (Second edition).
https://link.springer.com/book/10.1007/978-1-4419-7400-6

1
[email protected]

1
Contents

1 Differentiable manifolds 4
1.0 Topological spaces . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1 Differentiable manifolds . . . . . . . . . . . . . . . . . . . . . 6
1.2 Differentiable maps . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 Partition of unity . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2 Tangent vectors 13
2.1 Tangent vectors, tangent spaces . . . . . . . . . . . . . . . . . 13
2.2 The derivative of a map . . . . . . . . . . . . . . . . . . . . . 15
2.3 The regular level set theorem . . . . . . . . . . . . . . . . . . 17
2.4 Tangent vectors as derivations at a point . . . . . . . . . . . . 20

3 Vector fields 22
3.1 Vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2 Integral curves . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.3 The Lie bracket of vector fields . . . . . . . . . . . . . . . . . 24
3.4 Interpretation of the Lie bracket . . . . . . . . . . . . . . . . 27

4 Bundles 30
4.1 Fiber bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2 Vector bundles . . . . . . . . . . . . . . . . . . . . . . . . . . 32

5 Differential forms and integration 36


5.1 Forms on vector spaces . . . . . . . . . . . . . . . . . . . . . . 36
5.2 Differential forms on manifolds . . . . . . . . . . . . . . . . . 37
5.3 Orientation and volume forms . . . . . . . . . . . . . . . . . . 40
5.4 Integration on manifolds . . . . . . . . . . . . . . . . . . . . . 41

6 The exterior derivative and Stokes theorem 43


6.1 The exterior derivative in Rm . . . . . . . . . . . . . . . . . . 43
6.2 The exterior derivative on manifolds . . . . . . . . . . . . . . 44
6.3 Manifolds with boundary . . . . . . . . . . . . . . . . . . . . 45
6.4 Stokes’ theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 46

2
7 De Rham Cohomology 48
7.1 Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . 48
7.2 Homotopic maps and cohomology . . . . . . . . . . . . . . . . 50
7.3 The Mayer–Vietoris theorem . . . . . . . . . . . . . . . . . . . 52

8 Foliations 54
8.1 Immersed submanifolds . . . . . . . . . . . . . . . . . . . . . 54
8.2 Distributions and involutivity . . . . . . . . . . . . . . . . . . 55
8.3 The Frobenius theorem . . . . . . . . . . . . . . . . . . . . . . 55
8.4 Foliations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

9 Lie groups and Lie algebras 60


9.1 Lie groups, Lie subgroups . . . . . . . . . . . . . . . . . . . . 60
9.2 From Lie groups to Lie algebras . . . . . . . . . . . . . . . . . 63
9.3 The exponential map . . . . . . . . . . . . . . . . . . . . . . . 65
9.4 From Lie algebras to Lie groups . . . . . . . . . . . . . . . . . 67

CONTENTS 3
Chapter 1

Differentiable manifolds

1.0 Topological spaces

Definition 1.1 (Topological space). A topological space is (X, σ) where


X is a set and σ a family of subsets of X, called open sets, such that:

• ∅, X ∈ σ

• i∈I Ui ∈ σ whenever Ui ∈ σ for all i


S

• i<n Ui ∈ σ whenever Ui ∈ σ for all i (finite intersection)


T

Let (X, σ) be a topological space.

Definition 1.2 (Open neighbourhood). An open subset that contains p ∈


X is called a (open) neighbourhood of p.

Definition 1.3 (Subspace topology). If Y ⊂ X then (Y, σY ) is a topolog-


ical space, where
σY = {U ∩ Y | U ∈ σ}.
We call σY the subspace topology.

Example. Endowing R2 with the Euclidean topology, the subspace topology


on R × {0} ⊂ R2 is also the Euclidean topology. 

Definition 1.4 (Quotient topology). Let ∼ be an equivalence relation on


X. Consider π : X → X/ ∼. Then X/ ∼ is a topological space, where
the open sets are by definition the sets U such that π −1 (U ) is open in
X.

4
Definition 1.5 (Continuous functions). A function f : X1 → X2 is called
continuous iff ∀U ∈ σ2 : f −1 (U ) ∈ σ1 .

Definition 1.6 (Hausdorff). A topological space is called Hausdorff iff


∀x, y ∈ X, there exist neighbourhoods U of x, V of y such that U ∩V = ∅.

Example. Endow R2 \ {0} with the equivalence relation given by the thick
lines and the two half lines in the following figure. That is:
(
x = x0 if x 6= 0
(x, y) ∼ (x0 , y 0 ) ⇔
yy 0 > 0 if x = 0.

Then the quotient topology on (R2 \ {0})/ ∼ is not Hausdorff. 

X = R2 \ {0}

X

Figure 1.1: Example of a topology which is not Hausdorff

Definition 1.7 (Basis for a topology). A basis for the topology is S ⊂ σ


such that every open set of X is a union of elements of S.

Definition 1.8 (C2). A space (X, σ) is second countable if there exists a


countable basis.

Example. Rn is second countable. Indeed {B 1 (x) | x ∈ Qn , m ∈ N} is a


m
countable basis for the topology. Here Br (x) is the open ball with radius r
around x. 

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 5


1.1 Differentiable manifolds
Definition 1.9 (Topological manifold). A topological manifold M of di-
mension m is a second countable, Hausdorff topological space which is
locally homeomorphic to Rm .

Remark. ‘Locally homeomorphic to Rm ’ means that ∀p ∈ M, there exists a


neighborhood U of p and a homeomorphism φ : U → V ⊂ Rm . Recall that
open
homeomorphism means: bijective map that is continuous in both directions.

Definition 1.10 (Chart). The pair (U, φ) is called a chart.

Remark.
• Any subset of a Hausdorff space is Hausdorff
• Any subset of a C2 space is C2.
Recall that a map between open subsets of Rm is a diffeomorphism if it is
bijective, differentiable and the inverse is differentiable, where “differentiable”
means that all partial derivatives exist.

Definition 1.11 (Compatibility). Two charts (U1 , φ1 ) and (U2 , φ2 ) are


called smoothly compatible if

φ2 ◦ (φ1 )−1 |φ1 (U1 ∩U2 ) : φ1 (U1 ∩ U2 ) → φ2 (U1 ∩ U2 )

is a diffeomorphism.

Definition 1.12 (Smooth atlas). A smooth atlas for M is called a collec-


tion of charts {(Uα , φα )} such that α Uα = M and any two charts are
S
smoothly compatible.

Definition 1.13 (Maximal smooth atlas). A smooth atlas A is maximal


if: whenever B is a smooth atlas and A ⊂ B then B = A.

Definition 1.14 (Differentiable manifold ). A differentiable manifold (also


called a smooth manifold) is a topological manifold M together with a
maximal smooth atlas.

Remark. Given a smooth atlas A on a topological manifold M , there exists


a unique maximal smooth atlas containing it, namely

{(V, ψ) | (V, ψ) is smoothly compatible with all charts of A}.

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 6


U2
U1

φ1
φ2

V1
Rm V2
Rm
φ2 ◦ φ−1
1

Figure 1.2: Compatible charts

Example. Let U ⊂ Rn be open. Then U is a smooth manifold: an atlas is


{(U, Id)}. Take the maximal smooth atlas containing it. 
Example. Let S n := {x ∈ Rn+1 | kxk = 1}. The sphere S n with the sub-
space topology is Hausdorff and C2, simply because Rn+1 is. Two charts are
given by the stereographic projections from the Northpole N and Southpole
S:
(x1 , . . . , xn )
φN : S n \ {N } → Rn : (x1 , . . . , xn+1 ) 7→
1 − xn+1
(x 1 , . . . , xn )
φS : S n \ {S} → Rn : (x1 , . . . , xn+1 ) 7→ .
1 + xn+1

Now, φN and φS are homeomorphisms. Furthermore, kφN (p)k·kφS (p)k =


1, which allows us to calculate the inverse of φN . Hence
y
(φS ◦ φ−1 n n
N )|φN (S n \{N,S}) : R \ {0} → R \ {0} : y 7→ ,
kyk2

so φN and φS are smoothly compatible. Take the maximal smooth atlas


containing φN and φS . 
Remark. We could have started with other points P, Q ∈ S n instead of
N, S. The smooth atlases {φP , φQ } and {φN , φS } would be different, but
they define the same maximal smooth atlas.

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 7


1.2 Differentiable maps
Let M be a smooth manifold.

Definition 1.15 (Smooth function). A function f : M → R is differen-


tiable (or smooth) at p ∈ M iff ∃ a chart (U, φ) around p such that
f ◦ φ−1 is differentiable in φ(p).

M
f
U R
p

φ(U )

Rn

Figure 1.3: Smooth function from M to R

Remark. If f ◦ φ−1 is differentiable at φ(p) for a chart (U, φ), then f ◦ ψ −1 is


also differentiable at ψ(p), for any other chart (V, ψ) (in the maximal atlas
of M ).

Proof. We want to argue that f ◦ ψ −1 is smooth.

f ◦ ψ −1 = (f ◦ φ−1 ) ◦ (φ ◦ ψ −1 ) .
| {z } | {z }
C∞ C∞

Notation. We write C ∞ (M ) to denote all smooth functions M → R.

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 8


Definition 1.16 (Smooth map). f : M → N is differentiable at p ∈ M
iff

• it is continuous

• there exist charts (UM , φM ) around p and (UN , φN ) around f (p)


such that φN ◦ f ◦ φ−1
M is differentiable at φM (p)

Remark. The map φN ◦ f ◦ φ−1 M is defined on φM (UM ∩ f


−1 (U )). The
N
continuity of f ensures that this is an open neighborhood of φM (p) in Rm ,
hence it makes sense to talk about the differentiability of the above map at
φM (p).

M N
UM f UN
p f (p)

φM
φN

φN ◦ f ◦ φ−1
M
Rm Rn

Figure 1.4: Smooth function from M to N

Remark. A map f being a differentiable and a homeomorphism does not


imply that f is a diffeomorphism (which also include the differentiability
of the inverse). For instance, f : R → R, x 7→ x3 is not a diffeomorphism,

because the inverse x 7→ 3 x is not differentiable at zero.

1.3 Partition of unity

Definition 1.17 (Partition of unity). A partition of unity is a family


{eα }α∈A of smooth functions eα : M → [0, 1] such that

• For all p ∈ M, there exists a neighborhood U of p such that the


set {α ∈ A : eα |U 6≡ 0} is finite.


P
eα ≡ 1

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 9


Definition 1.18 (Subordinate). Let {Uα }α∈A be an open cover of M . A
partition of unity {eα }α∈A is subordinate to the cover ⇔ supp(eα ) ⊂ Uα ,
where
supp(eα ) = {p ∈ M : eα (p) 6= 0}.

Proposition 1.19. Let {Uα }α∈A be an open cover of M . Then there


exists a partition of unity subordinate to it.

This is useful in the following way: we can define functions, or vector


fields sα on open subsets of M which locally look like Rn . Then we create a
partition ofPunity subordinate to this cover, and paste the functions smoothly
together: eα · sα . This gives a smooth function (or vector field) on the
whole of M . The proof uses that the topology of M is second countable,
which is one of the reasons of requiring C2 in the definition of a manifold.

Proof. We present the idea of the proof, assuming M compact. For all
q ∈ M , choose α ∈ A such that q ∈ Uα . Let ψq be a functiona such that
ψq (q) = 1 supported in Uα . This is always possible, since the manifold
locally looks like Rn .
Since M is compact, the open cover {(ψq )P >0 q∈M has a finite
−1 (R )}
m
subcover. So ∃q1 , . . . , qm ∈ M such that ψ = i=1 ψqi > 0 on M . Now,
ψ
define φi = ψqi , which satisfies φi = 1 and supp(φi ) = supp(ψqi ) ⊂ Uα
P
for some α ∈ A. Now, one can rearrange these functions such that they
are indexed by A.
a
Such functions are sometime called bump functions.

1.4 Submanifolds
Let N be a smooth manifold of dimension n.

Definition 1.20 (Submanifold). A subset M ⊂ N is a submanifold of


dimension m if ∀p ∈ M there exists a chart (U, φ) such that

φ(U ∩ M ) = φ(U ) ∩ (Rm × {0}) .

Remark. A chart as above is called adapted to M .


Remark. The set M itself inherits the structure of a smooth manifold, with
smooth atlas given by

{(U ∩ M, φ|U ∩M ) : (U, φ) is a chart of N adapted to M }.

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 10


N

p M {0} × Rn−m
φ

Rn φ(U )

Rm × {0}

Figure 1.5: Definition of a submanifold

Example. The submanifolds of the same dimension as N are exactly the


open sets of N . The (connected) submanifolds with dimension 0 are exactly
the points. 
Example. The union of the axes {0} × R ∪ R × {0} is not a submanifold of
R2 . The problematic point is the origin (the “cross”). 
Example. The unit circle S 1 ⊂ R2 is a submanifold of R2 . The idea is that,
for each point of S 1 , we need to find a chart that ‘flattens’ a part of the circle
to a line. Define

φa : R2 − (R≥0 × {0}) −→ (0, 2π) × R+ φb : R2 − (R≤0 × {0}) −→ (−π, π) × R+


   
r cos θ r cos θ
7−→ (θ, r) 7−→ (θ, r).
r sin θ r sin θ

Note that 0 is not included in either of the charts, but that is not a problem.
We have that φa (S 1 − {(1, 0)}) = (0, 2π) × {1} and φb (S 1 − {(−1, 0)}) =
(−π, π)×{1}. This is already flat, but per definition of submanifold, we need
to move this to zero. So instead of {φa , φb }, use {φa − (0, 1), φb − (0, 1)}.

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 11


R2
1
φa

0 2π

1 R2

φb

−π 0 π

Figure 1.6: A circle is a submanifold of the plane.


Example. Consider the torus, S 1 × S 1 , where S 1 = {z ∈ C : |z| = 1}. Fix
α ∈ R. Now consider

M = {(e2πit , e2πiαt ) : t ∈ R},

a subset of the torus. This is a submanifold of the torus iff α ∈ Q (this


happens exactly when the spiral closes up). When α ∈
/ Q, M is dense in the
torus. 

Figure 1.7: Two examples of submanifolds of the torus. On the left, α = 7


2
and on the right α = 4.

CHAPTER 1. DIFFERENTIABLE MANIFOLDS 12


Chapter 2

Tangent vectors

2.1 Tangent vectors, tangent spaces


Note. This and the next section do not follow neither Lee’s nor Tu’s book.
Remark. Given a submanifold M of Rn , one can define tangent vectors to
M at some point p ∈ M as the collection of dt
d
0
γ(t) where γ : (−ε, ε) → Rn
is smooth and Im γ ⊂ M , γ(0) = p. This uses the ambient space Rn , hence
to define tangent vectors to manifolds we cannot proceed in the same way.
Let M be a smooth manifold of dimension m.

Definition 2.1 (Tangent vector). A tangent vector of M at p is an equiv-


alence class [γ] of smooth curves γ : (−ε, ε) → M with γ(0) = p, where

γ1 ∼ γ2 ⇔ ∃ a chart (U, φ) containing p s.t. (φ ◦ γ1 )0 (0) = (φ ◦ γ2 )0 (0).

Lemma 2.2. If (φ ◦ γ1 )0 (0) = (φ ◦ γ2 )0 (0) for a chart φ, then the same


holds for all charts ψ.

Proof. The linear map Dφ(p) (ψ ◦ φ−1 ) : Rm → Rm sends (φ ◦ γi )0 (0) to


(ψ ◦ γi )0 (0) because of the chain rule, for i = 1, 2. Now, as (φ ◦ γ1 )0 (0) =
(φ ◦ γ2 )0 (0), the vectors obtained applying Dφ(p) (ψ ◦ φ−1 ) must also be
the same.

13
V
U
p

φ
ψ

φ(U ) ψ ◦ φ−1 ψ(V )


Rm Rm

Figure 2.1: Charts in the proof of Lemma 2.2

Definition 2.3 (Tangent space). ∀p ∈ M , the set of all tangent vectors


at p is denoted Tp M , and called the tangent space at p.

Proposition 2.4. Tp M is a vector space, of dimension equal to dim(M ).

Proof. Let (U, φ) be a chart. We obtain a map

Tp M → Rm , [γ] 7→ (φ ◦ γ)0 (0).

It’s well-defined and injective by the definition of tangent vector. It is


also surjective. Indeed, given v ∈ Rm , take the straight line t 7→ φ(p)+tv
in Rm and then apply φ−1 , to obtain the following curve on M :

γ(t) = φ−1 (φ(p) + tv).

Now [γ] maps to v.


Now that we’ve proved that this is a bijection, we immediately get a
vector space structure on Tp M by “transporting” the one on Rm . Sup-
pose we had started with another chart (V, ψ), then we would have
obtained the same vector space structure on Tp M , because

Dφ(p) (ψ ◦ φ−1 ) : Rm → Rm

is a linear isomorphism making this diagram commute:

CHAPTER 2. TANGENT VECTORS 14


[γ] ∈ Tp M

Dφ(p) (ψ◦φ−1 )
(φ ◦ γ)0 (0) ∈ Rm (ψ ◦ γ)0 (0) ∈ Rm

Remark. Let W ⊂ Rm open, p ∈ W . There is a canonical linear isomorphism


Tp W → Rm , [γ] → γ 0 (0). To see this, take φ = Id in the proof of the previous
lemma.
When you choose a chart around p ∈ M , you get a basis of Tp M .

Definition 2.5 (Basis of Tp M induced by a chart). Let p ∈ M , (φ, U ) a


chart with p ∈ U whose components we denote by x1 , . . . , xm (hence
xi : U → R). By means of the isomorphism of vector spaces

Tp M → Rm : [γ] 7→ (φ ◦ γ)0 (0),

the standard basis of Rm induces a basis of Tp M , which we denote by


∂x1 p , . . . , ∂xm p .
∂ ∂

2.2 The derivative of a map

Definition 2.6 (Derivative of a smooth map). If f : M → N is differen-


tiable, then its derivative at p ∈ M is

(f∗ )p : Tp M → Tf (p) N : [γ] 7→ [f ◦ γ].

Proposition 2.7. (f∗ )p is well defined and linear.

Proof. Let φM be a chart for M near p. Let φN be a chart for N near


f (p). Consider the commutative diagram

(f∗ )p
[γ] ∈ Tp M Tf (p) N 3 [f ◦ γ]

DφM (p) (φN ◦f ◦φ−1


M )
(φM ◦ γ)0 (0) ∈ Rm Rn 3 (φN ◦ (f ◦ γ))0 (0)

As the two vertical maps are linear isomorphisms (as we saw in the

CHAPTER 2. TANGENT VECTORS 15


proof of Proposition 2.4) and the horizontal map is linear (being the
derivative of a smooth map between open subsets of Euclidean space),
the composition (f∗ )p is also linear.

f g
Proposition 2.8 (Chain rule). If M → N → L are smooth maps, then
∀p ∈ M we have (g ◦ f )∗ p = (g∗ )f (p) ◦ (f∗ )p : Tp M → T(g◦f )(p) L.

Remark. Let p ∈ M and (U, φ) a chart around p. One can show that φ : U →
φ(U ) is a diffeomorphism of manifolds. (Notice that since U is an open set of
M and φ(U ) is an open set of Rm , both carry manifold structures). We have
a commutative diagram of isomorphisms, where the bottom isomorphism is
the one of the last remark.
[γ] ∈ Tp M
(φ∗ )p

(φ ◦ γ)0 (0) ∈ Rm ∼
= Tφ(p) (φ(U ))

In other words, the map Tp M → Rm from the proof of Proposition 2.4, under
the identification given in the previous remark, is (φ∗ )p . In particular

∈ Tp M
∂xi p

and the i-th standard basis vector of Tφ(p) φ(U ) ∼


= Rm correspond under
(φ∗ )p .


is basis of Tp U
∂xi p

Basis of Tφ(p) φ(U )


Rm

Figure 2.2: The i-th standard basis vector and ∂


∂xi p
correspond under (φ∗ )p .

CHAPTER 2. TANGENT VECTORS 16


2.3 The regular level set theorem
Remark. Let f : U → V be a diffeomorphism between open subsets of Rn .
Then ∀q ∈ U , Dq f : Rn → Rn is an isomorphism, because its inverse is
Df (q) f −1 .
Conversely we have:

Lemma 2.9 (Inverse function theorem in Rn ). Let U ⊂ Rn open, f : U →


Rn smooth s.t. Dq f : Rn → Rn is an isomorphism for some q ∈ U . Then
there exists a neighbourhood V ⊂ U of q such that f |V : V → f (V ) is
a diffeomorphism.

Corollary 2.10 (Inverse function theorem for manifolds). Let f : M → N


be a smooth map p ∈ M , such that f∗ (p) : Tp M → Tf (p) N is an
isomorphism. Then there exists a neighbourhood W of p s.t. the map
f |W : W → f (W ) is a diffeomorphism.

Idea of the proof: this is a local statement, and by means of charts, locally
every manifolds can be identified with an open subset of Rn .
Example. Consider f : R → S 1 ⊂ C, t 7→ e2πit . This is not a diffeomor-
phism. But f∗ (t) is an isomorphism for all t ∈ R. So locally f restricts to a
diffeomorphism onto it image. 
Given k ≥ n, we denote

π : Rn × Rk−n → Rn , (v, w) 7→ v.

Lemma 2.11 (Submersion theorem in Rn ). Let U be a neighbourhood of


the origin 0 in Rn × Rk−n and f : U → Rn smooth such that

(D0 f )|Rn ×{0} : Rn × {0} → Rn

is an isomorphism. Then there exists a diffeomorphism τ between neigh-


bourhoods in Rn × Rk−n such that f ◦ τ −1 = π.

This theorem states that precomposing f with a diffeomorphism, we can


arrange that it becomes the projection on the first components.

Proof. Consider τ := (f1 , . . . fn , xn+1 , . . . , xk ) : U → Rn × Rk−m , which


we can write concisely as (f, IdRk−n ). Consider its derivative at the

CHAPTER 2. TANGENT VECTORS 17


Rn × Rk−n

f
Rn π

Figure 2.3: The submersion theorem. The dotted lines denote the preimages
of points of Rn under f and π.

origin  
(D0 f )|Rn ×{0} ∗
D0 τ = .
0 IdRk−n
This matrix is invertible, because it is an upper block matrix, (D0 f )|Rn ×{0}
is invertible, and IdRk−n is invertible. This means that τ is a diffeomor-
phism near 0, by the inverse function theorem. We have π ◦ τ = f .

Definition 2.12 (Regular value). Given a smooth map f : M → N , a


point c ∈ N is a regular value iff ∀p ∈ f −1 (c), the derivative (f∗ )p :
Tp M → Tc N is surjective.

Theorem 2.13. Let f : M k → N n be a smooth, and let c ∈ N be a


regular value s.t. f −1 (c) 6= ∅. Then

• f −1 (c) is a submanifold of M with dimension k − n

• ∀p ∈ f −1 (c) : Tp (f −1 (c)) = Ker(f∗ (p))

The first item states that the codimension is preserved when taking in-
verse images.
Example. Consider the “height function” f : S 2 → R : (x, y, z) 7→ z. Note
that −1 and 1 are not regular values. For all c ∈ (−1, 1), the preimage
f −1 (c) is a submanifold of S 2 , diffeomorphic to a circle. 

CHAPTER 2. TANGENT VECTORS 18


Figure 2.4: The “height function” f : S 2 → R. The derivative of f vanishes
at the Northpole and Southpole.

Proof. Fix p ∈ f −1 (c). Take charts (U, φM ) near p, and (V, φN ) near
f (p) = c, chosen so that φN (c) = 0. We know that Dφ(p) (φN ◦ f ◦ φ−1
M):
R → R is surjective (because c is a regular value). We can assume
k n

that
Dφ(p) (φN ◦ f ◦ φ−1
M) n
R ×{0}

is an isomorphism. (When we restrict a surjective linear map to a sub-


space transverse to the kernel, it becomes an isomorphism. If Rn × {0}
is not transverse to the kernel of Dφ(p) (φN ◦ f ◦ φ−1
M ), we can change the
chart φN by composing it with e.g. a rotation in Rn .)
By the last lemma (Submersion theorem in Rn ), there exists a dif-
feomorphism τ such that (φN ◦ f ◦ φ−1M )◦τ
−1 = π. This can be rewritten

as
φN ◦ f ◦ (τ ◦ φM )−1 = π.
The situation is summarized by the following diagram:

f
U V 3c
τ ◦φM φN .
π
open ⊂ Rk open ⊂ Rn 3 0

Hence τ ◦ φM is a chart of M adapted to f −1 (c).

For the second part, let p ∈ f −1 (c), and take a path γ : (−ε, ε) →
with φ(0) = p. Then
f −1 (c)

(f∗ )p [γ] = [f ◦ γ] = 0 ∈ Tc N

CHAPTER 2. TANGENT VECTORS 19


as f ◦γ ≡ c. This shows that Tp (f −1 (c)) ⊂ Ker(f∗ )p . Both vector spaces
have the same dimension, which gives the equality.

π
0

(τ ◦ φM )(f −1 (c) ∩ U )

Figure 2.5: The chart τ ◦ φM straightens out f −1 (c) (intersected with the
domain of the chart).

2.4 Tangent vectors as derivations at a point


Fix M = Rn . Recall that a tangent vector at p is an element of Tp Rn ∼
= Rn .

Definition 2.14 (Derivation at a point). A derivation at a point p ∈ Rn


is a linear map D : C ∞ (Rn ) → R satisfying the Leibniz rule:

D(f g) = D(f )g(p) + f (p)D(g).

Example. ∀v ∈ Rn , the directional derivative


C ∞ (Rn ) −→ R
X ∂f
f 7−→ (dp f )(v) = vi (p)
∂xi
i
is a derivation at p. This follows from the fact that partial derivatives obey
the Leibniz rule. 
Remark. For a constant function c, we have Dc = 0. Because D is linear, it
is enough to show this for c = 1. We have D1 = (D1)1 + 1(D1) = 2(D1), so
D1 = 0.

Proposition 2.15. The map

φ : Tp Rn −→ Derivations at p
 X ∂f 
v 7−→ f 7→ vi (p)
∂xi
i

is an isomorphism of vector spaces.

CHAPTER 2. TANGENT VECTORS 20


Proof.

• This formula really defines a derivation at p, as we’ve showed in


the previous example.

• It is clearly linear.

• To P
show that it’s injective, we check that the kernel is 0. If for all
∂f
f , i vi ∂xi
(p) = 0, then it is particular true for the functions xj ,
P ∂x
so vj = 0 for all j. In formulae: vj = i vi ∂xji (p) = 0.

• Surjectivity. Let D be a derivation at p. For all f ∈ C ∞ (Rn ), we


have X
f (x) = f (p) + (xi − pi )gi (x),
i
∂f
where gi (x) is a smooth function with gi (p) = ∂xi (p). (This is a
version of Taylor’s theorem.) Then
X
Df = 0 + D(xi )gi (p) + 0
i
X ∂f
= D(xi ) (p).
∂xi
i

So v = (D(x1 ), . . . , D(xn )) maps to D.

CHAPTER 2. TANGENT VECTORS 21


Chapter 3

Vector fields

3.1 Vector fields


DefinitionS3.1 (Vector field). A vector field on a manifold M is a map
X : M → p∈M Tp M , such that

• X(p) ∈ Tp M

• X satisfies the following smoothness condition: for any chart (U, φ),
writing X(p) = i ai (p) ∂x ∂
, all the coefficients ai : U → R are
P
ip
smooth.

Notation. We denote the set of all vector fields on M with X(M ).


Example. Let (U, φ) be a chart on M . Then ∂x ∂
i
is a vector field on U , for
all i = 1, . . . , dim(M ).
If U ⊂ Rn is open, using the chart Id, ∂x∂ 1 is just the vector field with
unit vectors pointing in the x1 direction. 

3.2 Integral curves

Definition 3.2 (Integral curve). Let X ∈ X(M ). A smooth curve γ :


(a, b) → M is an integral curve of X iff

γ̇(t) = X|γ(t)

for all t ∈ (a, b).

Remark. Here γ̇(t) is defined as (γ∗ )t (1), where (γ∗ )t : Tt (a, b) → Tγ(t) M ,
and Tt (a, b) ∼
= R, so using 1 as an input is valid. Another way to look at
it, in terms of tangent vectors as equivalence classes of curves: γ̇(t) equals
[s 7→ γ(s + t)].

22
Lemma 3.3. For all p ∈ M , there exists an ε > 0 and a unique integral
curve γp of X defined on (−ε, ε), starting at the point p.

Proposition 3.4. Let X be a vector field, p ∈ M . Then there exists a


neighborhood U of p, an ε > 0 and a unique smooth map

F : U × (−ε, ε) → M

s.t. for all q ∈ U , the curve γq defined by γq (t) = F (q, t) is an integral


curve of X with γq (0) = q.

Proof. Fix a chart (φ, V ) near p. In these coordinates, we have X =


i fi (x) ∂xi for some smooth functions fi . We need to show that there
P ∂

exists a neighboorhood W ⊂ φ(V ) of φ(p), ∃ε > 0 and ∃! y : W ×


(−ε, ε) → φ(V ) such that
(

∂t y(x, t) = f (y(x, t))
y(x, 0) = x

for all x. This holds by the fundamental theorem of ODE’s. (It says
that for each initial value x, there is a unique solution defined on a small
interval (−ε, ε), and the solution varies smoothly with x.)

Remark. The map F in the above proposition is called flow.


Example. On R2 take X = ∂x1 . Then the flow is Ft (x1 , x2 ) = (x1 + t, x2 ). 

Example. On R2 let X = x ∂y∂ ∂


− y ∂x . Then the images of integral curves
are circles. The flow is given by Ft (x, y) = Rt (x, y), where Rt denotes the
rotation by the angle t. 
Example. Sometimes, we cannot continue the curve. For example, look at
R2 \ {(0, 0)} and X = ∂x∂ 1 . Then the maximal integral curve starting at
(−2, 0) is defined only on (−∞, 2). 

CHAPTER 3. VECTOR FIELDS 23


Definition 3.5 (1-parameter group of diffeomorphisms). A 1-parameter
group of diffeomorphisms on M is a smooth map F : M × R → M such
that, using the notation Ft (p) = F (p, t) (think of it as fixing t and
varying the point), one has

• Fs ◦ Ft = Fs+t ∀s, t

• F0 = Id

(It then follows that Ft is a diffeomorphism for all t.)

Proposition 3.6. Assume, for the sake of simplicity, that the flow of X
is defined on M ×R. Then F is a 1-parameter group of diffeomorphisms.

Proof. Let p ∈ M . By the uniqueness in Lemma 3.3, the composition of


the integral curves (γp )|[0,t] and (γγp (t) )|[0,s] (i.e. running along the first
curve and then along the second) equals the integral curve (γp )|[0,s+t] .
Hence

Fs (Ft (p)) = Fs (γp (t)) = γγp (t) (s) = γp (s + t) = Fs+t (p).

Remark. There is a bijection between

• vector fields on M whose flow is defined on M ×R (the biggest possible


domain), and

• 1-parameter groups of diffeomorphisms on M .

The bijection reads:

X flow F as above
X given by X(q) = d
dt 0 Ft (q) F : M × R → M.

3.3 The Lie bracket of vector fields


Recall that C ∞ (M ) = {smooth functions from M to R}.
Remark. C ∞ (M ) is an algebra: it is a vectorspace, but we can also multiply
two functions.

CHAPTER 3. VECTOR FIELDS 24


Definition 3.7 (Derivation of C ∞ (M )). A derivation of C ∞ (M ) is a lin-
ear map D : C ∞ (M ) → C ∞ (M ) such that

D(f g) = D(f )g + f D(g).

Remark. If D1 , D2 are derivations of C ∞ (M ), then the commutator D1 ◦


D2 − D2 ◦ D1 is also a derivation. But D1 ◦ D2 on its own is not a derivation!

Proposition 3.8. There is a linear map

Φ : X(M ) −→ Derivations of C ∞ (M )
X 7−→ (f 7→ f∗ (X)).

Here f∗ (X) is the function on M given by (f∗ (X))p = (f∗ )p (X|p ) ∈


Tf (p) R ∼
= R. We denote f∗ (X) =: X(f ).

Proof. We’ll show that ∀X ∈ X(M ), Φ(X) is a derivation. Let f, g ∈


C ∞ (M ), p ∈ M . Let γ be a curve in M such that γ(0) = p, [γ] = Xp .
Then

(f g)∗p (Xp ) = [(f g) ◦ γ] ∈ T(f g)(p) R


= ((f g) ◦ γ)0 (0) ∈ R
0
= (f ◦ γ) · (g ◦ γ) (0)
= (f ◦ γ)0 (0) · g(γ(0)) + f (γ(0)) · (g ◦ γ)0 (0)
= (f∗ )p (X|p ) · g(p) + f (p) · (g∗ )p (X|p ).

In the second-last equality we applied the product rule of calculus.

Remark. Φ is an isomorphism of vector spaces. This can be showed using


the material in §2.4.

Definition 3.9 (Lie bracket of vector fields). The Lie bracket of two vector
fields X, Y ∈ X(M ) is

[X, Y ] := X ◦ Y − Y ◦ X,

using the identification between X(M ) and the derivations of C ∞ (M ).

CHAPTER 3. VECTOR FIELDS 25


Definition 3.10 (Lie algebra). A Lie algebra is a vector space g with a
bilinear map [·, ·] : g × g → g such that

• [X, Y ] = −[Y, X] Skew symmetry

• [X, [Y, Z]] + [Y, [Z, X]] + [Z, [X, Y ]] = 0 Jacobi Identity

Example. (X(M ), [·, ·]) is a Lie algebra. 


Example. The square matrices M (n, R) with [A, B] := AB − BA form a Lie
algebra. The Jacobi identity holds as a consequence of the associativity of
matrix multiplication. 
Remark. Let (U, φ) be a chart on M . Then we get vector fields ∂
∂xi on
U ⊂ M.

• The vector field ∂xi ,



seen as a derivation, maps xj ∈ C ∞ (U ) to δij

• If X = i ai ∂xi ,

i bi ∂xi ,

with ai , bi ∈ C ∞ (U ), then
P P
Y =

X X  ∂bj !
∂aj ∂
[X, Y ] = ai − bi .
∂xi ∂xi ∂xj
j i

This
h canibe seen by applying [X, Y ] to the functions xi . In particular
∂xi , ∂xj = 0, because the ai and bi are constants here.
∂ ∂

Definition 3.11 (F -related vector fields). Let F : M → N be a smooth


map and X ∈ X(M ), Y ∈ X(N ). We say that X and Y are F -related
iff
(F∗ )p (Xp ) = YF (p)
for all points p ∈ M .

Remark. Equivalently: X and Y are F -related iff ∀g ∈ C ∞ (N ) we have


X(F ∗ g) = F ∗ (Y (g)). Here F ∗ is the pullback of functions, i.e. F ∗ (g) = g◦F .

Proposition 3.12. Suppose Xi is F -related to Yi for i = 1, 2. Then


[X1 , X2 ] is F -related to [Y1 , Y2 ].

Proof. Use X1 (X2 (F ∗ (g)) = X1 (F ∗ (Y2 (g))) = F ∗ (Y1 (Y2 (g))).

CHAPTER 3. VECTOR FIELDS 26


p

Figure 3.1: Two p-related vector fields, where p : R2 → R is the first projec-
tion.

3.4 Interpretation of the Lie bracket

Definition 3.13 (Pushforward of a vector field by a diffeomorphism). Given


a diffeomorphism φ : M → N and X ∈ X(M ), we denote by φ∗ X the
unique vector field on N such that X is φ-related to φ∗ X.

Explicitly, we have (φ∗ X)φ(p) = (φ∗ )p (Xp ) for all p ∈ M .

Lemma 3.14. Let φ : M → N be a diffeomorphism, let X ∈ X(M ) and


Y ∈ X(N ). Then Y = φ∗ (X) iff

FtY ◦ φ = φ ◦ FtX for all t s.t. FtX is defined.

Here F X denotes the flow of X, and similarly for Y .

Definition 3.15 (Lie derivative). Let X, Y ∈ X(M ). The Lie derivative


of Y in the direction of X is the vector field
d X
LX Y = (F−t )∗ Y.
dt t=0

CHAPTER 3. VECTOR FIELDS 27


Remark. At every point p,
d X d X
(LX Y )p = [(F−t )∗ Y ]p = [(F−t )∗ YFt (p) ].
dt t=0 dt t=0

These are all tangent vectors in Tp M .


Remark. One can show that LX Y = [X, Y ].

Proposition 3.16. Let X, Y ∈ X(M ). The following are equivalent:

a) LX Y = 0

b) (FtX )∗ Y = Y for all t

c) The flows of X and Y commute: FtX ◦ FsY = FsY ◦ FtX for all t, s.

Proof. • b) ⇒ a). LX Y = d X
dt |0 (F−t )∗ Y = d
dt |0 Y = 0.

• a) ⇒ b). Fix a point p. Consider t 7→ (F−t ∗ FtX (p) . This is a


X) Y

curve in Tp M , which at t = 0 equals Yp . We will show that this is


a constant curve by taking the derivative. For all t0 :
d X d X
(F−t )∗ (YFtX (p) ) = (F−t ) (YF X (p) )
0 −s ∗
dt t0 ds s=0 t0 +s
d 
X X
= (F−t )
0 ∗
(F ) Y
−s ∗ Fs (Ft (p))
X X
 ds s=0  0

X
= (F−t0 )∗ (LX Y )FtX (p) = 0,
0

where in the first equality we set t = t0 + s.

• b) ⇔ c). For all t, apply the last lemma to FtX : M → M .

Corollary 3.17. [X, Y ] = 0 iff the flows commute.


h i
Example. On R2 , we know ∂x∂ 1 , ∂x∂ 2 = 0. Indeed, the flows given by
(x, t) 7→ x + (t, 0) and (x, t) 7→ x + (0, t) commute. 

Proposition 3.18. Let V1 , . . . , Vk be pointwise linearly independent vec-


tor fields on M such that [Vi , Vj ] = 0. Then for every p ∈ M there is a
chart (U, (s1 , . . . , sn )) centered at p such that Vi = ∂s∂ i for i = 1, . . . , k.

CHAPTER 3. VECTOR FIELDS 28


Proof. Since this is a local statement, let’s assume that we’re working in
Rn , and p = 0. Assume that spani {Vi |0 } ⊕ ({0} × Rn−k ) = Rn . Denote
by θi the flow of Vi and let Ω be a small neighborhood of the origin in
{0} × Rn−k . Define

Φ : (−ε, ε)k × Ω −→ Rn
(s1 , . . . , sk , sk+1 , . . . , sn ) 7−→ (θ1 )s1 ◦ · · · ◦ (θk )sk (0, . . . , 0, sk+1 , . . . , sn )T .

Then ∂
∂si is Φ-related to Vi for i ≤ k. Indeed for i ≤ k you can compute
that  ∂ 
(D(s1 ,...,sn ) Φ) = Vi |Φ(s1 ,...,sn )
∂si
by using the curve ε 7→ (s1 , . . . , si + ε, . . . , sn ) and using the fact that
the flows θj commute by the last corollary. At the point p = 0 we clearly
have (D0 Φ)( ∂s∂ i ) = ∂x

i
|0 for i > k. Hence d0 Φ is an isomorphism.
By the inverse function theorem, there exists a neighbourhood V of
0 such that Φ|V : V → Φ(V ) =: U is a diffeomorphism. The desired
chart is the inverse of this diffeomorphism.

(θ2 )s2 ◦ (θ1 )s1 (0, 0, s3 )


(θ2 )s2 ◦ (θ1 )s1 (0, 0)
(θ1 )s1 (0, 0, s3 )
(θ1 )s1 (0, 0) (0, 0, s3 )

(0, 0)

(0, 0, 0)

Figure 3.2: Two examples illustrating the proof. On the left, n = k = 2.


On the right, n = 3 and k = 2.

CHAPTER 3. VECTOR FIELDS 29


Chapter 4

Bundles

4.1 Fiber bundles


Definition 4.1 (Fiber bundle). Let F be a manifold. A fiber bundle with
typical fiber F is a smooth surjective map

π:E→B

between manifolds s.t. for all x ∈ B, there exists an open neighbourhood


U and a diffeomorphism ψ : π −1 (U ) → U × F making this diagram
commute:
ψ
π −1 (U ) U ×F
π
π1

E = S 1 × (−ε, ε)

π −1 (U )
π

B = S1
U

Figure 4.1: Example of a product fiber bundle

Remark. The map π : E → B is a submersion, i.e. (π∗ )e is surjective in all

30
points e of E. Indeed: locally, the map π looks like the projection onto the
first factor π1 , as ψ is a diffeomorphism. So we can work with π1 , which is
a submersion.
Remark. For all x ∈ B, the fiber π −1 (x) is diffeomorphic to F . Indeed,
π −1 (x) ≈ π1−1 (x) = {x} × F ≈ F . So we have a family of manifolds (the
fibers) which all are diffeomorphic to F , and this family is parametrized by
B.
E is called the total space, B the base space and ψ a local trivilization.
A section is a smooth map f : B → E such that π ◦ f = IdB .

E = S 1 × (−ε, ε)
Image of a section

B = S1

Figure 4.2: Definition of a section


π
Example. Given manifolds F and B, the product bundle is B × F −→
1
B. 
Example. Let F be a manifold, φ : F → F a diffeomorphism. Define the
manifold E := ([0, 1] × F )/ ∼, where ∼ is given by

(0, p) ∼ (1, φ(p)).

Let B = S 1 . Then π : E → B is a fiber bundle with typical fiber F .


For instance, take F = (−ε, ε), φ = − Id. Then E is the Möbius strip.

Figure 4.3: Möbius strip

CHAPTER 4. BUNDLES 31

Example. The first projection π1 : \ {0} → R, (x, y) 7→ x is not a fiber
R2
bundle. A reason is that not all fibers are diffeomorphic. 

4.2 Vector bundles


Definition 4.2 (Vector bundle). Fix n ∈ N≥0 . A vector bundle of rank
n is a smooth surjective map π : E → B between manifolds E, B such
that

• Ep = π −1 (p) is a n-dimensional vector space for all p ∈ B

• For all p ∈ B, there exist a neighborhood U of p and a diffeo-


morphism ψ : E|U := π −1 (U ) → U × Rn such that the following
diagram commutes

ψ
π −1 (U ) U × Rn
π
π1
U

and ψ|Eq : Eq → {q} × Rn is a linear isomorphism, for all q ∈ U .

image of the zero sectionB E

Figure 4.4: Definition of a vector bundle

CHAPTER 4. BUNDLES 32
Remark. Equivalently, a vector bundle of rank n is a fiber bundle s.t. each
fiber is an n-dimensional vector space and trivializations are linear in each
fiber.
Remark. • There exists a canonical section B → E, q 7→ (zero vector in Eq ).
It is called zero section.

• Denote by Γ(E) the set of all sections of E, it is a vector space. Note


that we can also multiply a section B → E with a function B → R.
Hence Γ(E) is a module over the algebra C ∞ (B).
π
Example. Given a manifold B, take B × Rn −→
1
B. Notice Γ(B × Rn ) =
∞ n
C (B, R ). 

Proposition 4.3. Let B be a manifold of dimension n. Then


G
T B := Tp B
p∈B

is naturally a vector bundle of rank n, called the tangent bundle of B

Proof. Define π : T B → B, v ∈ Tp B 7→ p. Notice that π −1 (p) = Tp B is


a vector space.
We show that T B is a manifold. Consider a chart φ : U → φ(U ) of
B, denote its components by (x1 , . . . , xn ). It induces a bijection

ψ : (T B)|U −→ U × Rn
n
X ∂
ai 7−→ (p, ai ).
∂xi p
i=1

Now take a cover of B by charts (Uα , φα )α∈A . Consider the topology


on T B with basis ψα−1 (σ) for σ ⊂ Uα × Rn open and α ∈ A. Notice
that ψβ ◦ ψα−1 is a smooth map from (Uα ∩ Uβ ) × Rn to itself. Then
{((T B)|Uα ), ψα }α∈A is aa smooth atlas. Hence T B is a smooth manifold.
For all α ∈ A, the chart ψα : (T B)|Uα → Uα × Rn is a diffeomorphism
and linear on every fiber.
a
Strictly speaking, ψα is not a chart, because Uα × Rn is not an open subset of
R × Rn . But we can identify Uα with φα (Uα ), which is an open subset of Rn .
n

Remark. Γ(T B) = X(B)


Remark. Let π1 : E1 → B and π2 : E2 → B be a vector bundle with the
same base. Their direct sum (Whitney sum) is a vector bundle

E1 ⊕ E2 → B

CHAPTER 4. BUNDLES 33
with fiber over p ∈ B given by (E1 ⊕ E2 )p = (E1 )p ⊕ (E2 )p . Trivialisations
are given by taking the direct sum of the trivializations of E1 and of E2 .

Definition 4.4 (Vector subbundle). Let π : E → B be a vector bundle. A


vector subbundle is a subset D ⊂ E s.t. π|D : D → B, with the induced
smooth structure and vector space structure on the fibers, is a vector
bundle.

Definition 4.5 (Vector bundle morphism). Let πi : Ei → Bi be vector


bundles for i = 1, 2. A smooth map F : E1 → E2 is a vector bundle
morphism

• if there exists a smooth map f : B1 → B2 such that the following


diagram commutes

F
E1 E2
π1 π2
f
B1 B2

• and if for all p ∈ B1 , the map F |(E1 )p : (E1 )p → (E2 )f (p) is linear.

We call F an isomorphism if F is invertible and if F, F −1 are both vector


bundle morphisms (or equivalently, if F is a diffeomorphism).

Definition 4.6 (Frame). Given a vector bundle E → B, a frame is a


collection of sections that form a basis at every point.

Remark. E admits a frame iff E is isomorphic to the product vector bundle


B × Rn . In this case we call E a trivial vector bundle.
Example. Let S n ⊂ Rn+1 be the unit sphere. The tangent bundle T S n and
its orthogonal (T S n )⊥ are vector bundles over S n .
• (T S n )⊥ is a trivial bundle. Indeed, we can choose a unit normal vector
at each point as a frame.
• T S n is not a trivial bundle in general. (T S n only admits a frame when
n = 1, 3. The fact that T S 2 does not admit a frame is a consequence
of the Hairy ball theorem.)
• The Whitney sum T S n ⊕ (T S n )⊥ is isomorphic to Rn+1 × S n , which
is a trivial vector bundle. So this is an example of a sum of a trivial
and a non-trivial vector bundle being trivial.

CHAPTER 4. BUNDLES 34


S2

Figure 4.5: The Whitney sum of T S 2 ⊕ (T S 2 )⊥ is a trivial vector bundle.

CHAPTER 4. BUNDLES 35
Chapter 5

Differential forms and


integration

5.1 Forms on vector spaces


Let V be a finite dimensional real vector space.

Definition 5.1. For every k ≥ 1, we define


k
^
V ∗ := {ω : V × · · · × V → R | multilinear and skew symmetric}
| {z }
k times

Skew symmetric means: ω(. . . , v, . . . , w, . . .) = −ω(. . . , w, . . . , v, . . .).


We define 0 V ∗ = R
V

V1
Example. V∗ =V∗ 
V2
Example. V = bilinear maps V × V → R that are skew symmetric.
∗ 

Definition 5.2 (Wedge product). ∀`, k ≥ 0, the wedge product is


k
^ `
^ `+k
^
∧: V∗× V ∗ −→ V∗
1 X
(ω ∧ τ )(v1 , . . . , vk+` ) 7−→ (−1)σ ω(vσ(1) , . . . , vσ(k) ) · τ (vσ(k+1) , . . . , vσ(k+`) ),
k!`!
σ∈Sk+`

where Sk+` denotes the permutation group.


Vk V`
Remark. ω ∧ τ = (−1)k` τ ∧ ω, where ω ∈ V ∗ and τ ∈ V ∗.

36
Lemma 5.3. Suppose θ1 , · · · , θk ∈ V ∗ and v1 , . . . , vk ∈ V . We have

(θ1 ∧ · · · ∧ θk )(v1 , . . . , vk ) = det(θi (vj )).

Lemma 5.4. Let m = dim V . Suppose θ1 , . . . , θm is a basis of V ∗ . Then


∀k ≥ 1,
{θi1 ∧ · · · ∧ θik | 1 ≤ i1 < . . . < ik ≤ m}
Vk ∗
is a basis of V .
Vk
Remark. If k > m, the dimension of the space, then V ∗ is the zero vector
space.
Example. Suppose dim V = 3 and (θ1 , θ2 , θ3 ) is a basis of V ∗ , then

• 1 is a basis of 0 V ∗ = R
V

• {θ1 , θ2 , θ2 } is basis of 1 V ∗ = V ∗
V

• {θ1 ∧ θ2 , θ2 ∧ θ3 , θ1 ∧ θ3 } is a basis of 2 V ∗
V

• θ1 ∧ θ2 ∧ θ3 is a basis of 3 V ∗ .
V

Definition 5.5 (Pullback). Let f : V → W be a linear map. Then the


dual map is f ∗ : W ∗ → V ∗ , given by (f ∗ θ)(v) = θ(f (v)). More generally,
∀k ≥ 1, the pullback by f is
k
^ k
^
∗ ∗
f : W −→ V∗
ω 7−→ f ∗ ω

where
(f ∗ ω)(v1 , . . . , vk ) = ω(f (v1 ), . . . , f (vk )).

5.2 Differential forms on manifolds


Let M be a manifold, f ∈ C ∞ (M ). For all p ∈ M , define

(df )p := (f∗ )p : Tp M → Tf (p) R = R

This is a linear map, i.e. (df )p ∈ (Tp M )∗ =: Tp∗ M


Let (U, φ = (x1 , . . . , xn )) be a chart.

CHAPTER 5. DIFFERENTIAL FORMS AND INTEGRATION 37


Lemma 5.6. The set {dx1 |p , dx2 |p , . . . , dxn |p } is a basis of Tp∗ M . It is
the basis dual to ∂x
 ∂
i p

Proof. We have that ∂x ∂


i p = (φ
−1 )
∗,φ(p) ei , where ei is the i-th standard
basis vector. Hence, for every j,
 

= dxj |p (φ−1 )∗,φ(p) ei = (xj ◦ φ−1 )∗,φ(p) ei .

dxj |p i
∂x p

Since xj ◦ φ−1 is the jth projection, this expression is the jth component
of ei , which is δij .

Example. Consider R2 with standard coordinates x1 , x2 . Then dx1 |p : Tp R2 →


R is given by ∂x∂ 1 |p 7→ 1 and ∂x∂ 2 |p 7→ 0. 
Vk ∗
Let k ≥ 0. For every p ∈ M , consider the vector space Tp M .

Definition 5.7 (Differential form). A k-form on M is a map


k
G ^
α:M → Tp∗ M
p∈M

such that
Vk
• α(p) ∈ Tp∗ M for each p

• For any chart (U, φ), writing


X
α(p) = ai1 ...ik dxi1 |p ∧ · · · ∧ dxik |p ,
1≤i1 <...<ik ≤m

the coefficients ai1 ...ik : U → R are smooth.


Vk
Remark. In other words: a k-form is a section of the vector bundle T ∗M .
We denote
Ωk (M ) = {k-forms on M }.
Notice that Ω0 (M ) = C ∞ (M ).
Remark. Given α ∈ Ωk (M ) and vector fields X1 , . . . , Xk ∈ X(M ), we define
α(X1 , . . . , Xk ) ∈ C ∞ (M ) by

(α(X1 , . . . , Xk ))(p) := α(p)(X1 (p), . . . , Xk (p)).

Notice that for all f ∈ C ∞ (M ), α(f X1 , . . . Xk ) = f α(X1 , . . . , Xk ).

CHAPTER 5. DIFFERENTIAL FORMS AND INTEGRATION 38


Definition 5.8. Let F : M → N be a smooth map. The pullback of
k-forms is F ∗ : Ωk (N ) → Ωk (M ) :

(F ∗ ω)(p)(v1 , . . . vk ) = ω(F (p))((F∗ )p v1 , . . . , (F∗ )p vk )

where p ∈ M , vi ∈ Tp M .

We now study the pullback of differential forms on Rn .

Proposition 5.9. Let U ⊂ Rm open, V ⊂ Rn open, G : U → V smooth.


Denote with xj the standard coordinates on Rm and by yi those on Rn .
Then
Pm ∂Gi
1) G∗ (dyi ) = j=1 dxj ∈ Ω1 (U )
∂xj
2) If m = n, then

G∗ (f dy1 ∧ . . . ∧ dym ) = (f ◦ G) det(Jac G) dx1 ∧ . . . ∧ dxm .

Above Jac G denotes the Jacobian of G, i.e. the matrix representing the
derivative DG of G.

Proof. 1) We have
    !
∗ ∂ ∂ X ∂Gk ∂ ∂Gi
G (dyi ) = dyi G∗ = dyi = .
∂xj ∂xj ∂xj ∂yk ∂xj
k

2) We have G∗ (f dy1 ∧ . . . ∧ dym ) = (G∗ f )(G∗ dy1 ∧ . . . ∧ G∗ dym ).


Evaluating on ∂x∂ 1 , . . . , ∂x∂m , we get
  
∗ ∗ ∂
(G f ) det G (dyi ) = (f ◦ G) det(Jac G).
∂xj
| {z }
∂Gi
= by above
∂xj

CHAPTER 5. DIFFERENTIAL FORMS AND INTEGRATION 39


5.3 Orientation and volume forms
Definition 5.10 (Oriented atlas). An oriented atlas for M is a smooth
atlas {(Uα , φα )} s.t.
det D(φβ ◦ φ−1

α ) >0

for all α, β.

Definition 5.11 (Orientation). An orientation is a choice of maximal ori-


ented atlas.

Remark. Not all manifolds are orientable, e.g. the Möbius band and RP2 are
not.
Remark. Let V be vector space. On {ordered bases of V }, there is an equiv-
alence relation, given by: (v1 , . . . , vm ) ∼ (w1 , . . . , wm ) if the change of basis
has det > 0. An orientation of V is by definition a choice of one of the two
equivalence classes. An ordered basis of an oriented vector space is positive
if it belongs to the equivalence class that gives the orientation.
Now let M be a manifold. An orientation on M induces an orientation
on each tangent space Tp M .

Definition 5.12 (Volume form). A volume form on M m is an m-form Ω


such that Ω(p) 6= 0 for all p ∈ M .

Remark. If Ω is a volume form, then any other volume form looks like f Ω,
where f : M → R \ {0}.

Proposition 5.13. M is orientable iff there exists a volume form Ω

Proof. We denote by Vol := dx1 ∧ . . . ∧ dxm the standard volume form


on Rm .
⇐: choose an atlas consisting of charts such that (φ−1 α ) Ω = fα Vol,

where fα is a positive function. Then det D(φβ ◦ φα ) > 0 because of


−1

item 2 of the last proposition. So {(Uα , φα )} is an oriented atlas.


⇒: given an oriented atlas {(Uα , φα )}, choose a partition of unity
{eα } subordinate to it, and define
X
Ω= eα · (φα )∗ Vol.
α

Then Ω is a volume form, because eα ≥ 0 for all α and on Uα ∩ Uβ we

CHAPTER 5. DIFFERENTIAL FORMS AND INTEGRATION 40


have

(φβ )∗ Vol = (φα )∗ (φβ ◦ φ−1 ∗ ∗


α ) Vol = (a positive function) · (φα ) Vol.
| {z }
det(D(φβ ◦φ−1
α ))·Vol

5.4 Integration on manifolds


Let M m be a oriented manifold. We want to define ω, where ω ∈ Ωm (M )
R
M
has compact support.

Step 1. Assume the support supp(ω) := {p ∈ M : ω(p) 6= 0} is contained


in one chart (U, φ) of the oriented atlas. Write (φ−1 )∗ ω = f dx1 ∧ . . . ∧ dxm
for some f ∈ C ∞ (φ(U )).

Definition 5.14 (Integral of top form supported in a chart).


Z Z Z
−1 ∗
ω := (φ ) ω := f (x) dx1 dx2 · · · dxm ,
M φ(U ) φ(U )

where the right hand side is the multiple Riemann integral of the func-
tion f over φ(U ).

Remark. Recall the transformation rule in Rm . Let U, V ⊂ Rm open and


θ : V → U a diffeomorphism. Let f ∈ C ∞ (U ) be integrable. Then
Z Z
f (x) dx1 · · · dxm = (f ◦ θ)(y) det(Jac θ) dy1 · · · dym .
U V

ω is independent of the choice of chart in the oriented


R
Lemma 5.15. M
atlas.

Proof. Let (V, ψ) be another chart as above, and denote θ := φ ◦ ψ −1 .

CHAPTER 5. DIFFERENTIAL FORMS AND INTEGRATION 41


V
U
supp(ω)

φ ψ
Rm
φ(U ) θ ψ(V )
Rm

We want to find g ∈ C ∞ (ψ(V )) such that (ψ −1 )∗ ω = g(y) dy1 ∧


dy2 · · · ∧ dym . To do so we compute

(ψ −1 )∗ ω = (φ−1 ◦θ)∗ ω = θ∗ ((φ−1 )∗ ω) = (f ◦ θ) det(Jac θ) dy1 ∧. . .∧dym


| {z }
so this is g

using a Proposition from §5.2 in the last equality. By the transformation


rule
Z Z
f (x) dx1 dx2 · · · dxm = (f ◦ θ)(y)| det(Jac θ)| dy1 · · · dym ,
φ(U ) ψ(V ) | {z }
=g

finishing the proof. (The function on the right hand side equals g since
the absolute values can be removed, due to the fact that ψ and φ lie in
the oriented atlas of M .)

Step 2. For any ω ∈ Ωm (M ) with compact support, let {(Uα , φα )} be an


oriented atlas such that {α : supp(ω)∩Uα 6= ∅} is finite.
P Let eα be
P a partition
of unity subordinate to this cover. Notice that ω = ( eα )ω = (eα ω), and
the sum on the right is finite.

Definition 5.16 (Integral of top form).


Z XZ
ω= eα ω.
M α

Notice that each summand eα ω was defined in Step 1, since supp(eα ω) ⊂


R

Uα .
Remark. One can show: this definition is independent of the choice of ori-
ented atlas and partition of unity.

CHAPTER 5. DIFFERENTIAL FORMS AND INTEGRATION 42


Chapter 6

The exterior derivative and


Stokes theorem

6.1 The exterior derivative in Rm


Let U ⊂ Rm be open. Denote by x1 , . . . , xm the standard coordinates on
Rm .
Remark. If f ∈ C ∞ (U ), then df = f∗ lies in Ω1 (U ), because for all p ∈ U
we have (f∗ )p ∈ Tp∗ M . Expressing df in terms of the dxi , we have
X ∂f
df = dxi ,
∂xi
i
 
∂f
since (f∗ )p ∂
∂xi p = (Dp f )(ei ) = ∂xi (p).

Example. On R2 , we have d(xy) = ydx + xdy. 

Definition 6.1 (Exterior derivative on Rm ). Let U ⊂ Rm be open. The


exterior derivative (or de Rham differential)

d : Ωk (U ) −→ Ωk+1 (U )

is defined as follows. For k = 0: as above. For k > 0:


 
X X
d ai1 ...ik dxi1 ∧ · · · ∧ dxik  := (dai1 ...ik )∧dxi1 ∧· · ·∧dxik
1≤i1 <...<ik ≤m

Notice that here ai1 ...ik ∈ C ∞ (U ).


Example. On R3 consider ω = (x1 )3 dx2 ∧ dx3 , then

dω = 3(x1 )2 dx1 ∧ dx2 ∧ dx3 .

43

Example.
d (x1 )2 dx1 ∧ dx2 = 0


Proposition 6.2. The exterior derivative satisfies the following:

i) d is R-linear

ii) d(α ∧ β) = dα ∧ β + (−1)k α ∧ dβ, where α ∈ Ωk (U ).

iii) d2 = 0.

iv) If F : U → V smooth, then d(F ∗ ω) = F ∗ (dω).

Proof. iii) We prove it only for g ∈ C ∞ (U ) = Ω0 (U ).


 X ∂g 
d(dg) = d dxj
∂xj
j
X  ∂g 
= d ∧ dxj
∂xj
j
X ∂2g
= dxi ∧ dxj .
∂xi ∂xj
j,i

2
Now dxi ∧ dxj is anti-symmetric in i and j, while ∂x∂i ∂x
g
j
is sym-
metric in i and j (partial derivatives commute!), so d = 0.
2

iv) We prove it only for g ∈ C ∞ (U ). Let p ∈ U , v ∈ Tp U = Rm . Then

(F ∗ (dg))v = dg((F∗ )p v) = ((g ◦ F )∗ )p v = (F ∗ g)∗ v = d(F ∗ g)v,

using the chain rule in the second equality.

6.2 The exterior derivative on manifolds


Let M be a manifold.

CHAPTER 6. THE EXTERIOR DERIVATIVE AND STOKES 44


THEOREM
Definition 6.3 (Exterior derivative on manifolds). Let ω ∈ Ωk (M ), then
dω ∈ Ωk+1 (M ) is defined as follows: for all charts (U, φ),
  
(dω)|U = φ∗ d (φ−1 )∗ ω .

Notice that (φ−1 )∗ ω is a k-form on an open subset of Rm .


Remark. The above is well defined because if (V, ψ) is another chart, then
the map (ψ ◦ φ−1 )∗ commutes with d by part iv) of the previous proposition.

6.3 Manifolds with boundary

Definition 6.4. Hm = {(x1 , . . . , xm ) ∈ Rm | x1 ≤ 0}

Definition 6.5 (Differentiable map on a open subset of Hm ). Let U ⊂ Hm


be open. Then f : U → Rk is differentiable iff there exists an open Ũ ⊂
Rm such that U = Ũ ∩ Hm , and there exists f˜ : Ũ → Rk differentiable
such that f˜|U = f . In that case, for all p ∈ U , Dp f = Dp f˜.

Definition 6.6 (Manifold with boundary). A manifold with boundary of


dimension m consists of topological space M that is second countable
and Hausdorff, together with a maximal smooth atlas.
Here by smooth atlas we mean: an open cover {Uα } on M and
homeomorphisms φα from Uα to open subsets of Hm , such that φβ ◦ φ−1
α
is differentiable for all α, β.

Remark. For m = 1, additionally we allow homeomorphisms φα from Uα to


open subsets of {x ∈ R | x ≥ 0}.
Remark. In particular, all manifolds are manifolds with boundary.

Definition 6.7 (Boundary). The boundary of M is

∂M = {p ∈ M : ∃ chart (U, φ) : φ(p) ∈ ∂Hm },

where ∂Hm = {0} × Rm−1 .

Remark. If one charts satisfies this property, all charts satisfy this property.
Example. Consider M = {v ∈ Rm : kvk ≤ 1}. Then the boundary of M is
∂M = {v ∈ Rm : kvk = 1}. 
Remark. For all p ∈ ∂M , the tangent space Tp M is defined similarly to the
case of manifolds (without boundary), and has dimension m.

CHAPTER 6. THE EXTERIOR DERIVATIVE AND STOKES 45


THEOREM
Remark. The integration of differential forms on manifolds with boundary
is analog to the one for manifolds.
One can show the following:

Proposition 6.8. ∂M is a manifold, of one dimension less than M .

Proposition 6.9. An orientation on M induces an orientation on ∂M ,


as follows:

∀p ∈ ∂M : (v1 , . . . , vm−1 ) is a positive basis of Tp ∂M


⇔ (e, v1 , . . . , vm−1 ) is a positive basis of Tp M ,

where e ∈ Tp M is “outward pointing”.

Example. Note that R2 has a standard orientation (the one determined by


the ordered basis (e1 , e2 )). Hence the unit disk D ⊂ R2 too. The induced
orientation on the circle ∂D is the “anticlockwise” one.
To see this: (e, v1 ) as in the figure is a positive basis of R2 . Hence the
orientation that D induces on ∂D is the one for which v1 is a positive basis.

v1
e

Figure 6.1: Orientation of the boundary of a disk

6.4 Stokes’ theorem


Rb
We generalize the fundamental theorem of calculus: a f 0 (x)dx = f |ba .

CHAPTER 6. THE EXTERIOR DERIVATIVE AND STOKES 46


THEOREM
Theorem 6.10 (Stokes). Let M m be an oriented manifold with boundary
and ω ∈ Ωm−1 (M ) with compact support. Then
Z Z
dω = i∗ ω,
M ∂M

where i : ∂M → M is the inclusion and ∂M has the induced orientation.

Remark. Note that since ω has compact support, so does dω.


The idea of the proof is to use a partition of unity to reduce this to the case
where supp(ω) is contained in a chart, i.e. to reduce to M = Hm . There one
can apply suitably the fundamental theorem of calculus.
Example. Let D = {v ∈ R2 : kvk ≤ 1}. We have
Z Z Z

i (xdy − ydx) = d(xdy − ydx) = 2 dx ∧ dy = 2π.
S1 D D

Corollary 6.11. Let M be a manifold (without boundary) and ω ∈


Ωm−1 (M ) with compact support. Then
Z
dω = 0.
M

Corollary 6.12. If M is compact, orientable manifold with boundary,


then there is no smooth f : M → ∂M such that f |∂M = Id∂M .

Proof. Since M is orientable, ∂M is too. Therefore, there exists a vol-


ume form on the boundary, call it ω. If the boundary has several con-
nected components, choose ω so that the integral over each boundary
component is, say, positive. Suppose f exists. Then d(f ∗ ω) = f ∗ dω = 0,
because dω = 0 by degree reasons. So
Z Z Z Z
∗ ∗ ∗ ∗
0= d(f ω) = i f ω= (f ◦ i) ω = ω 6= 0,
M ∂M ∂M ∂M

using Stokes’ theorem in the second equality and in the inequality the
fact that ω is a volume form chosen as above.

CHAPTER 6. THE EXTERIOR DERIVATIVE AND STOKES 47


THEOREM
Chapter 7

De Rham Cohomology

7.1 Basic definitions


Let M be a manifold of dimension m.

Definition 7.1 (Closed forms). ω ∈ Ωk (M ) is called closed iff dω = 0.

Definition 7.2 (Exact forms). ω ∈ Ωk (M ) is called exact iff dα = ω for


some α ∈ Ωk−1 (M ).

We have
d0 d1 dm−1
0 Ω0 (M ) Ω1 (M ) ··· Ωm (M ) 0

with dk ◦ dk−1 = 0, so in particular the image of dk−1 is included in the


kernel of dk . This is an example of a cochain complex. Cochain because d
increases the degree.

Ωk−1 Ωk Ωk+1

exact
closed
dk−1 dk

Figure 7.1: Visualization of the de Rham cochain complex. Exact k-forms


lay per definition in the image of dk−1 , and closed k-forms in the kernel of
dk . As d2 = 0, the exact k-forms form a subspace of the closed k-forms.

48
Definition 7.3 (De Rham cohomology). The k-th de Rham cohomology
group of M is
Ker dk closed k-forms
H k (M ) := HdR
k
(M ) := = .
Im dk−1 exact k-forms

If ω ∈ Ωk (M ) is closed, we denote by [ω] ∈ H k (M ) its class.

Ωk−1 Ωk Ωk+1

dk−1 dk
π

Hk

Figure 7.2: Definition of the De Rham cohomology.

Remark. Each H k (M ) is a vector space. H k (M ) can be different from zero


only when 0 ≤ k ≤ m. If M is compact, one can show that H k (M ) is finite
dimensional for all k. This is not trivial, because the dimension of Ker dk is
usually infinite-dimensional.

Proposition 7.4. If M is connected, then H 0 (M ) ∼


= R.

Proof. For all f ∈ Ω0 (M ), f is closed if df = 0, so f is constant (M is


connected and f is smooth). f is exact iff it is zero, since Ω−1 (M ) =
{0}.

Remark. If M is not connected, we get Rk , with k being the number of


connected components.

Theorem 7.5. If M m is compact, connected and orientable, then H m (M ) ∼


=
R.

CHAPTER 7. DE RHAM COHOMOLOGY 49


Proof. Forms of ‘top degree’ are always closed. The idea is that we
construct a surjective map from the m-forms to R and then show that
the kernel is given precisely by the exact m-forms. Consider

I : Ωm (M ) −→ R
Z
ω 7−→ ω.
M

Then I Ris surjective: since M is orientable, there exists a volume form,


Ω, and M Ω 6= 0. We now argue that ker(I) = {exact m-forms}.
For the inclusion “⊃”: consider an exact m-form ω = dα. Then by
Stokes’ theorem Z Z
dα = α = 0.
M ∂M

Idea for the inclusion “⊂”: Assume M ω = 0, we need to show that


R

ω = dα for some α. Reduce to forms with support (necessarily compact)


contained in a chart. Then this reduces to the following: if f isRa function
x
on R with compact support and R f (x)dx = 0, then F (x) = −∞ f (t)dt
R

is a primitive of f with compact support (because when xR is really small,


this integral is zero, and when x is large enough, F (x) = R f (x)dx = 0).
This shows that
Ωm (M )
= H m (M ).
Ker(I)
But this quotient is isomorphic to the image of I, which is R.

7.2 Homotopic maps and cohomology

Lemma 7.6. Let f : N → M be a smooth map. Then

• If ω ∈ Ω(M ) is closed, then f ∗ ω is also closed.

• If ω ∈ Ω(M ) is exact, then f ∗ ω is also exact.

Proof. We only prove the first part. d(f ∗ ω) = f ∗ dω = f ∗ 0 = 0.

Given f : N → M , consider the pullback of differential forms f ∗ :


Ωk (M ) → Ωk (N ). This restricts to a map between closed forms, and also
between exact forms, so it induces a map on the level of cohomology groups:

H(f ) : H k (M ) −→ H k (N )
[ω] 7−→ [f ∗ ω].

CHAPTER 7. DE RHAM COHOMOLOGY 50


Remark. If f is a diffeomorphism, then it’s clear that H(f ) is an isomor-
phism. But there are more general maps that induces an isomorphism in
cohomology, as we now explain.

Definition 7.7 (Smooth homotopy). Let f, g : N → M be smooth maps.


A smooth homotopy between f and g is a smooth map h : N ×[0, 1] → M
such that h|N ×{0} = f and hN ×{1} = g .

In other words, h is a smooth family of maps that interpolate between f and


g. One can prove:

Proposition 7.8. If f, g : N → M are smoothly homotopic, then f ∗ and


g ∗ are cochain homotopic, i.e. there exists a linear map

K : Ω• (M ) → Ω•−1 (N ),

such that d ◦ K + K ◦ d = f ∗ − g ∗ .

d d
Ωk−1 (M ) Ωk (M ) Ωk+1 (M )
f ∗ −g ∗
K K
d d
Ωk−1 (N ) Ωk (N ) Ωk+1 (N )

Theorem 7.9. If f, g are smoothly homotopic, then H(f ) = H(g).

Proof. Let K = Ω• (M ) → Ω•−1 (M ) a cochain homotopy between f ∗


and g ∗ . Let ω ∈ Ωk (M ) be closed. Then

f ∗ ω − g ∗ ω = d(Kω) + K(|{z}
dω )
=0

is exact. Hence [f ∗ ω] = [g ∗ ω].

Corollary 7.10. Let M and N be homotopy equivalent, i.e. ∃α : N → M


and β : M → N such that α ◦ β ' IdM and β ◦ α ' IdN , where ' means
smoothly homotopic. Then H(α) : H(M ) → H(N ) is an isomorphism
with inverse H(β).

Proof. H(α) ◦ H(β) = H(β ◦ α) = H(IdN ), because of the previous

CHAPTER 7. DE RHAM COHOMOLOGY 51


theorem, and H(IdN ) = IdH(N ) . You can do the same for the reverse
order.
Remark (Poincaré Lemma). Let U ⊂ Rn be an open subset which is star
shaped1 .
Then U is homotopy equivalent to a point {p}, via the inclusion α :
{p} → U and the constant map β : U → {p}. Hence,
(
k ∼ k ∼ R if k = 0
H (U ) = H ({p}) =
0 if k > 0
Notice that Rn is star-shaped, and every point of a manifold has a star-
shaped neighborhood.
Example. If E → M is a vector bundle, then E is homotopic equivalent to
M , as we can shrink the fibers (which are vector spaces) to the zero section.
So the cohomology of M is the same as that of E: H(E) ∼ = H(M ). 
Remark. Given manifolds M, N , if H(M ) = ∼
6 H(N ), then the two manifolds
are not homotopic equivalent.

7.3 The Mayer–Vietoris theorem

Theorem 7.11. Suppose a manifold admits a cover by 2 open sets U


and V , i.e M = U ∪ V . Then

··· H k−1 (U ∩ V )

Ak Bk
H k (M ) H k (U ) ⊕ H k (V ) H k (U ∩ V )
δk

H k+1 (M ) ···

is a long exact sequence, i.e. the image of any arrow is the kernel of the
next one. Here,

Ak [ω] := [ω|U ] ⊕ [ω|V ]


Bk (α ⊕ β) := β|U ∩V − α|U ∩V .

Remark. The proof uses the smooth maps


iU ∩V,V
U ∩ V −−−−→ U t V → M
iU ∩V,U

1
∃p such that for all u ∈ U the line segment between u and p lies in U .

CHAPTER 7. DE RHAM COHOMOLOGY 52


induced by inclusions.
Example. Let’s consider the sphere S m with m ≥ 2. Write

Sm = U ∪ V

where U = S m \ {North pole}, V = S m \ {South pole}. Notice that

• U ' {p},

• V ' {p},

• U ∩ V ' S m−1 ,

where ' means homotopic equivalent. Applying the theorem and doing some
diagram-chasing, we get
(
k m ∼ R if k = 0, m
H (S ) =
0 otherwise.


CHAPTER 7. DE RHAM COHOMOLOGY 53


Chapter 8

Foliations

8.1 Immersed submanifolds


Definition 8.1 (Immersion). An immersion is a smooth map F : H → M
such that (F∗ )p is injective for all p ∈ H.

Definition 8.2 (Immersed submanifold). An immersed submanifold of M


is a subset H ⊂ M with a topology (not necessarily the subspace topol-
ogy) and a smooth manifold structure such that the inclusion i : H → M
is an immersion.

Remark. In that case, for all p ∈ H, the map (i∗ )p is an isomorphism from
Tp H to (i∗ )p (Tp H) ⊂ Tp M .
Remark. All submanifolds are immersed submanifolds. Immersed submani-
folds are not necessarily submanifolds.
Remark. If H is an immersed submanifold, then for all p ∈ H, there exists
an open neighbourhood U in H (“open” w.r.t. the topology of H) such that
U is a submanifold of M .
Example. For all λ ∈ R consider F : R → S 1 × S 1 : t 7→ (e2πit , e2πiλt ). If
λ ∈ Q, then F (R) is a submanifold of S 1 × S 1 . If λ 6∈ Q, then F (R) is just
an immersed submanifold with the smooth structure given by the bijection
R∼= F (R) obtained from F . The topology on F (R) differs from the subspace
topology induced by S 1 × S 1 .

S1 × S1
F i

R F (R)

54
8.2 Distributions and involutivity

Definition 8.3 (Distribution). A distribution D on M is a subbundle of


TM.
For all p ∈ M , we get a subspace Dp ⊂ Tp M of constant dimension varying
smoothly with p.

Definition 8.4 (Involutive distribution). A distribution D is involutive iff


for all X, Y ∈ Γ(D), we have [X, Y ] ∈ Γ(D)

Remark. D is involutive iff for all p ∈ M , there exists a neighborhood U ⊂ M


and a frame X1 , . . . , Xk ∈ Γ(D|U ) such that [Xi , Xj ] ∈ Γ(D|U ).

Definition 8.5 (Integral manifold). An integral manifold of D is a (non-


empty) immersed submanifold H of M such that Tp H = Dp for all
p ∈ H.

Example. Let X be a nowhere vanishing vector field on M . Then D given


by Dp = span Xp is a rank 1 distribution, thus involutive. The image of any
integral curve of X is an integral manifold of D. 
n o
Example. On R3 , D = span ∂x ∂ ∂
, ∂y is an involutive rank-2 distribution.
The planes {z = c} ⊂ R3 are integral manifolds. 
n o
Example. On R3 , D = span ∂x ∂ ∂
, ∂y ∂
− x ∂z is not involutive. Indeed,
 
∂ ∂ ∂ ∂
, −x =− 6∈ D.
∂x ∂y ∂z ∂z


Proposition 8.6. Let D be a distribution. Suppose that every point of


M is contained in an integral manifold of D. Then D is involutive.

Proof. Let X, Y ∈ Γ(D), p ∈ M . Let H be an integral manifold through


p. Then X|H , Y |H are tangent to H, so [X, Y ]|H = [X|H , Y |H ] is tangent
to H. (To see this equality, use the naturality of the Lie bracket applied
to the smooth map i : H → M .) In particular, [X, Y ]p ∈ Tp H = Dp .

8.3 The Frobenius theorem


Let D be a rank k distribution on M m .

CHAPTER 8. FOLIATIONS 55
Definition 8.7 (Flat chart for D). A chart (U, φ) is flat for D if φ(U ) =
I1 × · · · × In ⊂ Rn and D = span ∂x∂ 1 , . . . , ∂x∂ k on U . Here, Ii are open
intervals.

Definition 8.8 (Completely integrable distribution). D is completely inte-


grable if for all p ∈ M , there exists a flat chart containing p.

Theorem 8.9 (Frobenius). D is completely integrable iff D is involutive.

Notice that the condition on the left is a local one, while the condition on
the right is an infinitesimal one (thus easier to check).

Proof. ⇒ Apply the last proposition.


Alternatively: At each p ∈ M , take a flat chart. Then D =
span ∂x∂ 1 , . . . , ∂x∂ k and ∂x
∂ ∂

i , ∂xj = 0 ∈ Γ(D).

⇐ One can show that near every point p ∈ M , D has a frame of vector
fields X1 , . . . , Xk s.t. [Xi , Xj ] = 0. Then apply Prop. 3.18. (The
last proposition in §3.4, stating that there is a chart (U, (s1 , . . . , sn ))
centered at p such that Xi = ∂s∂ i for i = 1, . . . , k.)

One can prove:

Proposition 8.10. Let D be an involutive rank-k distribution, H an


integral manifold. For any flat chart (U, (x1 , . . . , xn )) for D, H ∩ U is
the union of countably many open subsets of slices {xi = const} (i > k).

Example. On the torus S 1 × S 1 , let D = span{ ∂x∂ 1 + λ ∂x∂ 2 }, where λ ∈ R \ Q.


Every line with slope λ in S 1 × S 1 is an integral manifold, and the countable
condition holds. 

8.4 Foliations
Let M n be a smooth manifold.

CHAPTER 8. FOLIATIONS 56
Definition 8.11 (Foliation). A rank k foliation is a collection {Lα }α∈A of
k-dimensional, connected, immersed submanifolds of M , called leaves,
such that

• M = α∈A Lα (disjoint union)


F

• For all p ∈ M , there exists a chart (U, φ) around p such that


φ(U ) = I1 × · · · × In and for all α ∈ A the following holds: U ∩ Lα
is a countable union of slices {xk+1 = const, . . . , xn = const}, or
empty.

Rn−k

φ
M
Rk

Figure 8.1: Definition of a foliation

Example. On R2 \ {0}, circles with radius r > 0 form a foliation. 


Example. On R2 ,the following is not a foliation. (It’s impossible to straighten
all leaves nearby the x-axis.)

CHAPTER 8. FOLIATIONS 57
Figure 8.2: Not a foliation.


The following theorem (sometimes called “Global Frobenius Theorem”)
gives a bijection between global objects on M and infinitesimal objects.

Theorem 8.12. Let M be a manifold. There is a bijection

Foliations ↔ Involutive distributions


{Lα } 7→ D s.t. Dp = Tp (leaf through p)

The inverse map reads

D 7→ {maximal connected integral manifolds of D}.

Proof. We show that both maps are well-defined. It is easy to see that
they are inverses of each other.

→: D, as defined above, is a distribution. Through every point of


M there passes an integral manifold, hence D is involutive by
Proposition 8.6.

←: By the Frobenius theorem, D is completely integrable, in particu-


lar for every point p of M there is an integral manifold through it.
One can show that there is a maximal connected integral manifold
through it (because the union of all integral manifolds through p
is again an integral manifold). The decomposition of M into the
above maximal connected integral manifolds is a foliation, since D
is completely integrable and by Proposition 8.10.

CHAPTER 8. FOLIATIONS 58
CHAPTER 8. FOLIATIONS 59
Chapter 9

Lie groups and Lie algebras

9.1 Lie groups, Lie subgroups

Definition 9.1 (Lie group). A Lie group is a group which is a mani-


fold, such that the multiplication m : G × G → G and the inversion
i : G → G, g 7→ g −1 are differentiable.

Definition 9.2 (Lie group morphism). A Lie group morphism is a group


morphism which is smooth.

Example. (Rn , +) is a Lie group. Check, for instance, that (x, y) 7→ x + y is


smooth. 
Example. We check that

GL(n, R) = {A ∈ Mat(n, R) : A invertible}

is a Lie group. Here Mat(n, R) denotes the real n × n matrices. The set
GL(n, R) is open in the vector space Mat(n, R) ∼
2
= Rn , so it’s a manifold. It
is also a group, and the multiplication is smooth, because
X
(AB)ij = Ajk Bki
k

is a polynomial in the entries of A and B. The inversion is also smooth: for


n = 2, for instance,
 −1  
a b 1 d −b
= ,
c d ad − bc −c a

and similarly for arbitrary n. 


Remark. GL(n, R) has 2 connected components: det > 0 and det < 0.

60
Example. We check that

SL(n, R) = {A ∈ Mat(n, R) : det(A) = 1}

is a Lie group.
The map det : Mat(n, R) → R is smooth, because det(A) is a polynomial
in the entries of A. We check that 1 ∈ R is a regular value of det, i.e. for all
A ∈ SL(n, R), the derivative dA det : TA Mat(n, R) = Mat(n, R) → T1 R = R
is surjective:
 
d d
(dA det) (t + 1)A = det((t + 1)A)
dt 0 dt 0
d
= (t + 1)n det A
dt 0
= n det(A) = n 6= 0.

Hence SL(n, R) = det−1 (1) is a submanifold of Mat(n, R), by the regular


value theorem, and thus a submanifold of its open set GL(n, R). The mul-
tiplication and inversion of SL(n, R) are smooth, being the restrictions of
those of GL(n, R). 
Example. The following are all examples of Lie groups:

• O(n), defined as matrices with A−1 = AT , which has two components.

• SO(n), defined as matrices with A−1 = AT and det(A) = 1. It is the


connected component of the identity of O(n).
T
• U(n), defined as matrices A ∈ Mat(n, C) such that AA = 1.
T
• SU(n), defined as matrices A ∈ Mat(n, C) such that AA = 1 and
det A = 1.

Definition 9.3 (Left translation, left invariant vector field). Let G be a Lie
group. For all g ∈ G, the diffeomorphism

Lg : G → G : h 7→ gh

is called left translation. A vector field X ∈ X(G) is left invariant iff

∀g ∈ G : (Lg )∗ X = X.

Notice that Lg is the restriction to {g} × G of the multiplication map m :


G × G → G.

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 61


Remark. There is a linear isomorphism

Te G −→ X(G)L := {left-invariant vector fields}


v 7−→ ←

v where (←

v ) = ((L ) ) v.
g g ∗ e

The idea is that if you have a left invariant vector field, then it is deter-
mined by its value at any point, for example e, so that Xg = [(Lg )∗ X]g =
((Lg )∗ )e (Xe ) ∈ Tg G.
It follows that:

• The tangent bundle T G is a trivial vector bundle. Indeed,

G × Te G −→ T G
(g, v) 7−→ ((Lg )∗ )e v.

is a vector bundle isomorphism.

• G admits a volume form, hence it’s orientable.

Example. Of all the spheres, only S 0 , S 1 , S 3 are Lie groups. (S 2 is not a Lie
group; for instance, the Hairy ball theorem implies that the tangent bundle
is not trivial.) 

Proposition 9.4. A connected Lie group G is generated (as a group) by


any open neighbourhood W of identity.

Proof. Let H be the subgroup generated by W , i.e. finite products of


elements of W and W −1 .

H is open. Indeed W1 = W ∪ W −1 is open. W2 := W1 · W1 =


g∈W1 gW1 is open as the left translation is smooth and any union
S
of open subsets is open. Similarly, for all k we get that Wk :=
W1 · Wk−1 is open. Thus H = ∪k≥1 Wk is open.

H is non-empty, as e ∈ H.

The complement G − H is open. Indeed,


[
G−H = gH
g6∈H

is also open, as the union of open sets.

Since G is connected, it follows that G − H is empty, i.e. H = G.

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 62


Definition 9.5 (Lie subgroup). Let G be a Lie group. A Lie subgroup H
is

• an (abstract) subgroup of G

• which is an immersed submanifold

such that H becomes a Lie group with the induced group and manifold
structures.

Remark. H might not be a submanifold.


Example. G = S 1 × S 1 is a Lie group, since S 1 = U (1), and the product of
two Lie groups is again a Lie group.
Let λ ∈ R, and let

H = {(e2πit , e2πiλt ) : t ∈ R}.

We check that H is a Lie subgroup of G.


Indeed: H is a subgroup of G. It is also an immersed submanifold (as we
saw earlier). The multiplication and inversion are smooth. (To check this:
the smooth structure on H is obtained from the one on R, and the induced
multiplication is the addition on R.) 

Proposition 9.6. Let G be a Lie group, let H be a subgroup and also a


submanifold. Then H is a Lie subgroup.

Proof. There is an obvious manifold structure on H, since it is a sub-


manifold of G. We have to check that the multiplication m : H ×H → H
and the inversion i : H → H are smooth. We just do the latter: the
inversion of G is smooth, so restricting it to a submanifold, we again get
a smooth map.
One can show:

Theorem 9.7. Let G be a Lie group, H a subgroup. If H ⊂ G is closed


in the topological sense, then H is a submanifold, and therefore a Lie
subgroup.

9.2 From Lie groups to Lie algebras


Recall:

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 63


Definition 9.8 (Lie algebra). A Lie algebra is a vector space g equipped
with a bilinear, skew symmetric map [·, ·] : g × g → g, such that

[x, [y, z]] + [z, [x, y]] + [y, [z, x]] = 0.

We now define morphisms and the “subobjects” of Lie algebras:

Definition 9.9 (Lie algebra morphism). A map φ : (g, [·, ·]g ) → (h, [·, ·]h )
is a Lie algebra morphism if it preserves all the structure:

• the map φ is linear

• φ([x, y]g ) = [φ(x), φ(y)]h ∀x, y ∈ g.

Definition 9.10 (Lie subalgebra). A Lie subalgebra of g is a vector sub-


space h, such that
[x, y] ∈ h ∀x, y ∈ h.

We now show three propositions, labelled (A), (B), (C), about


(A) Lie groups
(B) Lie group morphisms
(C) Lie subgroups.
Later we will encounter their “converses”.

Proposition 9.11 (A). Let G be a Lie group, then g := Te G is a Lie


algebra, with bracket
[v, w] := [←

v ,←
−]| .
w e

Proof. Recall the linear isomorphism Te G → X(G)L , v 7→ ← −


v . The Lie
bracket of two left invariant vector fields is again a left invariant vector
field:

(Lg )∗ [←

v ,←
−] = [(L ) ←
w − ←− ←− ← −
g ∗ v , (Lg )∗ w ] = [ v , w ],

where in the first equation we used the naturality of the Lie bracket.
So X(G)L is a Lie algebra. Now use the above linear isomorphism to
transport this Lie algebra structure to Te G.

Example. Consider GL(n, R). Then Te (GL(n, R)) is all n × n matrices, i.e.
Mat(n, R). So Mat(n, R) has an induced Lie algebra structure. Its bracket
is
[A, B] = AB − BA.

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 64


(This is not at all obvious, it requires a bit of computation) 

Proposition 9.12 (B). Let Φ : G → H be a Lie group morphism (i.e. a


smooth group homomorphism). Then

de Φ := (Φ∗ )e : Te G → Te H

is a Lie algebra morphism.

←−−−−
Proof. For all v ∈ Te G, the vector field ←

v is Φ-related to (de Φ)v. Indeed,

(dg Φ)((←

v )g ) = (dg Φ)((de Lg )(v)) = de (Φ ◦ Lg )(v) = de (LΦ(g) ◦ Φ)(v),

where the last equation holds because Φ is a group homomorphism.


Now, pulling derivatives apart again, we see that the above equals
←−−−−
(LΦ(g) )∗ ((de Φ)v) = ((de Φ)v)Φ(g) .

Take v1 , v2 ∈ Te G. Applying the previous statement and the naturality


←−−−−− ←−−−−−
of the Lie bracket we obtain that [← v−1 , ←
v−2 ] is Φ-related to [(de Φ)v1 , (de Φ)v2 ].
In particular, at g = e,
h←−−−−− ←−−−−−i
(de Φ)([←
v−1 , ←
v−2 ]e ) = (de Φ)v1 , (de Φ)v2 .
e

Proposition 9.13 (C). Let H be a Lie subgroup of G. Then Te H is a


Lie subalgebra of Te G.

Proof. We know that the inclusion is Lie group morphism. Therefore, its
derivative Te H → Te G is a Lie algebra morphism by the last proposition.

Example. SL(n, R) is a Lie subgroup of GL(n, R), hence the Lie algebra of
SL(n, R) is a Lie subalgebra of GL(n, R). It turns out that the Lie algebra
of SL(n, R) are exactly the traceless matrices. 

9.3 The exponential map


Let G be a Lie group, g = Te G.

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 65


Lemma 9.14. For all v ∈ Te G, there is a unique morphism of Lie groups
γv : (R, +) → G with the property that γv0 (0) = v.

Notice: there are lots of curves γ with velocity γ 0 (0) = v, but by the lemma
there exists only one which is also a group homomorphism, i.e. γv (s)·γv (t) =
γv (s + t).

Proof. Existence: The left invariant vector field ←



v is complete, i.e. in-
tegral curves are defined for all times. Let

γ:R→G

be the integral curve of ← −


v starting at e. Clearly γ 0 (0) = v. Further γ
is a group homomorphism, because for every fixed s, we have that the
curves t 7→ Lγ(s) γ(t) and t 7→ γ(s + t) agree. (At t = 0, both of these go
through γ(s), and they are also integral curvesa of ← −
v .)
Uniqueness: let γ : R → G be a Lie group morphism with γ 0 (0) = v.
Since ∂t

is a left-invariant vector field on R, by the proof of Proposi-
←−−−−−−− ←−−
tion 9.12 it is γ-related to (d γ)( ∂ | ) = γ 0 (0) = ←
0 ∂t 0

v . Hence for every
t0 ∈ R:
∂
=←


(dt0 γ) v |γ(t0 ) .
∂t t0

The left hand side is just γ 0 (t0 ). Hence γ is the (unique) integral curve
of ←

v starting at e.
a
For t 7→ γ(s + t), it’s trivial. For t 7→ Lγ(s) γ(t), we have that this is an integral
curve of (Lγ(s) )∗ ←

v , which is the same as ← −
v.

Definition 9.15 (Exponential map). Let G be a Lie group with Lie alge-
bra g. The exponential map is

exp : g → G, v 7→ γv (1).

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 66


Proposition 9.16. a) exp(tv) = γv (t) for all t ∈ R and v ∈ g.

b) There exists a neighborhood U of 0 ∈ g such that

exp |U : U → exp(U )

is a diffeomorphism onto an open subset of G. (This allows to


study G close to the identity element e by studying the Lie algebra.
This also defines a chart near e.)

c) If H ⊂ G is a Lie subgroup. Then expH : Te H → H is the


restriction of expG : Te G → G.

Proof. a) The curves s 7→ γtv (s) and s 7→ γv (st) agree, as they are
both Lie group morphism (R, +) → G with the same velocity tv
at s = 0. Take s = 1.

b) Idea: check that d0 exp : T0 g = g → Te G = g is the identity on g.


Then apply the inverse function theorem.

Example. For GL(n, R), we have

exp : Te GL(n, R) = Mat(n, R) −→ GL(n, R)



A
X An
A 7−→ e := .
n!
n=0

Proof. For all A ∈ Mat(n, R), the curve

(R, +) → GL(n, R), t 7→ etA

is a group morphism, since etA · esA = e(t+s)A . Furthermore,


d d
etA = (I + tA + O(t2 )) = A.
dt 0 dt 0
So this curve is γA .

9.4 From Lie algebras to Lie groups


Let (g, [·, ·]) be a (finite dimensional) Lie algebra.

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 67


Definition 9.17 (Lie group integrating a given Lie algebra). A Lie group
G integrates g iff Te G is isomorphic to g (as Lie algebras).

We will show:

Theorem 9.18 (A). Up to isomorphism, there exists a unique simply


connected Lie group GSC integrating g.

Remark. All other connected Lie groups integrating g are quotients of GSC
by discrete normal subgroups.
Example. Let g = (R, [·, ·] = 0). Then GSC = (R, +). Note that U (1) =
S 1 = R/Z is also a Lie group that integrates g. 
Example. Let g = {A ∈ Mat(n, R) : A + AT = 0}. The Lie group SO(n)
integrates g. 

Proposition 9.19 (C). Let G be a Lie group and h a Lie subalgebra of


g = Te G. There exists a unique connected Lie subgroup H whose Lie
algebra is h.

Example. Let G = U (1) × U (1), h = span(1, λ) ⊂ R2 = Te G, where λ ∈ R.


Then H = {(eit , eiλt ) : t ∈ R}. 

Proof. Sketch: Denote by D the distribution on G given by

Dg := (Lg )∗ h.

It is involutive because it is spanned by left-invariant vector fields and


h is a Lie subalgebra.
By the global Frobenius theorem, there is a (unique) foliation of G
whose leaves are tangent to D. Notice that the foliation is invariant
under left-translation: (Lg )(Sg0 ) = Sgg0 , where Sg0 denotes the leaf of D
through g 0 .
One can show that H := Se , the leaf of the foliation through e, is a
Lie subgroup with Lie algebra h. (Clearly Se is an immersed manifold
with Te Se = h. To show that it is a subgroup, use the left-invariance
of the foliation. One can show that multiplication and inversion are
smooth.) Further, one can show uniqueness.

Corollary 9.20. Given a Lie group G, there is a bijection

{Connected Lie subgroups of G} ↔ {Lie subalgebras of Te G}.

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 68


Remark. The nice bijection in this corollary justifies the (slightly involved)
definition of Lie subgroup.
Once can show:

Proposition 9.21 (B). Let G be a simply connected Lie group, H a Lie


group, and Ψ : Te G → Te H a Lie algebra morphism. Then there is a
unique Lie group morphism Φ : G → H such that de Φ = Ψ.

This proposition allows us to easily prove the uniqueness in Theorem


(A):

Proof. Let G1 , G2 be simply connected Lie groups and Ψ : Te G1 → Te G2


a Lie algebra isomorphism. Then there exist

• a Lie group morphism Φ : G1 → G2 s.t. de Φ = Ψ,

• a Lie group morphism χ : G2 → G1 s.t. de χ = Ψ−1 .

Since χ ◦ Φ and IdG1 are both Lie group morphisms with derivative
IdTe G1 , they must agree. Similarly, Φ◦χ = IdG2 . So Φ is an isomorphism.

We sketch the proof of the existence in Theorem (A), i.e.: given a Lie
algebra g, there exists a simply connected Lie group integrating it.

Proof. Idea of proof: g is isomorphic to a Lie subalgebra g0 of Mat(n, R)


for some n (Ado’s theorem).
Let G0 be the unique connected Lie subgroup of of GL(n, R) with
Lie algebra g0 . Take GSC to be the universal cover of G0 . (The universal
cover is again a Lie group, and simply connected.)

CHAPTER 9. LIE GROUPS AND LIE ALGEBRAS 69

You might also like