DG Notes Berkeley
DG Notes Berkeley
Contents
3 Differential forms 16
3.1 Exterior derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2 De Rham cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 Orientations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Partitions of unity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.5 Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.6 Stokes’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.7 Applications of Stokes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1
5.3 Homotopy invariance of de Rham cohomology . . . . . . . . . . . . . . . . . 32
8 Riemannian geometry 38
8.1 Riemannian metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
8.2 The Levi-Civita connection . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
8.3 The Riemann tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
8.4 Hodge theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2
1 Manifolds and smooth maps
1.1 Manifolds
Definition. A topological n-manifold is a topological space X such that for all p ∈ X
there exists an open neighborhood U of p, an open set V ⊆ Rn and a homeomorphism
φ : U → V . We also require that X is Hausdorff and second countable.
Remark. One can show that for spaces locally homeomorphic to Rn the condition “Haus-
dorff + second countable” is equivalent to “metrisable + has countably many components”
Maps φ as in the definition are called charts for X. A collection of charts whose domains
cover X is called an atlas for X.
If φα : Uα → Vα , φβ : Uβ → Vβ are overlapping charts, then
φβ ◦ φ−1
α : φα (Uα ∩ Uβ ) → φβ (Uα ∩ Uβ )
f ◦ φ−1
α : φα (Uα ∩ W ) → R
3
endowed with the subspace topology. We define an atlas on S n . Let U± = S n \{(0, . . . , 0, ±1)}.
≃
→ Rn by
Define φ± : U± −
1
φ± (x1 , . . . , xn+1 ) = (x1 , . . . , xn ).
1 ∓ xn+1
(φ ◦ γ1 )′ (0) = (φ ◦ γ2 )′ (0)
Proof. Given a chart φ about p, the map πpφ : {curves at p} → Rn factors through Tp X
and thus induces an injection Tp X ,−→ Rn . It is easily seen to be surjective. This bijection
defines a vector space structure on Tp X. Different charts give rise to different bijections
but they are related by a linear automorphism of Rn , hence they all induce the same vector
space structure on Tp M .
∂
Definition. Given a chart φ about p with coordinates x1 , . . . , xn , define ∂x i
∈ Tp X to be
φ −1 n
(πp ) (ei ) where ei ∈ R is the i-th standard basis vector. We will often abbreviate this
by ∂xi or ∂i .
Warning. ∂xi depends on the choice of the whole coordinate system x1 , . . . , xn , not just
xi .
4
Lemma 1.4. On chart overlaps we have
X ∂xj n
∂ ∂
=
∂yi ∂yi p ∂xj
j=1
∂x
Here ∂yji |p := ∂y∂ i xj := φj (ψ −1 (ψ(p) + tei ))′ (0) where φ, ψ are charts inducing the local
coordinates xj , yi .
Proof. We have
∂
= (πpψ )−1 (ei )
∂yi
= (πpφ )−1 (πpφ ◦ (πpψ )−1 )(ei )
= (πpφ )−1 (t 7→ φ(ψ −1 (p) + tei ))′ (0)
Xn
= (πpφ )−1 (t 7→ φj (ψ −1 (p) + tei ))′ (0)ej
j=0
n
X ∂xj
= (πpφ )−1 ej
∂yi p
j=0
n
X ∂xj ∂
= .
∂yi p ∂yj
j=0
1.3 Derivatives
Fix manifolds X, Y and a smooth map F : X → Y .
Definition. The derivative of F at p ∈ X is the map Dp F : Tp X → TF (p) Y defined by
[γ] 7→ [F ◦γ]. Sometimes we will write it as F∗ ,called pushforward by F on tangent vectors.
Lemma 1.5. The map Dp F is well-defined and linear.
Proof. Let φ, ψ be charts about p, resp. F (p). We have πFψ (p) (F ◦ γ) = (ψ ◦ F ◦ γ)′ (0) =
((ψ ◦ F ◦ φ−1 ) ◦ (φ ◦ γ))′ (0) = T πpφ (γ) where T is the derivative of ψ ◦ F ◦ φ−1 at φ(p).
Dp F
Tp X TF (p) X
πpφ ψ
πF (p)
T
Rdim X Rdim Y
5
Suppose that {xi }, {yj } are coordinates associated to φ resp. ψ. Then ψ◦F ◦φ−1 expresses
∂y
the yj as functions of the xi via F , so T = ∂xji p and
∂y
j
Dp F (∂xi ) = DF ((πpφ )−1 (ei )) = (πFψ (p) )−1 (T ei ) =
X ψ
(πF (p) )−1 ej .
∂xi p
j
Remarks.
(i) For X = Rn , Y = Rm , the new notion of derivative coincides with the standard one
from multivariable calculus.
∂f
(ii) Given f : X → R, we have Dp f (∂xi ) = ∂xi p .
Then Dp (G ◦ F ) = DF (p) G ◦ Dp F .
Proof. By definition.
6
Notice: We could have chosen ψ ◦ F as φ. W.r.t. the charts ψ ◦ F, ψ the map F looks like
the identity in local coordinates.
Proposition 1.8. Suppose F is an immersion at p. Given local coordinates x1 , . . . , xn
on X about p, there exist local coordinates y1 , . . . , ym on Y about F (p) w.r.t. which F
looks like Rn = Rn × 0 ,−
→ Rn × Rm−n = Rm . Similarly, if F is a submersion of p, then
given coordinates y about F (p) there exist coordinates x about p w.r.t. which F looks like
Rn → Rn /(0 × Rn−m ) = Rm .
Proof. Exercise.
1.5 Submanifolds
Fix an n-manifold X.
Definition. A subset Z ⊆ X is a submanifold of codimension k if for all p ∈ Z there exist
local coordinates x1 , . . . , xn on X about p such that Z is given locally by x1 = · · · = xk = 0.
(Formally: on the domain of the chart Z = {x1 = · · · = xk = 0}).
Z is a properly embedded submanifold if the same holds for all p ∈ X (not merely p ∈ Z).
E.g. 0 × R ⊆ R2 is a properly embedded submanifold. 0 × R∗ ⊆ R2 is a submanifold but
not properly embedded.
Given a codimension k submanifold Z ⊆ X,
• Equip Z with the subspace topology (automatically Hausdorff and 2nd-countable
because X is)
• For p ∈ Z can choose local coordinates x1 , . . . , xn on X as in the definition. Then
xk+1 , . . . , xn define a chart on Z about p.
• Transition functions are smooth since original transition functions on X were smooth.
Proposition 1.9. A codimension k submanifold Z ⊆ X is naturally an (n − k)-manifold.
The inclusion map ι : Z → X is a smooth immersion and a homeomorphism onto its
image. Composition with ι gives a bijection
ι◦−
{smooth maps to Z} −−→ {smooth maps to X
with image ⊆ Z }.
7
Example. The inclusion S n ,−→ Rn+1 is an embedding. So S n is a submanifold of Rn+1
and the induced manifold structure agrees with the one we already defined.
Proposition 1.11. If F : X → Y is a smooth map, and q ∈ Y is a regular value, then
F −1 (q) is a submanifold of X of codimension dim Y .
Tp Y + Tp Z = Tp X.
8
2 Vector bundles and tensors
9
• Total space: T X = p∈X Tp X. This is a manifold via a pseudo-atlas: Given
F
a
F coordinate patch Un on X with coordinates
P xi we get a pseudo-chart φ :
T
p∈U p X → U × R given by (p, a ∂
i xi ) →
7 (p, (a1 , . . . , an )). These make
T X into a manifold.
• Projection π : (p, v) 7→ p.
• The pseudo-charts give the local trivializations.
Definition. A section of a vector bundle π : E → B is a smooth map s : B → E such
that π ◦ s = idB .
Examples.
(i) Every vector bundle has a zero section given by s(b) = 0 for all b ∈ B.
(ii) A vector field on X is a section of the tangent bundle T X
Definition. Given vector bundles πi : Ei → Bi (i = 1, 2) and a smooth map F : B1 → B2 ,
a morphism of vector bundles E1 → E2 covering F is a smooth map G : E1 → E2 such
that π2 ◦ G = F ◦ π1 and for all p ∈ B1 the restricted map Gb : (E1 )p → (E2 )F (p) is linear.
If B1 = B2 = B, an isomorphism of vector bundles over B is a morphism covering idB
with a two-sided inverse. Equivalently, a diffeomorphism of the total space that induces
linear isomorphisms (E1 )p → (E2 )p .
Example. Consider S 1 = {eiθ | θ ∈ R} ⊆ C. The vector field ∂θ is defined and non-zero
in each fibre. So we get an isomorphism
S1 × R → T S1
(eiθ , a) 7→ (eiθ , a∂θ )
10
2.2 Vector bundles by gluing
To define a vector bundle over B (of rank k) it suffices to give an open cover {Uα } of B
and for all α, β a smooth map gβα : Uα ∩ Uβ → GLk (R) satisfying
(i) gαα = constant map with value idRk .
(ii) gγα = gγβ gβα for all α, β, γ (cocycle condition)
−1
Note that from (i) and (ii) it follows that gαβ = gβα .
Given this data, define
× Rk (|{z}
`
E= α Uα b , v) ∼ (|{z}
b , gβα (b)v)
∈Uα ∈Uα
Corollary 2.2. Two bundles are isomorphic iff they can be trivialized over a common
open cover with the same transition functions.
Proof. If both can be trivialized over {Uα } with transition functions gβα , then they are
both isomorphic to the above construction. Converse is clear.
Example. Define
the Möbius line bundle M → RP1 to be trivialized over U0 , U1 with
g10 = sign xx01 .
Claim: This is isomorphic to ORP1 (−1). Suffices to show we can modify the trivializations
of M to make the transition function become xx01 instead of sign xx01 . Let’s rescale the
trivialization Φi of M by a smooth map ψi : Ui → R∗ = GL1 (R). Excplicitly, consider the
trivialization
(p, v) 7→ (p, ψi (p) pr2 (Φi (p, v)))
11
This changes g10 to ψ1
g10 . Left to choose ψ0 , ψ1 such that ψ1 /ψ0 = |x1 |/|x0 |. One choice
rψ0 r
x21 x20
that works is ψ1 = x2 +x2 , ψ0 = x2 +x 2.
0 1 0 1
Proof. Define the map e : {functions about p} → Tp∗ X, f 7→ ([γ] 7→ (f ◦ γ)′ (0)) = Dp f . In
P P ∂f
local coordinates: f 7→ ( ai ∂xi 7→ ai ∂xi ). We see that e(xj ) are the dual basis to ∂xj ,
so e is surjective and by definition e(f1 ) = e(f2 ) iff f1 , f2 agree to first order about p.
12
Definition. A section of T ∗ X is a 1-form.
Unwarning! dxi only depends on xi , not the other xj (unlike ∂xi ).
Definition. Given a smooth map F : X → Y . The map (Dp F )∨ : TF∗ (p) Y → Tp∗ X is
called pullback by F , denoted F ∗ .
Lemma 2.5. Given F : X → Y and a smooth function g on Y , we have F ∗ dg =
d(F ∗ g) := d(g ◦ F ).
Proof. Given [γ] ∈ Tp X, we have (F ∗ dg)[γ] = dg(Dp F [γ]) = dg([F ◦ γ]) = (g ◦ F ◦ γ)′ (0) =
d(g ◦ F )[γ].
13
2.6 Index notation
From now on, indices on local coordinates will be superscripts: x1 , . . . , xn .
A section T of T X ⊗ T ∗ X ⊗ T X (a specific kind of tensor of type (2, 1)) can be written
in local coordinates xi uniquely as
X ik
T = T j ∂i ⊗ dxj ⊗ ∂k
i,j,k
T = T ij k ∂i ⊗ dxj ⊗ ∂k
14
of type (0, r) and contracting with v(1) , . . . , v(r) .
1
Warning. Some people include the factor r! .
We can refer to the components of this tensor as αi1 ...ir (this is the coefficient of dxi1 ⊗
· · · ⊗ dxir ). When the ij form a multi-index I, i.e. i1 < i2 · · · < ir , this agrees with αI .
Example. On R2 we view dx1 ∧ dx2 as dx1 ⊗ dx2 − dx2 ⊗ dx1 . A general 2-form looks
like α12 dx1 ∧ dx2 = αij dxi ⊗ dxj where α21 = −α12 and α11 = α22 = 0.
In summation convention, it is correct to say α = αi1 ...ir dxi1 ⊗ · · · ⊗ dxir but NOT
α = αi1 ...ir dxi1 ∧ · · · ∧ dxir (the last sum would be r!α).
If α = α1 ∧ · · · ∧ αr , then α(v) = det(αi (v(j) )).
where we apply DF −1 (q) F on the T X factor and (Dq (F −1 ))∨ on the T ∗ X factor.
In this setting F∗ = (F −1 )∗ and vice versa.
15
3 Differential forms
∂αi j i i
Definition. The exterior derivative dα is ∂x j dx ∧ dx = dαi ∧ dx . By the above calcu-
lation this is well-defined (independent of local coordinates).
∂αI
More general, given a p-form α = αI dxI we define dα := dαI ∧dxI = ∂xj
dxj ∧dxI . Again,
this is well-defined.
Proposition 3.1. d satisfies the following:
(i) It is R-linear.
(ii) It agrees with the differential on 0-forms.
(iii) d2 = 0.
(iv) It commutes with pullback, i.e. F ∗ (dα) = d(F ∗ α).
(v) Graded Leibniz rule: Given a p-form and a q-form β:
∂ 2 αI
2 ∂αI j I
d α=d dx ∧ dx = dxk ∧ dxj ∧ dxI = 0
∂xj ∂xk ∂xj
16
∂ 2 αI
since ∂xk ∂xj
is symmetric in j, k, but dxk ∧ dxj is antisymmetric.
(iv) Write α locally as αI dy I . Then
F ∗ (dα) = F ∗ (dαI ∧ dy i1 ∧ · · · ∧ dy ip )
= F ∗ (dαI ) ∧ F ∗ (dy i1 ) ∧ · · · ∧ F ∗ (dy ip )
= d((F ∗ αI )d(F ∗ y i1 ) ∧ · · · ∧ d(F ∗ y ip )
= d(F ∗ α)
using (iii) and (v) (we sum over multi-indices i1 < · · · < ip ).
(v) Let α = αI dxI , β = βJ dxJ . Then
17
Example. Let X = S 1 . We know that
R
r = 0,
r 1
HdR (S ) = ? r = 1,
0 r ̸= 0, 1.
A 1-form on S 1 can be written uniquely as f (θ)dθ. All 1-forms are closed. Define a map
I : Ω1 (S 1 ) −→ R,
Z 2π
f (θ)dθ 7−→ f (θ)dθ
0
This is linear and non-zero, hence surjective. Claim: ker I = B 1 (S 1 ). Proof: If f dθ = dg,
then f = ∂g ∂θ , so I(f dθ) = g(2π) − g(0) = 0. Conversely, if I(f dθ) = 0, define g(θ) =
Rθ 1 1
0 f (t)dt. Then we have f dθ = dg. Thus we get an isomorphism I : HdR (S ) ≃ R.
18
3.3 Orientations
Vn
Definition. An orientation of a n-dimensional vector space is a non-zero element of V
modulo positive rescalings.
An orientation of a vector bundle E → X of rank k is a nowhere-zero section of k E,
V
modulo rescaling by positive smooth functions. E is orientable if there exists an orientation
for it, and oriented if it is equipped with a choice of orientation.
Note: E is orientable iff k E is trivial. E.g. any trivial bundle is orientable. The tauto-
V
logical bundle over RPn is not orientable.
Definition. A manifold X is orientable/oriented if T X → X is.
E.g. S n is orientable for all n. RPn is orientable iff n is odd.
Definition. A volume form on n-manifold X is a nowhere-zero n-form.
A volume form ω defines an orientation (basis e1 , . . . , en for Tp X is positively oriented iff
ω(e1 , . . . , en ) > 0) and conversely an orientation defines a volume form modulo rescaling
by positive smooth functions.
Lemma 3.6. For any open cover {Uα } of X, there exists a partition of unity subordinate
to it.
3.5 Integration
Fix an oriented n-manifold X and a compactly supported n-form ω on X.
R
Definition. The integral of ω over X, denoted X ω, is defined as follows:
• Cover X by coordinate patches Uα with coordinates xiα . WLOG these are positively
oriented, i.e. ∂x1α ∧ · · · ∧ ∂xnα represents the orientation.
• Pick a Rsubordinate
P Rpartition1 of unity {ρα }. Write ρα ω = f/ alphadx1α ∧ · · · ∧ dxnα .
Define X ω = α Rn fα dxα . . . dxnα .
19
R
Lemma 3.7. The integral X ω is well-defined.
Proof. Suppose we cover X by patches Vβ with coords yβ1 , . . . , yβn . Take a partition of
unity σβ subordinate to this cover. Locally write σβ ω = gβ dyβ1 ∧ · · · ∧ dyβn . We want to
show α Rn fα dx1α . . . dxnα = β Rn gβ dyβ1 . . . dyβn . On overlaps Uα ∩ Vβ we have
P R P R
∂y j
So σβ fα = ρα gβ det( ∂xβi ). Since both y and x are oriented in the same way, we have
α
∂y j ∂y j
det( ∂xβi ) = | det( ∂xβi )|. So
α α
XZ XZ
fα dx1α . . . dxnα = σβ fα dx1α . . . dxnα
α Rn α,β Rn
XZ ∂yβj
= ρα gβ det( )dx1α . . . dxnα
Rn ∂xiα
α,β
XZ
= ρα gβ dyβ1 . . . dyβn
α,β Rn
XZ
= ρα gβ dyβ1 . . . dyβn
β Rn
Note: Since ω is compactly supported and partitions of unity are locally finite, all sums
appearing are actually finite.
20
(iii) If X is a m-w-b and Y is an ordinary manifold, then X × Y is a manifold with
boundary. Then X × Y is a manifold-with-boundary. ∂(X × Y ) = (∂X) × Y and
(X × Y )◦ = X ◦ × Y .
If both X and Y are manifolds-with-boundary, then in general X × Y is a manifold-
with-corners.
Definition. If X is an oriented
V n-manifold with boundary, then ∂X is oriented as follows.
Given p ∈ ∂X, pick oX ∈ n Tp X representing the orientation of X. Pick a vector
n ∈ Tp X transverse
Vn−1 to ∂X and pointing outwards. Orient ∂X at p by the unique o∂X ∈
Tp ∂X ≤ n−1 Tp X satisfying oX = n ∧ o∂X .
V
Proof. Step 1: Cover X by coordinate patches and pick a subordinate partition of unity
ρα . Then Z Z X XZ XZ
ω= ρα ω = ρα ω = ρα ω
∂X ∂X α α ∂X α ∂Uα
and Z Z X XZ
dω = d( ρα ω) = d(ρα ω).
X X α α Uα
So it suffices to prove the result when X is a coordinate patch. WLOG X = R≥0 × Rn−1 .
Orientation of ∂(R≥0 ×Rn−1 ) is −∂x2 ∧· · ·∧∂xn . So the right side above is exactly this.
Historical fact: Stokes put this theorem as a problem in the Cambridge exam (basically
Part III).
Example. X = {x ∈ R2 : ∥x∥ ≤ a}. Area of X =
Z Z
1
dx ∧ dy = d(xdy − ydx)
X X 2
21
Z
1
= xdy − ydx
2 ∂X
Z
1
= r2 dθ
2 ∂X
a2
Z
= dθ
2 ∂X
= πa2
Proof. By the Leibniz rule d(α ∧ β) = (dα) ∧ β + (−1)p−1 α ∧ dβ. Integrating and applying
Stokes gives Z Z Z
α∧β = (dα) ∧ β + (−1)p−1 α ∧ dβ
∂X X X
n (X) ̸= 0.
Corollary 3.11. If X is a compact, orientable n-manifold, then HdR
22
4 Connections on vector bundles
4.1 Connections
Fix a rank k vector bundle π : E → B. Given a section s, we can view it locally under each
trivialization Φα as an Rk -valued function which we will denote by vα (= pr2 ◦Φα ◦ s|Uα ).
The naive derivative is dvα , which we can view as a local E-valued 1-form via Φ−1
α . Under
a different trivialization Φβ , s becomes vβ = gβα vα . Taking the naive derivative and
transferring the answer to the α-trivialization gives
−1 −1
gβα d(gβα vα ) = gβα (dgβα )vα + dvα
So the answer is trivialization-dependent via the action of the gl(k, R)-valued 1-form
−1
gβα dgβα on vα .
Definition. A connection A on E comprises a gl(k, R)-valued 1-form Aα on Uα for each
∼
trivialization Φα : π −1 (Uα ) −
→ Uα × Rk such that on overlaps we have
−1 −1
Aα = gβα Aβ gβα + gβα dgβα (∗)
23
Example (Trivial connection). Suppose E → B admits a global trivialization Φα . We
can define a connection A by Aα = 0, then defining Aη for all other trivializations by (∗).
A section is horizontal iff it is locally constant under Φα .
Lemma 4.1. Given a connection A on E → B, the covariant derivative
dA : Γ(E) → Ω1 (E)
It is clearly R-linear and inherits the Leibniz rule from dA0 . So it corresponds to a unique
connection on T X.
Lemma 4.2. Any vector bundle admits a connection.
Proof. Trivialize E over Uα with transition functions gβα as usual. Pick a partition of
unity ρα subordinate to this cover. Now define
X
−1
Aα = ργ gγα dgγα
γ
It remains to prove that this satisfies the transformation law (∗). We have
X
−1 −1
gβα Aβ gβα = ργ gβα (gγβ dgγβ )gβα
γ
X
−1
= ργ gγα (d(gγβ gβα ) − gγβ dgβα )
γ
X X
−1 −1
= ργ gγα dgγα − ργ gβα dgβα
γ γ
−1
= Aα − gβα dgβα
24
4.2 Connections vs End(E)
Fix rank k vector bundle E → B. Let ρ : GL(k, R) → GL(gl(k, R)) be the representation
ρ(A)(M ) = AM A−1 for A ∈ GL(k, R) and M ∈ gl(k, R).
Definition. End(E) is the vector bundle over B of rank k 2 with total space
a
End(Eb )
b∈B
If E is trivialized over Uα with transition functions gβα , then End(E) is trivialized over
the same sets with transition functions ρ(gβα ).
−1
A section M of End(E) is locally a gl(k, R)-valued function Mα such that Mβ = gβα Mα gβα .
Equivalently End(E) = E ⊗ E . ∨
Proof. Just prove that everything is compatible with the transition functions.
−1 −1 −1 −1 −1
Aα + ∆α = gβα Aβ gβα + gβα dgβα + gαβ ∆β gαβ = gβα (Aβ + ∆β )gβα + gβα dgβα
For the other direction, verify that A′ − A transforms correctly, i.e. like a section of
Ω1 (End(E)).
dA (s ⊗ ω) = (dA s) ∧ ω + s ⊗ dω
25
= Fα ∧ σα
where Fα = dAα +Aα ∧Aα . Then one can check that this transforms like a End(E)-valued
2-form.
26
Definition. Given a vector v0 ∈ Eγ(0) , the unique horizontal section s of γ ∗ E starting at
v0 is the horizontal lift of γ to E (starting at v0 ). The vector s(1) ∈ Eγ(1) is the parallel
transport of v0 along γ. Doing this for all v0 gives a linear map Pγ : Eγ(0) → Eγ(1) . If γ
is a loop, i.e. γ(0) = γ(1), then Pγ : Eγ(0) → Eγ(0) is the monodromy or holonomy of A
around γ.
Examples.
(i) Consider T S 2 with the “orthogonal projection” connection. Given path γ on S 2 and
v0 ∈ Tγ(0) S 2 , the horizontal lift is the map v : [0, 1] → T S 2 such that
• v(t) ∈ Tγ(t) S 2 for all t.
• v̇(t) in R3 is orthogonal to Tγ(t) S 2 , so that the orthogonal projection to Tγ(t) S 2
is 0
2 1 1 0 0 −1
(ii) Returning to R → R × S with connection Aα = f dx + g dθ.
0 −1 1 0
1 0
Consider γ(t) = (t, 0). Horizontal lift v of γ starting at v0 satisfies v̇+f v=
0 −1
−λ
e 0 Rt
0. So v(t) = λ v0 where λ = 0 f (x, 0)dx. Similarly, the monodromy
0 e
cos φ sin φ R 2π
around γ(t) = (0, 2πt) is where φ = 0 g(0, θ)dθ.
− sin φ cos φ
∂ 2 Pa,b
|a=b=0 = −Fij (p)
∂a∂b
Proof. For formal proof see Sheet 3. We will give an intuitive sketch proof ignoring analysis
details, but these details can be filled in (e.g. see Nicolaescu Proposition 3.3.14).
Parallel transport in the xi direction satisfies v̇ = −Ai v, so Pγ1 = I −aAi (p)+. . . where . . .
means higher order terms which will wash out. Similarly Pγ2 = I − bAj (γ1 (1)) + · · · = I −
∂A
b(Aj (p)+a ∂xij (p))+. . . . So Pγ2 ◦Pγ1 = I −aAi (p)−bAj (p)+abAj (p)Ai (p)−ab ∂A ∂xi
i
(p)+. . . .
27
∂bi
Similarly Pγ4 ◦ Pγ3 = I + aAi (p) + bAj (p) + ab ∂x j
(p) + abAj Ai (p) + . . . . So
∂Ai ∂Aj
Pa,b = I + ab (p) − (p) + Aj (p)Ai (p) − Ai (p)Aj (p) + a(. . . ) + b(. . . ) + . . .
∂xj ∂xi
So
∂ 2 Pa,b ∂Ai ∂Aj
|a=b=0 = j
(p) − (p) + Aj (p)Ai (p) − Ai (p)Aj (p) = −Fij (p)
∂a∂b ∂x ∂xi
V2
Corollary 4.7. If v ∈ Ep is such that F (p)v ̸= 0 in Ep ⊗ Tp∗ B, then there does no
exist a local horizontal section s about p with s(p) = v.
Proof. If such an s exists, then Pγ1 (v) = s(γ1 (1)), similarly for the other paths, so Pa,b (v) =
s(p) = v for all a, b. So by the Proposition for all i, j we have −Fij (p)v = 0, hence
F (p)v = 0.
Example. Consider R → S 1 with Aα = Cdθ. Local horizontal sections exist and have
the form vα = Ke−cθ . This extends to a global section if C = 0. If C ̸= 0, then this does
not extend, due to the presence of non-trivial monodromy e−2πC .
Summary. Curvature is the local obstruction from the existence of horizontal sections,
monodromy is the global obstruction.
28
5 Flows and Lie derivatives
5.1 Flows
Fix a manifold X and a vector field v on X. Given a point p ∈ X, can try to flow along
v from p, i.e. solve the ODE γ̇(t) = v(γ(t)) and γ(0) = p. By standard ODE theory,
solutions exist locally and are unique. Solutions are called integral curves of v.
Definition (Non-standard). A flow domain is an open neighborhood U of X × 0 in X × R
such that for all p ∈ X the set U ∩ (p × R) is connected.
Definition. A local flow of v comprises a flow domain U and a smooth map Φ : U → X
such that
• Φ(−, 0) = idX .
• d
dt Φ(p, t) = v(Φ(p, t)) for all (p, t) ∈ U .
We write Φt (p) for Φ(p, t).
Previous ODE discussion plus smooth dependence on initial conditions, tells us that ocal
flows exist and are unique in the sense that if (U1 , Φ1 ) and (U2 , Φ2 ) are local flows, then
Φ1 = Φ2 on U1 ∩ U2 .
A vector field is complete if it has a global flow, i.e. one with U = X × R. Not all vector
fields are complete, e.g. x2 ∂x on R. But if v is compactly supported, then v is complete.
(Idea: for each p ∈ X, there exists Up neighborhood of p and εp > 0 such that a flow exists
on Up × (−εp , εp ). By compactness, get local flow on U = X × (−ε, ε) for some ε > 0. Can
then define a global flow by Φt = (Φt/N )N for n ≫ 0.)
Lemma 5.1. If Φ is a local flow of v, then Φs+t = Φs ◦ Φt whenever this makes sense.
So in particular (Φt/N )N = Φt when this makes sense, and Φ−t = (Φt )−1 .
Proof. Fix p ∈ X, fix t. Let q ∈ Φt (p). Then γ1 (s) := Φs+t (p), γ2 (s) = Φs ◦ Φt (p). These
two curves both satisfy γ̇i = v ◦ γi and γi (0) = q. So by uniqueness of solutions to ODEs
get γ1 = γ2 .
29
Definition. For a tensor or form T on X, its Lie derivative is
d
Lv T = |t=0 (Φt )∗ T
dt
Lv f = df (v).
d t d
Proof. At each point p we have Lv f = dt |t=0 f (Φ (p)) = df ( dt |t=0 Φt (p)) = df (v).
We have
d
Lv α = |t=0 (Φt )∗ α
dt
d
= |t=0 (αi ◦ Φt )d(xi ◦ Φt )
dt
= (Lv αi )dxi + αi d(Lv xi )
∂αi
= v j j dxi + αi dv i
∂x
∂αi ∂v j
= (v j j + αj i )dxi
∂x ∂x
Lv (S ⊗ T ) = (Lv S) ⊗ T + S ⊗ (Lv T )
30
Proof. Pullback by Φt commutes with contraction and with ⊗. Then proceed as in the
proof of the ordinary Leibniz rule.
Proof. By first part of the previous lemma, for any 1-form α we have Lv (αi wi ) = (Lv α)i wi +
αi (Lv w)i , so we get by the lemma before
i j
j ∂(αi w ) j ∂αi j ∂v
v = v +α wi + αi (Lv w)i
∂xj ∂xj ∂xi
Hence
∂αi ∂wi j i ∂αi i ∂v
j
v j wi + v j
α i = v w + α j w + αi (Lv w)i
∂xj ∂xj ∂xj ∂xi
and thus
∂wi i ∂v
j
αi (Lv w)i = v j αi − α j w .
∂xj ∂xi
This holds for all α, so
∂wi j ∂v
i
(Lv w)i = v j − w .
∂xj ∂xj
[v, w] := Lv w = −Lw v.
This operation makes the Γ(T X) into a Lie algebra, i.e. a vector space equipped with a
bilinear operation [·, ·] satisfying
• [v, v] = 0 for all v (alternating)
• [u, [v, w]] + [v, [w, u]] + [w, [u, v]] = 0 for all u, v, w (Jacobi identity)
Lemma 5.6. If F : X → Y is a diffeomorphism, then for any vector field v on Y , and
any tensor T on Y , we have
F ∗ (Lv T ) = LF ∗ v (F ∗ T )
Proof. We have
d
F ∗ (Lv T ) = F ∗ |t=0 (Φt )∗ T
dt
d
= |t=0 F ∗ (Φt )∗ T
dt
31
d
= |t=0 F ∗ (Φt )∗ (F ∗ )−1 F ∗ T
dt
d
= |t=0 (F −1 ◦ Φt ◦ F )∗ F ∗ T
dt
But F −1 ◦ Φt ◦ F is a flow of F ∗ v.
Lv α = (dιv α) + ιv dα
Lv (F ∗ α) = d(ι∂t F ∗ α) + 0
So
Z 1
F1∗ α − F0∗ α = i∗0 (Φt )∗ d(ι∂t F ∗ α)dt
0
Z 1
= i∗t d(ι∂t F ∗ α)dt
0
32
Z 1
=d i∗t ι∂t F ∗ αdt
0
So F1∗ α − F0∗ α is exact and thus F0 , F1 induce the same map on de Rham cohomology.
33
6 Foliation and Frobenius integrability
6.1 Foliations
If F : X → Y is a submersion, then X decomposes into slices F −1 (q) which are submani-
folds of dimension dim X − dim Y .
A k-foliation on X is a local decomposition of X into k-dimensional slices, but the slices
need not globally form submanifolds.
Example. Consider X = T 2 = R2 /Z2 . For any α ∈ R, we can locally slice X into lines
of slope α. If α is irrational, then the slices do not globally form submanifolds.
Definition. An atlas on X is k-foliated if the transition functions φβ ◦ φ−1 α are locally
of the form Rk × Rn−k ∋ (x, y) 7→ (ζ(x, y), η(y)) ∈ Rk × Rn−k . Two k-foliated atlases are
equivalent if their union is k-foliated, and a k-foliation is an equivalence class of k-foliated
atlases. We will usually write associated local coordinates as x1 , . . . , xk , y 1 , . . . , y n−k . Slices
are given locally by y = const.
Example. If F : X → Y is a submersion, then foliated charts correspond to local
coordinates on X in which F corresponds to projection onto the last n − k components.
6.2 Distributions
Fix an n-manifold X.
Definition. A k-plane distribution on X is a rank k subbundle D of T X.
Example. ⟨∂x , ∂y ⟩ and ⟨∂x + y∂z , ∂y ⟩ each define a 2-plane distribution on R3 . Note that
⟨∂x , ∂y ⟩ = ker dz and ⟨∂x + y∂z, ∂y ⟩ = ker(dz − ydx).
In general a k-plane distribution can be written locally as the kernel as the kernel of n − k
1-forms.
Example. If X is equipped with a k-foliation, with foliated coordinates x, y as usual, then
Tn−k
⟨∂x1 , . . . , ∂xk ⟩ = i=1 ker dy i is a k-plane distribution. It describes the tangent spaces to
the slices.
Definition. A k-plane distribution D is integrable if it arises from a k-foliation in this
way
34
If k = 1, then every distribution is integrable: Locally D = ⟨v⟩ for some vector field v,
then X is foliated by integral curves of v.
Proof. Both conditions are local, so it suffices to work near a point p. If D is integrable
with local foliated coordinates x, y, then D = ⟨∂x1 , . . . , ∂xk ⟩. Can easily check by hand
that for any smooth coefficients f i , g i , [f i ∂xi , g j ∂xj ] ∈ D.
Conversely, suppose D is closed under [·, ·]. We want to show that there exist local coordi-
nates x, y such that D = ⟨∂x1 , . . . , ∂xk ⟩. First choose local coordinates s1 , . . . , sk , t1 , . . . , tn−k
about p such that D = ⟨∂s1 , . . . , ∂sk ⟩ at p. WLOG p corresponds to s = 0, t = P0. Locally
there exist uniquely determined smooth functions aij such that vi := ∂si + aij ∂tj lies
in D. Let Φti be the flow of vi . Now define F : open neighborhood of 0 in Rn → X by
1 k
F (x, y) = Φx1 ◦ · · · ◦ Φxk (s = 0, t = y). This has F (0) = p and D0 F (∂xi ) = vi = ∂si and
D0 F (∂yi ) = ∂tj . So D0 F is an isomorphism, hence F defines a parametrization near p. It
suffices to show that ∂xi = vi . Suppose the Φi all commute, then for each i we would have
d 1 i k
∂xi = |h=0 Φx1 · · · Φxi +h · · · Φxk (0, y)
dh
d i 1
ci . . . Φxk (0, y)
= |h=0 Φxi +h Φx1 · · · Φ k
dh
i 1 k
= vi (Φxi Φx1 · · · Φ
ci . . . Φx (0, y))
k
= vi
so we would be done.
i j j i
Left to show: Φxi ◦Φxj = Φxj ◦Φxi for all i, j, i.e. that [vi , vj ] =
P0 for all i, j. We know that
D is closed under [·, ·], so there exist bijl such that [vi , vj ] = l bijl vl . Equate coefficients
of ∂sl to get all bijl = 0.
Example. Consider D = ⟨∂x + y∂z , ∂y ⟩ on R3 . This is not closed under [, ] since [∂x +
y∂z , ∂y ] = −∂z ∈
/ D. So D is not integrable.
∂f ∂f ∂f
By hand: If D were tangent to a surface f = const, then we would have ∂x +y ∂z = ∂y = 0.
∂2f ∂2f
So 0 = ∂x∂y + ∂f∂z + y ∂y∂z =
∂f
∂z . Then also ∂f
∂x = −y ∂f
∂z = 0 and so df = 0. So {f = const}
is not a surface!
35
7 Connections on vector bundles with extra
structure
7.1 Connections on T X
Suppose A is a connection on E = T X → X. Given local coordinates x1 , . . . , xn on X, we
get a trivialization of E by ∂x1 , . . . , ∂xn . Call this a coordinate trivialization. We typically
write the induced local connection 1-form as Γijk dxk where i, j are the matrix indices on
gl(n, R). So for a vector field v we have (dA v)i = dv i + Γijk v j dxk .
Warning. The Γijk do not transform like a tensor of type (1, 2). But the space of
connections on E is an affine space for Ω1 (End E) = Γ(E ⊗ E ∨ ⊗ T ∗ X) = Γ(T X ⊗ T ∗ X ⊗
T ∗ X), i.e. the space of tensors of type (1, 2).
Definition. The solder form θ is the E-valued 1-form that corresponds to the fibrewise
identity map under E ⊗ T ∗ X = T X ⊗ T ∗ X = End(T X).
The torsion T of A is the E-valued 2-form dA θ. A is torsion-free if T = 0.
In a coordinate trivialization θ = ei ⊗ dxi , so T = d(ei ⊗ dxi ) + Aα ∧ (ei ⊗ dxi ) =
Γj ik ej ⊗ dxk ∧ dxi .
γ̈ i + Γijk γ̇ j γ̇ k = 0.
Note that
• A connection on T X induces connections on T ∗ X and all bundles of tensors and
forms. If we had taken the covariant derivative of θ as a tensor of type (1, 1), we
would have got 0 automatically.
36
• The curvature of A is an End(E)-valued 2-form, which we can view as a tensor of
type (1, 3) F ijkl that is antisymmetric in k, l.
Definition. An orthogonal vector bundle is a vector bundle equipped with an inner prod-
uct g. A trivialization Φα is orthogonal if under Φα , g becomes the standard inner product
on Rk .
Note: Transition functions between orthogonal trivializations take values in O(k).
Fix an orthogonal vector bundle (E, g) → B.
Lemma 7.3. E can be covered by orthogonal trivializations.
37
8 Riemannian geometry
A section T ij of End(T X) lies in o(T X) iff T ij gik = −T ik gji , i.e. Tkj = −Tjk . When
dxi ⊗dxj +dxj ⊗dxi
writing coordinate expressions, we use dxi dxj to mean 2 , e.g. the standard
Riemannian metric on Rn is gEucl = i (dxi )2
P
Proof. We will prove the more generally statement that the map {orthogonal connections} →
Ω2 (T X) sending a connection to its torsion, is a bijection.
Fix an arbitrary orthogonal connection A0 . Any other orthogonal connection A can be
written uniquely as A0 + ∆ for an o(E)-valued 1-form ∆. We will show that the map
Ω1 (o(E)) → Ω2 (T X), ∆ 7→ TA0 +∆ − TA0 is a bijection. This map sends ∆ to ∆ ∧ θ, i.e
∆ikj − ∆ijk . (If A0 is locally Γijk , then A0 + ∆ is Γijk + ∆ijk , so (TA0 +∆ − TA0 )ijk =
(Γ + ∆)ikj − (Γ + ∆)ijk − (Γikj − Γijk )).
It is induced by the bundle morphism F : o(T X)⊗T ∗ X → T X ⊗∧2 T ∗ X given by wedging
with θ. So it suffices to show that F is an isomorphism which we can do fibrewise.
n ∗
Note both bundles have rank n 2 since o(T X) ⊗ T X = {∆ijk | ∆ijk = −∆jik } and
38
T X ⊗ ∧2 T ∗ X = {T ijk : T kjk = −T ikj }. So it is enough to show that ∆ 7→ ∆ ∧ θ is injective,
i.e. that if ∆ijk satisfies ∆ijk = −∆jik and ∆ijk = ∆ikj (∆ ∈ ker), then ∆ = 0. But if
∆ satisfies these two conditions, then ∆ijk = −∆jik = −∆jki = ∆kji = ∆kij = −∆ikj =
−∆ijk .
Definition. This is the Levi-Civita connection on (X, g). Its components Γijk are called
Christoffel symbols.
The explicit coordinate expressions are
1
Γijk = (∂j gik + ∂k gji − gi gjk ) .
2
39
In particular, there is a distinguished unit volume form ω.
Given a p-form β there exists a unique (n − p)-form ∗β such that for all p-forms α
α ∧ ∗β = ⟨α, β⟩ω
E.g. ∗αI = ±αJ where J = {1, . . . , n} \ I.
Definition. The map
∗ : Ωp (X) → Ωn−p (X)
is the Hodge star operator.
By considering its action on the αI , can see that it is a fibrewise isometry and ∗2 =
(−1)p(n−p) idΩp (X) .
Example. Take R3 with the standard metric and orientation. Then ω = dx1 ∧ dx2 ∧ dx3 ,
so ∗dx1 = dx2 ∧ dx3 , ∗(dx2 ∧ dx3 ) = dx1 and cyclically.
assume X is compact. Define an inner product on Ωp (X) by ⟨α, β⟩X = X ⟨α, β⟩ω =
R
Now
R
X α ∧ ∗β. For (p − 1)-form α, p-form β we have
Z
⟨dα, β⟩X = (dα) ∧ ∗β
Z X
40
Idea: Hp (X) = ker ∆ = ker d ∩ ker δ = ker d ∩ (im d)⊥ ∼ p
= ker d/ im d = HdR (X)
Theorem 8.5 (Hodge decomposition). The space Hp (X) is finite-dimensional and we
have orthogonal decompositions
41
9 Lie groups and principal bundles
42
Proof. Given left-invariant vector fields v, w, we have for all g ∈ G that
Definition. The Lie bracket on g is defined by [ξ, η] = ζ where ζ is the unique element
of g such that [ℓξ , ℓη ] = ℓζ . It inherits alternating, bilinear, Jacobi from the Lie bracket of
vector fields.
43
• Sections s over U ⊆ B correspond to trivializations Φ over U :
– Given Φ, define s by s(b) = Φ−1 (b, e)
– Given s, define Φ by Φ(s(b)g) = (b, g).
9.4 Connections
Fix a principal G-bundle P → B. Write Rg : P → P for the right action of g.
Definition. A connection on P is a g-valued 1-form A on P , satisfying:
• A(p · ξ) = ξ for p ∈ P, ξ ∈ g.
• Rg∗ A = Adg−1 A (A is equivariant).
Given a local section sα (or equivalently a trivialization Φα ), the local connection 1-form
Aα is s∗α A.
Lemma 9.3. On overlaps we have Aα = Adg−1 Aβ + (Lg−1 )∗ dgβα .
βα βα
Conversely, given Aα transforming this way, they arise from a unique connection A on P .
Proof. Sheet 4.
44