Lecture_Notes_on_Smooth_Manifolds
Lecture_Notes_on_Smooth_Manifolds
MANIFOLDS AND
DIFFERENTIAL TOPOLOGY
WRITTEN BY
BAZINGA
This is a set of notes that I have written on basics of differentiable and smooth
manifolds, and some other topics that I liked a lot while learning, specifically de
Rham cohomology theory and its applications in differential topology. I was deeply
amazed by the beauty of these elegant results in differential topology, and the con-
crete geometric (in contrast to homological algebraic) approach to cohomology is
what I believe to be the most pedagogical approach to a first exposure in algebraic
topology.
The content in the majority of the text come from some of my old notes on differ-
ential manifolds that I wrote while self-studying and while taking graduate courses
on relevant topics. So technically it does not follow any textbooks or standard ref-
erences. However, the style of my writing was greatly influenced by [1], a series of
wonderful lectures in the WE-Heraeus International Winter School on Gravity and
Light by Frederic P. Schuller, and [2]. The readers are encouraged to read these
references if needed, since they certainly do a better job than what I did in these
notes.
The reader hoping to use these notes should have a solid foundation in multi-
variable calculus and linear algebra, and a certain level of mathematical maturity.
Ideally, they should also have a good background in point set topology, as I will only
review the relevant concepts very quickly in the beginning. Readers who does not
have relevant background mentioned above may find themselves having difficulties
to comprehend the text, and may want to do some extra reading on the background
material to catch up along the way.
I am deeply indebted to my professors: Xianzhe Dai, Guofang Wei, Rugang
Ye, Fedor Manin, and Daryl Cooper, for teaching me the theory of differentiable
manifolds, Riemannian geometry, and differential (algebraic) topology. I would also
like to thank my friends Merrick Hua and Benjamin Chung, for taking graduate
topology classes with me and doing a group reading on Bott and Tu’s book. Our
collective notes inspired the majority of Chapter 6.
Contents
I Basic Theory 7
0 Review of Basic Topology 8
0.1 Basic Concepts, Continuity . . . . . . . . . . . . . . . . . . . . . . . . 8
0.2 Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
0.3 Product Topology, Connectedness . . . . . . . . . . . . . . . . . . . . 11
1 Manifolds 13
1.1 Topological Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.2 Smooth Manifolds and Smooth Maps . . . . . . . . . . . . . . . . . . 15
1.3 Examples of Smooth Manifolds . . . . . . . . . . . . . . . . . . . . . 17
1.4 Partition of Unity and Applications . . . . . . . . . . . . . . . . . . . 19
3 Vector Bundles 37
3.1 Basic Definitions and Examples . . . . . . . . . . . . . . . . . . . . . 37
3.2 Tangent Bundle and Cotangent Bundle . . . . . . . . . . . . . . . . . 38
3.3 Vector, Covector, and Tensor Fields . . . . . . . . . . . . . . . . . . . 41
4 Differential Forms 43
4.1 Symmetric and Alternating Tensors, Wedge Product . . . . . . . . . 43
4.2 Differential Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3 Orientation on Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.4 Pullback and Exterior Differentiation . . . . . . . . . . . . . . . . . . 50
5 Integration on Manifolds 55
5.1 Basic Definitions and Properties . . . . . . . . . . . . . . . . . . . . . 55
5.2 Manifolds with Boundary and Boundary Orientation . . . . . . . . . 59
5.3 Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4 Applications of Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . 64
CONTENTS 6
Index 139
Bibliography 142
Part I
Basic Theory
0 | Review of Basic Topology
In this chapter we review the basic idea of topology and continuity defined on general
topological spaces. Here is the definition of a topology.
1. ∅, M ∈ O;
2. U, V ∈ O implies U ∩ V ∈ O;
S
3. If Ui ∈ O for some i ∈ A, then i∈A Ui ∈ O.
One example. Let M = {1, 2, 3}. Let OM = {∅, {1, 2, 3}}. You can immediately
verify that this is indeed a topology on M . We have two trivial topologies: the
smallest possible topology ∅, and the biggest possible topology P(M ), which is the
set of all subsets of M .
We introduce the standard topology on Rn .
You should verify that this topology OS on Rn satisfies the axioms above, thus
is indeed a topology. This topology is the the set of open sets in Euclidean space.
Definition 0.3. Suppose some set M is given some topology O. Together, (M, O)
is called a topological space.
Definition 0.4. Let (M, OM ) and (N, ON ) be topological spaces. Then a map
f : M → N is continuous if for all V ∈ ON , we have f −1 (V ) ∈ OM . A continuous
bijection with a continuous inverse is called a homeomorphism.
(g ◦ f )−1 (V ) = {m ∈ M : g ◦ f (m) ∈ V }
= {m ∈ M : f (m) ∈ g −1 (V )}
= f −1 (g −1 (V )).
O|S := {U ∩ S : U ∈ OM }.
The definition above says that every element in O|S must be in the form U ∩ S,
where U ∈ OM . The good thing about this definition is shown below.
Theorem 0.8. Given topological spaces (S, O|S ) and (N, ON ), the restricted map
f |S : S → N is continuous.
Proof. Let V ∈ ON . Since f : M → N is continuous, we have Pref (V ) ∈ OM .
Consider
f −1 (V ) ∩ S = {m ∈ M : f (m) ∈ V } ∩ {m ∈ S : S ⊂ M }
= {m ∈ S : f (m) ∈ V } = (f |S )−1 (V ).
it suffices to show that f −1 (B(p, ε)) is open for any p ∈ R and ε > 0.
Let x ∈ f −1 (B(p, ε)). Then f (x) ∈ B(p, ε), or equivalently |f (x) − p| < ε. Set
η = ε − |f (x) − p| > 0.
Since f is continous at x, we may choose δ > 0 such that |x0 − x| < δ implies
|f (x0 ) − f (x)| < η. In other words f (B(x, δ)) ⊂ B(f (x), η). Then by triangle
inequality:
|f (x0 ) − p| ≤ |f (x0 ) − f (x)| + |f (x) − p| < η + |f (x) − p| < ε.
Thus f (x0 ) ∈ B(p, ε) for all x0 ∈ B(x, δ). Hence we have shown that there exists
δ > 0 such that B(x, δ) ⊂ f −1 (B(p, ε)). This shows that f −1 (B(p, ε)) is open.
Conversely, we show that a continuous function f : R → R between topological
spaces is continuous in the δ-ε sense. Fix ε > 0, p ∈ R. Since f −1 (B(f (p), ε)) is an
open set (by assumption) that contains p, we can choose δ > 0 such that B(p, δ) ⊂
f −1 (B(f (p), ε)). Thus f (B(p, δ)) ⊂ B(f (p), ε). This is exactly the definition of δ-ε
continuity.
0.2 Compactness
Definition 0.10. A topological
S space X is called compact if for any open cover
{Uα }α∈A of X (i.e., X = α∈A Uα ), there exists a finite subcover. That is, there
exists a finite subcollection {U1 , ..., Un } ⊂ {Uα }α∈A such that X = U1 ∪ · · · ∪ Un .
In Rn , we know exactly what compact sets look like:
Theorem 0.11 (Heine-Borel). A set K ⊂ Rn is compact iff it is closed and bounded,
i.e., K ⊂ B(0, r) for some sufficiently large r > 0.
Example 0.12. Here are some examples of compact spaces:
• By Heine-Borel theorem, any closed interval [a, b] ⊂ R is compact. So is finite
union of them.
• Let X = {p1 , ..., pn } be a finite set. Let T be the collection of all subsets of
X. Then with respect to this topology, X is compact.
• Let X be a compact space and f : X → Y a continuous map between topo-
logical spaces. Then f (X) is a compact subspace of Y (subspace topology).
Proof. Let p ∈ X be any point and let s = f (p) ∈ Y . Then consider the set
S := {q ∈ X : f (q) = s}.
Since S = f −1 (q) and f is continuous, we see that S is closed. On the other hand,
for every q ∈ S, by locally constant hypothesis, there exists an open set Uq ⊂ X
such S that f (q 0 ) = s for all q 0 ∈ Uq . That is, f is constant on Uq with value s. Thus
S = q∈S Uq . Hence S is open. Since S is closed and open at the same time, and by
assumption p ∈ S, so S 6= ∅. Then since X is connected, we conclude that S = X,
i.e., f is globally constant.
Theorem 0.20. The interval [0, 1] with subspace topology of R, is connected. Sim-
ilarly, any interval [a, b] is connected.
Proof. Suppose [0, 1] is disconnected and U, V be two disjoint nonempty open subsets
such that U ∪ V = [0, 1]. Define a function f : [0, 1] → {±1} via
(
+1 x ∈ U
f (x) =
−1 x ∈ V.
Theorem 0.21. A path connected space is connected. (The converse fails, but in
the case of smooth manifolds, it is true)
Proof. Suppose X is path connected but disconnected. Then we can find nonempty
open sets U, V such that U ∪ V = X. Since both are nonempty, take p ∈ U, q ∈ V .
Since X is path connected, we may find a path γ : [0, 1] → X with γ(0) = p, γ(1) = q.
Since U, V are disjoint, γ −1 (U ), γ −1 (V ) are disjoint open sets that covers [0, 1]. But
[0, 1] is connected, contradiction.
1 | Manifolds
This is called a chart transition map, for the obvious reason that it serves as a
transition map that transform one set of coordinates of points in Uα ∩ Uβ to another
set. Note that τα,β is continuous. See Figure 1.1 for illustration.
φ(U )
For example, say if we have φ ◦ γ be a continuous map, then we better hope ψ ◦ γ
to be also continuous. Now,
ψ ◦ γ = ψ ◦ (φ−1 ◦ φ) ◦ γ = (ψ ◦ φ−1 ) ◦ (φ ◦ γ).
Since φ ◦ γ is continuous, and ψ ◦ φ−1 is continuous (since both ψ and φ−1 are
continuous), we have ψ ◦ γ be a continuous map.
However, if φ ◦ γ is differentiable, it need not be true that ψ ◦ γ is also differ-
entiable. This is due to the fact that ψ ◦ φ−1 only needs to be continuous. The
composition of a continuous map and a differentiable map is not necessarily differ-
entiable. Thus we need stronger condition for differentiability on manifolds, which
we will come back in the next section.
15 CHAPTER 1. MANIFOLDS
Definition 1.5. Let (U, φ), (V, ψ) ∈ A ⊂ A∞ be two charts of (M, O). We say
these two charts are ♥-compatible if either U ∩ V = ∅, or the chart translation
maps ψ ◦ φ−1 and (ψ ◦ φ−1 )−1 = φ ◦ ψ −1 have property ♥.
For example, if the two chart translation maps are at least once differentiable
(C ), then we say the two charts (U, φ), (V, ψ) are C 1 -compatible.
1
Definition 1.6. An atlas A♥ is ♥-compatible if any two charts (U, φ), (V, ψ) ∈ A♥
are ♥-compatible.
Now, we can solve the problem mentioned above. Let ♥ be the property that
the maps are infinitely differentiable, or C ∞ . Then we restrict our choice of charts
by only choosing charts in some AC ∞ . Then the chart transition map ψ ◦ φ−1 is
differentiable. Thus ψ ◦ γ = (ψ ◦ φ−1 ) ◦ (φ ◦ γ) is differentiable given that φ ◦ γ is
differentiable.
For the rest of this notes, we will always focus on smooth manifolds (M, O, AC ∞ ).
Smooth manifolds are sometimes also called C ∞ -manifolds. Usually, we will just
write M , assuming the information O, AC ∞ is clear. Usually we just refer to smooth
manifolds as manifolds to be concise.
It is generally not easy to construct a maximal smooth atlas on a topological
manifold. However, the next theorem saves us and suggests that constructing a
smoothly compatible atlas is sufficient to determine a smooth structure:
Proof. Let A be a smooth atlas for M , and let A∗ be the set of all charts that are
smoothly compatible with every chart in A. We need to show:
1.2. SMOOTH MANIFOLDS AND SMOOTH MAPS 16
ψ̃◦f ◦φ̃−1
φ̃(U ) ψ̃(V )
φ̃ ψ̃
f
φ̃◦φ−1 U V ψ̃◦ψ −1
φ ψ
φ(U ) ψ(V )
ψ◦f ◦φ−1
17 CHAPTER 1. MANIFOLDS
we see that
ψ̃ ◦ f ◦ φ̃−1 = (ψ̃ ◦ ψ −1 ) ◦ (ψ ◦ f ◦ φ−1 ) ◦ (φ̃ ◦ φ−1 )−1
is a composition of smooth maps in Euclidean spaces, hence is also smooth.
where
Ui+ = {(x1 , ..., xn+1 ) ∈ S n : xi > 0} Ui− = {(x1 , ..., xn+1 ) ∈ S n : xi < 0}
and the homeomorphisms φ± ±
i : Ui → B(0, 1) ⊂ R is given by
n
φ± 1
i (x , ..., x
n+1
) = (x1 , ..., x̂i , ..., xn+1 ).
It is easy to check that the defined maps are indeed homeomorphisms, with inverse
p
(φ± −1 1
i ) (a , ..., a ) = (a , ..., ±
n 1
1 − |a|2 , ai , ..., an ).
Hence we have a chart on S n , making it a topological manifold. Now to check
that AS n defines a smooth structure, we need to check that the transition maps are
smooth. In the case where i < j, we have
p
φ± ± −1 1
i ◦ (φj ) (u , ..., u ) = (u , ..., û , ..., ±
n 1 i 1 − |u|2 , ..., u)
which is clearly smooth on B(0, 1). A similar formula holds for i > j. When i = j,
the transition map is just identity on B(0, 1). Hence charts in AS n are smoothly
compatible, hence defining a smooth structure on S n . Thus S n is a smooth manifold
of dimension n.
1.3. EXAMPLES OF SMOOTH MANIFOLDS 18
Lemma 1.16. Any two norms on a finite-dimensional real vector space is equivalent,
so in particular they all give the same topology to V .
P P
Proof. Let k · k be the norm on V given by k i ai ei k = ( i |ai |2 )1/2 . Let F :
V → R≥0 be any norm. It suffices to prove that F is equivalent to k · k. Define
µ := max1≤i≤n F (ei ) < ∞. Then by Cauchy-Schwarz inequality one has
X X X √
F (x) = F ( ai e i ) ≤ |ai |F (ei ) ≤ µ |ai | ≤ µ nkxk. (1.1)
i i i
Since by the reverse triangle inequality one has F (x−y) ≥ F (x)−F (y) and similarly
F (x − y) = F (y − x) ≥ F (y) − F (x), one has
√
|F (x) − F (y)| ≤ |F (x − y)| ≤ µ nkx − yk.
Hence F is continuous on (V, k · k). Then F attains a minimum on the compact set
K = {x ∈ V : kxk = 1}
φ(x1 , ..., xn ) = x1 e1 + · · · + xn en .
Hence we see that (V, φ−1 ) is a chart. That is, the coordinates on V are given by
components of a vector with respect to the basis eα ’s. If e01 , ..., e0n is another basis
for V and φ0 : Rn → V is another chart that is defined by the same recipe but with
respect to the basis e01 , ..., e0n , then there is an invertible matrix T = (Tβα ) such that
X
eβ = Tβα e0α .
β
Thus we see that the chart transition map φ0 ◦ φ is given by the linear map Rn → Rn
which is exactly (Tβα ). Since this is an invertible linear map, it is a diffeomorphism
from Rn to itself. Thus φ, φ0 are smoothly compatible. The collection of all such
charts thus defines a smooth structure, called the standard smooth structure
on V .
19 CHAPTER 1. MANIFOLDS
AU = {(V, φ) ∈ AM : V ⊂ U }.
This clearly is a smoothly compatible atlas and gives U a smooth structure. Thus,
we call any open subset of a smnooth manifold M an open submanifold of M .
Example 1.18 (Matrix groups). Let M (n, R) denote the space of n × n matrices
with real coefficients. This is a real vector space of dimension n2 , hence a smooth
manifold of the same dimension by our example above. Let GL(n, R) denote the set
of invertible n × n matrices with real entries. Then GL(n, R) is an open subset of
M (n, R) because GL(n, R) = det−1 (R∗ ) where det : M (n, R) → R is the determi-
nant function (which is continuous because it only involves polynomial operations).
Hence GL(n, R) is a smooth manifold of dimension n2 as well.
1. 0 ≤ ρα ≤ 1 for all α;
3. Each p ∈ M has a neighborhood which intersects only finitely many Supp(ρα )’s;
P
4. α ρα (p) = 1 for all p ∈ M . By the previous condition, at every point this is
only a finite sum, hence is well-defined.
1.4. PARTITION OF UNITY AND APPLICATIONS 20
Our goal is to prove that for any open cover U of M , there exists a partition of
unity subordinate to U. However, the proof requires several very technical lemmas
in advance. Since the existence of partition of unity subordinate to any open cover
of a manifold is such an important result, we will present the full proof for future
references. However, this is a good point for impatient readers to stop for now, and
come back to it later when needed. We now begin the technical preparation required
for the proof.
Construction 1.22. Consider the function f1 : R → R given by
(
e−1/x x > 0
f1 (x) =
0 x ≤ 0.
It is a standard exercise in calculus to check that f1 is smooth (check all derivative
exists by L’Hospital rule). One sees that 0 ≤ f1 (x) ≤ 1 when x > 0 and f1 (x) = 0
when x ≤ 0. See Figure 1.2a.
Next, we consider f2 : R → R defined by
f1 (x)
f2 (x) = .
f1 (x) + f1 (1 − x)
Then one sees that f2 is a nonnegative smooth function, such that f2 (x) = 0 when
x ≤ 0, 0 < f2 (x) < 1 when 0 < x < 1, and f2 (x) = 1 when x ≥ 1. See Figure 1.2b.
Finally, consider the function f3 : Rn → R defined by
f3 (x) = f2 (2 − |x|).
Then one sees that f3 is a bump nonnegative function that is 1 on B(0, 1), and
decays smoothly until it vanishes outside of B(0, 2). See Figure 1.2c.
(c) Plot of f3
Figure 1.2: The plot of the three functions defined in Construction 1.22.
Lemma 1.24. For any topological manifold M , there exists a countable collection
of open sets {Xi }, which is called an exhaustion of M , so that
1. For each j, the closure X j is compact;
φ(p) ∈ B ⊂ φ(U )
and an element S in the basis B such that p ∈ S ⊂ φ−1 (B). Then we see that
X 1 ⊂ Yi 1 ∪ · ∪ Yi k .
Now define
X 2 = Y 2 ∪ Yi 1 ∪ · ∪ Yi k .
1.4. PARTITION OF UNITY AND APPLICATIONS 22
Lemma 1.25. For any open covering U = {Uα } of a topological manifold M , one
can find two countable family of open covers V = {Vj } and W = {Wj } such that
(i) For each j, V j is compact and V j ⊂ Wj ;
Uα(p)
M
Wp
Vp p
Figure 1.3: Since Uα(p) ∩ (Xj+2 − X j−1 ) is the intersection of two ooen sets contain-
ing p, it is open. Intersecting this open set with a chart near p and map it down to
Rm under the homeomorphism φ, we get an open set U 0 ⊂ Rm under φ containing
φ(p). Then choose V 0 , W 0 ⊂ U 0 such that φ(p) ∈ V 0 ⊂ W 0 . Then set Vp = φ−1 (V 0 )
and Wp = φ−1 (W 0 ).
Now for each j, since the strip X j+1 −Xj is compact, one can choose finitely many
points p1,j , ..., pkj ,j so that Vp1,j , ..., Vpkj ,j is an open cover of X j+1 − Xj . Denote all
these Vpk,j ’s by V1 , V2 , V3 , ..., and the corresponding Wpk,j ’s by W1 , W2 , W3 , .... Then
V = {Vj } and W = {Wj } clearly satisfies (i) and (ii) as in the statement of the
lemma. For (iii), the local finiteness property of W follows from the fact that there
are only finitely many Wk ’s intersecting Xj+1 − X j−1 .
23 CHAPTER 1. MANIFOLDS
Lemma 1.26. Closure of unions of locally finite sets in a topological space is the
union of closure of these sets. In other let {Si : i ∈ I} be a collection of
S words, S
locally finite sets on M , then we have i∈I Si = i∈I S i .
S S S S
Proof. Since Si ⊂ i∈I Si , we have S i ⊂ i∈I Si , hence i∈I S i ⊂ i∈I Si . So the
nontrivial part is the inclusion in the other direction.
S To prove the other inclusion, by the definition of closure it suffices to prove S that
i∈I S i is closed. To this end, we show its complement is open: Take x ∈ / i∈I S i .
Since we have a locally finite collection, we may find an open neighborhood U of x
that only intersects finitely many Si ’s. Say U intersects Si1 , ..., Sin . Now consider
the new open set
U 0 := U \ (S i1 ∪ · · · ∪ S in ).
Note that U 0 is still an open set containing x, and it does not intersect any of the Si ’s.
We claim that this implies U 0 does not intersect any of the Si ’s either: Indeed, since
the complement (U 0 )c of U 0 is a closed set containing all the Si ’s, by Sminimality of
the closure, this means it also contains all the S i ’s. Hence x ∈ U 0 ⊂ ( i∈I S i )c .
Theorem 1.27. Let M be a smooth manifold, and {Uα } an open cover of M . Then
there exists a partition of unity {ρα } subordinate to {Uα }.
0 ≤ ϕj ≤ 1 ϕj | V j ≡ 1 Supp(ϕj ) ⊂ Wj .
P
Since W is a locally finite covering, the function ϕ = j ϕj is a well-defined smooth
function on M . Since each ϕj is nonnegative, and V is a covering of M , ϕ is strictly
positive on M . PIt follows that the functions ψj = ϕj /ϕ are smooth and satisfy
0 ≤ ψj ≤ 1 and j ψj = 1.
Next, we need to re-index the family {ψj } to get the required partition of unity
subordinate to {Uα }. For each j, we fix an index α(j) so that Wj ⊂ Uα(j) , and define
X
ρα = ψj .
{j:α(j)=α}
Note that the right hand side is a finite sum near each point, so it does define a
smooth function. Moreover, By Lemma 1.26, we have
[ [ [
Supp(ρα ) = Supp(ψj ) = Supp(ψj ) = Supp(ψj ) ⊂ Uα .
{j:α(j)=α} {j:α(j)=α} {j:α(j)=α}
That is, we take all possible curves that passes p, and find their velocity at point
p. These velocity vector spans a imaginary plane that forms the tangent space Tp M .
Now it is natural to ask whether Tp M , if it is like any plane in Euclidean space,
forms a vector space. We define the addition and multiplication on Tp M and show
that it is indeed a vector space.
27 CHAPTER 2. TANGENT AND COTANGENT SPACES
⊕ : Tp M × Tp M → Hom(C ∞ (M ), R)
: R × Tp M → Hom(C ∞ (M ), R)
to be
(λ vγ,p )(f ) := λ · vγ,p (f ),
where happens in Tp M and · happens in R.
where the third step follows from the chain rule. This finished the second half of the
proof. Now, for the first part, make a choice of chart (U, φ) where p ∈ U . Define
the curve σ : R → M to be
Notice that
Now,
Notice that f ◦φ−1 : Rm → R, and φ◦σ : R → Rm are functions that are very easy to
differentiate by using the multidimensional chain rule. We continue the calculation
and use chain rule to get
From now on, we will use this convention throughout this notes, to be concise.
The notation (∂/∂xi )p looks deceptively simple and familar (after all it is just
like the partial derivative), but it is fundamentally different from the familiar partial
29 CHAPTER 2. TANGENT AND COTANGENT SPACES
derivative that we know. So, from now on, whenever it appears, what we mean is
the operation
∂ ∞ ∼ ∂(f ◦ φ−1 )
: C (M ) −
→R f 7→ (φ(p))
∂xi p ∂xi
dγ j d(φ ◦ γ i )
(0) = =0
dλ dλ
unless j = i. That is, a curve γ such that its image φ ◦ γ in Rm is constant for every
j 6= i. The existence and uniqueness theorem of ODEs guarantee that this curve is
unique at least in a neighborhood of p. See Figure 2.1.
M
∂
x2
φ(U )
x2 ∂x1 p
φ x1
x 1 p
φ(p)
γ φ◦γ
U
Note that the notion of (∂/∂xi )p only makes sense if a chart (U, φ) around p is
specified, as φ is explicitly needed in its definition. But overall we would like to
show the following:
Proof. We have already shown that any vector vγ,p ∈ Tp M can be written as a linear
combination of these vectors in the discussion above. So we only need to show that
these vectors are linearly independent. We will show that
i ∂
λ = 0 implies λi = 0 for all i.
∂xi p
2.2. PUSHFORWARD OF A SMOOTH MAP 30
Proof. This is immediate since given any chart (U, φ) around p, we have shown that
(∂/∂x1 )p , ..., (∂/∂xm )p is a basis.
Example 2.8. Let p = (p1 , ..., pn ) ∈ Rn . The tangent space Tp Rn consists of all
linear maps C ∞ (Rn ) → R defined by C 1 -curves γ : R → Rn . Let x1 , ..., xn be the
standard coordinate on Rn . A basis for Tp Rn consists of tangent vectors of the
constant curves η i (λ) = (p1 , ..., pi + λ, ..., pn ) at λ = 0. It is easy to see that the
associated basis (∂/∂xi )p turns out to be exactly the directional derivatives in the
ith direction, evaluated at p. Let e1 = (1, ..., 0), ..., en = (0, ..., 1) be the standard
basis for Rn . Then the identification
X n Xn
∂ ∼
=
v= v i
i
−
→ v i ei ∈ Rn
i=1
∂x p i=1
gives an isomorphism Tp Rn ∼
= Rn of vector spaces.
Note that this definition makes sense, because f ◦ F ∈ C ∞ (M ), and v acts linearly
on C ∞ (M ). This map dFp is called the differential or pushforward of F at p.
See Figure 2.2 for a more geometric description of the pushforward.
M N
F (γ)
v
dFp (v)
γ F
F (p)
p
Figure 2.2: If v is the tangent vector of a curve γ at p and γ(0) = p, then for
any g ∈ C ∞ (M ), we have v(g) = (g ◦ γ)0 (0). Then using the definition dFp (v)(f ) =
v(f ◦ F ) = (f ◦ F ◦ γ)0 (0). Therefore, the geometric picture of dFp can be viewed
as follows: it maps v, the tangent vector of γ ⊂ M at p, to the tangent vector of
F ◦ γ ⊂ N at F (p).
Proof. 1. This is clear from definition, since v is linear; 2. This is also clear because
for f ∈ C ∞ (P ) and v ∈ Tp M , we have
This map is a diffeomorphism (in fact a chart) by Example 1.15. Therefore, we can
make sense of dϕw : Tw V → Tϕ(w) Rn . Then the three isomorphisms in the diagram
below defines an identification isomorphism θ : V → Tw V :
θ
V Tw V
ϕ dϕw
Rn ∼
=
Tϕ(w) Rn
Example 2.12 (Another way of viewing basis for Tp M ). Let (U, φ) be a chart
around p ∈ M . Then we may view φ−1 : φ(U ) → U a smooth map (in fact a
diffeomorphism) between manifolds. Note that because curves on U are curves on
M as well, and curves on M restricts to curves on U , this gives an identification
Tp U ∼
= Tp M .Since φ−1 is a diffeomorphism, the pushforward
Hence we see that the basis for Tp M induced by the chart (U, φ) is just the image
of the standard basis in Tφ(p) φ(U ) ⊂ Rn under the isomorphism d(φ−1 )φ(p) of vector
spaces.
We conclude this section by studying the expression of the differential dF :
Tp M → TF (p) N in coordinates.
Theorem 2.13. With respect to the basis for Tp M induced by a chart (U, φ) around p
and the basis for TF (p) N induced by a chart (V, ψ) of N containing F (U ), the matrix
representation of the linear map dFp : Tp M → TF (p) N is given by the Jacobian of
F̂ := ψ ◦ F ◦ φ−1 : φ(U ) → ψ(V ) evaluated at φ(p).
Proof. Let (∂/∂x1 )p , ..., (∂/∂xm )p a basis for Tp U induced by the chart (U, φ), and
similarly let (∂/∂y 1 )F (p) , ..., (∂/∂y n )F (p) a basis for TF (p) V induced by the chart
(V, ψ). We would like to study dFp in matrix form with respect to these base.
Now let f ∈ C ∞ (N ) be arbitrary, by chain rule we have
!
∂ ∂
dFp i
f= i
f ◦ F ◦ φ−1 (φ(p))
∂x p ∂x
∂
= i
f ◦ ψ −1 ◦ ψ ◦ F ◦ φ−1 (φ(p))
∂x
∂(f ◦ ψ −1 ) ∂(ψ ◦ F ◦ φ−1 )j
= (ψ(F (p))) · (φ(p))
∂y j ∂xi
∂ F̂ j ∂
= (φ(p)) · f
∂xi ∂y j F (p)
This is easy to remember because of its resemblance to the chain rule formula in
multivariable calculus. This also generalizes to arbitrary tangent vectors v ∈ Tp M .
We first write v as a linear combination of basis vectors, then by above we get
j
∂ i ∂y ∂ ∂
v = vi i
= v i j
= ṽ j j ,
∂x ∂x ∂y ∂y
or simply
∂y j
ṽ j = v i .
∂xi
Example 2.14 (Tangent space of the circle). In this example, we explicitly calculate
the tangent space of S 1 at a point (a, b).
Rotating the circle if necessary, we may assume (a, b) lies on the upper half circle,
i.e. b > 0. Then a local parametrization is given by ψ : (−1, 1) → S 1 , where
√
ψ(x) = (x, 1 − x2 ).
Note that since dψx : R → R2 , this map should take a real number and
√ gives a vector,
which is done by scalar multiplication. Now if we let x = a, then 1 − x2 = b since
(a, b) is on the upper half circle. Therefore
1 1 −b
T(a,b) (S ) = dψa (R) = R
1
= R(−b) =R .
−a/b −a/b a
This is what we expect: the tangent space of the circle at the point p = (a, b) should
be spanned by the vector (−b, a), which is perpendicular to the radial vector (a, b)
at the point p.
2.3 Tensors
We provide a section that briefly reviews the concept of tensors.
Definition 2.15. Let (V, +, ×) be a finite dimensional vector space. An (r, s)
tensor T is a multilinear map
∗ ∗ ∗
T :V · · · × V} × V
| × V {z | × V {z· · · × V } → R.
r s
Multilinear means the tensor is linear in each slots. Say if T is a (1, 1) tensor.
Let φ, ψ ∈ V ∗ , let v, w ∈ V , and let λ ∈ F (some field, e.g. real or complex). Then
Example 2.16. Let P(R) be the vector space of polynomials on R and let g :
P(R) × P(R) → R be defined as
Z 1
g(p, q) = p(x)q(x)dx.
−1
Then you can verify that g is a (2, 0) tensor. From linear algebra, you know this is
an inner product. More generally speaking, inner products are (2, 0) tensors.
Of course, φ : V → R is a (1, 0) tensor.
Vectors are (0, 1) tensors. This is because, for finite dimensional vector space,
∗ ∗ ∼
(V ) = V . The identification is given by the correspondence that for any v ∈ V
and α ∈ V ∗ , we may define v(α) := α(v) ∈ R, hence viewing v as an element in V ∗ .
Similar to choosing a basis in vector space V , we can choose a basis for V ∗ .
Definition 2.17. Let V be a finite dimensional vector space with basis e1 , · · · , en .
Define ε1 , · · · , εn ∈ V ∗ where
εa (eb ) = δba .
ε1 , · · · , εn are uniquely determined by the choice of e1 , · · · , en . It is easy to verify
that ε1 , ..., εn form a basis for V ∗ . We call ε1 , · · · , εn the dual basis for V .
Example 2.18. Consider the vector space P3 (R), the set of real polynomials with
degree no more than 3. Let
e0 (x) = 1 e1 (x) = x e2 (x) = x2 e3 (x) = x3 ,
Let
1 a
εa = ∂
a! x=0
a
Then ε (eb ) = δba is a dual basis.
Definition 2.19. Let T be an (r, s) tensor on a finite dimensional vector space V
(dim V = n < ∞). Let e1 , · · · , en be the chosen basis for V . We can define the
components of T by
Tji11,··· ,js = T (ej1 , · · · , ejs , ε , · · · , ε ),
,··· ,ir i1 ir
XV
dim XV
dim
T (v, φ) = T ( v j ej , φ i εi )
j=1 i=1
XX
= φi v T (ej , εi )
j
i j
XX
= φi v j Tji
i j
= φi v Tji ,
j
where the second step is by the multilinear property of T . In the last step we use
Einstein summation convention and contract the indices.
35 CHAPTER 2. TANGENT AND COTANGENT SPACES
to the chart induced by the identity map on R with 1 ∈ R, then the two definitions
coincide, as the formula above is exactly Theorem 2.13.
Theorem 2.23. Let (U, φ) be a chart and xi : U → R be the smooth function defined
for any p ∈ U by giving the ith component of φ(p) ∈ Rm . Then the gradient vectors
(dx1 )p , (dx2 )p , · · · , (dxm )p ∈ Tp∗ M
form a basis for Tp∗ M .
Proof. By the calculation above, it is easy to see that
!
∂ ∂(xi ◦ φ−1 ) ∂xi 1
(dxi )p = (φ(p)) = (x , ..., xm ) = δji .
∂xj p ∂xj ∂xj
That is, (dxi )p ’s are the dual basis to the basis ∂x∂ j p ’s in Tp M . Hence it is a basis
for Tp∗ M .
Now we can talk about the change of components of a covector under a change of
charts. Similar to our discussion on change of components of vectors, let (U, φ), (V, ψ)
be two overlapping charts containing p, and let xi , y j be the respective coordinates.
Then we have
(dxi )p vγ,p = vγ,p (xi ) = (xi ◦ γ)0 (λ0 ) = (xi ◦ ψ −1 ◦ ψ ◦ γ)0 (λ0 )
∂(xi ◦ ψ −1 )
= (y j ◦ γ)0 (λ0 ) · (ψ(p))
∂y j
∂(xi ◦ ψ −1 )
= vγ,p (y j ) · (ψ(p))
∂y j
∂(xi ◦ ψ −1 )
= (ψ(p)) · (dy j )p vγ,p .
∂y j
2.4. COTANGENT SPACES 36
Note that xi ◦ ψ −1 : ψ(U ∩ V ) → R is the function that gives the ith component
of the map φ ◦ ψ −1 : ψ(U ∩ V ) → φ(U ∩ V ). Similar to our previous argument on
change of components of tangent vectors, we may view xi as a function of y 1 , ..., y m ’s,
so that xi (y 1 , ..., y m ) is the ith component of φ ◦ ψ −1 (y 1 , ..., y m ). So the equation
above can be conveniently written as
i
i ∂x
(dx )p = (dy j )p .
∂y j p
α = αi (dxi )p .
or simply
∂xi
α̃j = αi .
∂y j
Remark 2.24. Coming back to gradient dfp of a smooth function f ∈ C ∞ (M ). At
a local chart around p we may write dfp as a linear combination dfp = βi (dxi )p . To
figure out what the coefficient βi is, we note that
∂ i ∂
dfp = βi dx = βi δji = βj .
∂xj ∂xj
∂(f ◦ φ−1 )
βj = (φ(p)).
∂xj
Note that in the case where M = Rn , and φ is the identity map, we have exactly
∂f
dfp = j
(dxj )p .
∂x
Therefore, differentials should really be viewed as elements in the cotangent space
Tp∗ Rn . There are higher dimensional generalizations of them called differential forms,
which we will get to later.
3 | Vector Bundles
There is also the notion of a topological bundle, where the definition is the same
as above, except that we only require E, M to be topological spaces, and π : E → M
be a continuous and surjective map.
Note that from the definition above, in particular the surjectivity of π, we see
that as a set, [
E= π −1 (p)
p∈M
(i) For each p ∈ M , the fiber Ep = π −1 ({p}) over p is endowed with the structure
of a k-dimensional real vector space.
σ(M )
p
M
π −1 (p)
Step 1
Define the tangent boundle (on the level of set) as
a
T M := Tp M,
p∈M
which is the disjoint union of all the tangent space Tp M . Here, the disjoint union is
the set defined by a
Tp M = {(p, v) : p ∈ M, v ∈ Tp M }.
p∈M
39 CHAPTER 3. VECTOR BUNDLES
Step 2
Define our map π : T M → M such that
π : (p, v) 7→ p,
where p ∈ M, v ∈ Tp M . By the definition of T M , since we unite over all p ∈ M , we
know this map π will hit all points on M . Hence π is a surjective map. Now we have
π : T M → M where T M only has a set structure and we don’t know whether it is
smooth or not. π is only a surjective map for now. We don’t know if it is smooth
either. M is a smooth manifold by assumption. If we want to make this triple a
bundle, then we have to make T M a smooth manifold.
Step 3
Now, T M is just a set. We first make T M a topological manifold first by giving
it a proper topology. Let us require OT M to be the coarsest topology to make π
continuous. We define
OT M := {π −1 (U ) : U ∈ O}.
We claim that this is indeed a topology.
Proof. We verify that this axiom satisfies the definition of a topology, i.e., Definition
0.1:
1. Note that ∅ ∈ O, and π −1 (∅) = ∅ since π is surjective, we have ∅ ∈ OT M . Also,
M ∈ O and π −1 (M ) = T M since π surjective, we have T M ∈ O.
3. Let Ui0 ∈ OT M for some i ∈ I. Then there exists Ui ∈ O such that we have
Ui0 = π −1 (Ui ) for the same i ∈ I. Then
[ [
Ui0 = π −1 (Uj ).
i∈I j∈I
S
But since Uj ∈ O, we have i∈I Ui0 ∈ OT M .
This concludes the proof.
Step 4
Now with a given topology, we construct a C ∞ atlas from A on M . Define
AT M := {(π −1 (U ), ξφ ) : (U, φ) ∈ A},
where ξφ : π −1 (U ) → R2m defined by the rule
ξφ : (p, v) 7→ (x1 (p), ..., xm (p), (dx1 )p (v), ..., (dxm )p (v)).
3.2. TANGENT BUNDLE AND COTANGENT BUNDLE 40
Recall that xi (p) is the ith coordinate of φ(p) ∈ Rm . All this is saying is that given
a chart (U, φ) containing p, we take the first m coordinate of (p, v) ∈ T M under ξφ
to be the coordinates of p under φ; then for the second m coordinates, express v as
a linear combination of basis tangent vectors in Tp M induced by the chart φ (see
the previous chapter),
j ∂
v=v .
∂xj p
Then our second m coordinates will be v 1 , ..., v m . It is easy to see that the inverse
of ξφ is given by
!
∂
ξφ−1 : (α1 , ..., αm , β 1 , ..., β m ) 7→ φ−1 (α1 , ..., αm ), β j ∈ T M.
∂xj p
Step 5
To show that we have constructed a smooth atlas, we still need to show that the
chart transition maps are smooth. Hence, let (U, φ), (V, ψ) be two overlapping charts
on M , and xi , y j be the respective coordinates. Then
!
∂
ξψ ◦ ξφ−1 (α1 , ..., αm , β 1 , ..., β m ) = ξψ φ−1 (α1 , ..., αm ), β j
∂xj p
k
!
∂y ∂
= ξψ φ−1 (α1 , ..., αm ), β j j
∂x ∂y k p
1 m
−1 j ∂y j ∂y
= ψ ◦ φ (α , ...α ), β
1 m
, ..., β .
∂xj ∂xj
This is clearly smooth with respect to α1 , ..., αm , β 1 , ..., β m , because recall that
∂y k /∂xj = ∂(ψ ◦ φ)/∂xj by our definition, and since A is a smooth atlas, we know
that ψ◦φ−1 ∈ C ∞ . By Theorem 1.9, we see that AT M determines a smooth structure
on T M , making T M a smooth manifold.
Conclusion
This completes the construction. See what have we got: we defined a smooth
manifold T M , and a smooth and surjective map π : T M → M . And finally, we get
a bundle T M , which is called the tangent bundle of M , as it is constructed by
gluing pieces of tangent spaces of M . Concretely, we have constructed a space that
contains the information of all points of M along with all possible tangent spaces of
M . To a physicist, M can be thought of as a space where the particle of interests
move along, and the particle, when at a point p ∈ M , has a velocity v ∈ Tp M . As
the particle moves along M , physicists may want to write down the pair (p, v), for
p ∈ M all points the particle was at, and v the corresponding velocity at p. What
physicist will get eventually is a curve in T M .
Remark 3.4 (The Cotangent Bundle). One can construct cotangent bundle of a
smooth manifold M m in basically the same way: In step 1, we define the cotangent
41 CHAPTER 3. VECTOR BUNDLES
`
bundle T ∗ M = p∈M Tp∗ M ; The projection map π and the topology OT ∗ M is defined
exactly as in Step 2 and Step 3 above; The chart construction (π −1 (U ), ξφ ) for a
chart (U, φ) is similar to that in step 4: the first m coordinates of ξφ (p, α) for
p ∈ U, α ∈ Tp∗ M is still given by the coordinates φ(p) ∈ Rm , and then express
α = αi (dxi )p , then the last m coordinates will be α1 , ..., αm ; Similar to Step 5, one
can show that this indeed defines a smooth atlas.
Example 3.5 (Tangent Bundle of S 1 ). The tangent bundle T S 1 of the circle S 1 is
diffeomorphic to S 1 × R, which is shown in Figure 3.1. To see this, note that for
any p = (a, b) ∈ S 1 ⊂ R2 , we have dim Tp S 1 = 1, as S 1 is one dimensional. In fact,
the calculation in Example 2.14 shows that T(a,b) S 1 is spanned by a single tangent
vector ∂/∂θ|(a,b) = (−b, a). Hence we can define a map ψ : T S 1 → S 1 × R via
ψ(p, λ · ∂/∂θ|p ) = (p, λ). It is easy to check that ψ is a diffeomorphism.
(g X)f = g · X(f ),
where we used C ∞ (M ) rather than R because the tensor field need not to be a
constant. Instead, it varies from point to point.
Remark 3.10. Notice that, for every p ∈ M , and X ∈ Γ(T M ), we have X(p) ∈
Tp M . Hence, at a point, a vector field is just a vector in the tangent space. Similarly
for α ∈ Γ(T ∗ M ), we have α(p) ∈ Tp∗ M , hence at a point a coverctor field is just a
covector. Similarly, at a point p, a tensor field is just a tensor in the sense of Section
2.3, where V = Tp M .
4 | Differential Forms
· · · αk : (v1 , ..., vk ) 7→ α1 (v1 )α2 (v2 ) · · · αk (vk ). Then we have two types of important
tensors: If α ∈ T 0,k (V ), then α is said to be symmetric if
α(v1 , ..., vk ) = α(vσ(1) , ..., vσ(k) )
Q
for any (v1 , ..., vk ) ∈ k V and σ ∈ Sk . Equivalently, one can define α to be the
unchanged by interchanging any pair of arguments. We say α is alternating if
α(v1 , ..., vk ) = sgn(σ)α(vσ(1) , ..., vσ(k) )
Q
for any (v1 , ..., vk ) ∈ k V and σ ∈ Sk . Equivalently, one can define α to be the
tensor that takes the negative of the original value by interchanging any pair of
arguments.
Now, given any α ∈ T 0,k (V ), we can define its symmetrization S (α) to be
1 X
S (α)(v1 , ..., vk ) = α(vσ(1) , ..., vσ(k) )
k! σ∈S
k
It is clear that S (α) is a symmetric tensor. One can also define the antisym-
metrization A (α) to be
1 X
A (α)(v1 , ..., vk ) = sgn(σ)α(vσ(1) , ..., vσ(k) ).
k! σ∈S
k
1 X sgn(η)
= α(vη(1) , ..., vη(k) )
k! η∈S sgn(τ )
k
where we made the substitution η = σ ◦ τ and we used sgn(η) = sgn(σ) sgn(τ ). But
since sgn(τ ) = ±1, so dividing by sgn(τ ) is the same as multiplying by it.
∧k V ∗ = A (⊗k V ∗ ) = {σ ∈ ⊗k V ∗ : σ alternating}.
Definition 4.2. Let V n be a vector space, and ε1 , ..., εn any basis for V ∗ . Let
I = (i1 , ..., ik ) be a multi-index of length k, where 1 ≤ is ≤ n. We define the
so called elementary alternating tensor εI to be the tensor such that for any
v1 , ..., vk ∈ V ,
εi1 (v1 ) · · · εi1 (vk )
εI (v1 , ..., vk ) := .. ..
. ··· . .
εik (v1 ) · · · εik (vk )
By definition, εI is alternating because the determinant is, hence εI ∈ ∧k V ∗ .
Lemma 4.3. Let e1 , ..., en be a basis for V , and ε1 , ..., εn be the dual basis for V ∗ .
Let εI be defined as above where I = (i1 , ..., ik ).
δji11 · · · δji1k
εI (ej1 , ..., ejk ) = ... · · · .. =: δ I .
. J
δjik1 · · · δjikk
Proof. For (1), if I has an repeated index, then the determinant in the definition
of εI will have two identical rows, hence is equal to zero; Next, (2) is immediate
because interchanging I by σ amounts to interchanging columns of the determinant
in the definition of εI by σ; Finally, (3) is just obvious from the definition of εI .
Theorem 4.4. Let V n be a vector space with basis e1 , ..., en and ε1 , ..., εn the dual
basis for V ∗ . Then, for each positive k ≤ n, the collection of εI with increasing
multi-index of length k, i.e.,
Proof. To show that E spans ∧k V ∗ : For each multi-index I = (i1 , ..., ik ), define
αI := α(ei1 , ..., eik ). Since α is alternating, αI = 0 if I has an repeated index and
45 CHAPTER 4. DIFFERENTIAL FORMS
To see this, let J = (j1 , ..., jk ) be any multi-index set. Note that by the definition
of δJI above, we can see that δJI = 0 if I or J has an repeated index or if J is not a
permutation of I; and δJI = sgn σ if neither I nor J has a repeated index and J = Iσ
for some σ ∈ Sk . Then by Lemma 4.3,
X X
αI εI (ej1 , ..., ejk ) = αI δJI = αJ = α(ej1 , ..., ejk ).
I increasing I increasing
εI ∧ εJ = εIJ .
Proof. By linearity, it suffices to show that εI ∧ εJ and εIJ give the same result for
any sequence (ep1 , ..., epk+l ) of basis vectors. Denote P = (p1 , ..., pk+l ).
We first reduce the problem by restricting our attention to nontrivial multi-index
P in which the equation is not obviously true. Note that:
• If P contains a repeated index, then since both sides are clearly alternating
tensors, both sides are again zero.
• We will discuss the case where P = IJ and has no repeated index later.
4.1. SYMMETRIC AND ALTERNATING TENSORS, WEDGE PRODUCT 46
By Lemma 4.3 again, the only terms in the sum above that is nonzero are the ones
in which σ permutes the first k indices and the last l indices of P separately. In
other words, σ = τ η where τ ∈ Sk permutes the index 1, ..., k, and η ∈ Sl permutes
the index k + 1, ..., k + l. Since sgn(τ η) = (sgn τ )(sgn η), we have
1 X
εI ∧ εJ (ep1 , ..., epk+l ) = (sgn τ )(sgn η)εI (epτ (1) , ..., epτ (k) )εJ (epk+η(1) , ..., epk+η(l) )
k!l! τ ∈S ,η∈S
k l
1 X 1X
= (sgn τ )εI (epτ (1) , ..., epτ (k) ) · (sgn η)εJ (epk+η(1) , ..., epk+η(l) )
k! τ ∈S l! η∈S
k l
where the second to last inequality is because εI , εJ are already alternating, and
the last inequality is because of the assumption P = IJ. Note that since P = IJ,
clearly εIJ (eP ) = 1 by Lemma 4.3, so indeed εI ∧ εJ = εIJ .
Proposition 4.7. The wedge product is bilinear, associative, and anticommutative,
i.e., ω ∧ η = (−1)deg ω deg η η ∧ ω.
Proof. Bilinearity is immediate from definition of wedge product, because both the
tensor product and alternation are bilinear. To prove associativity, note on basis
elements associativity holds:
The general case follows from bilinearity. To show anticommutativity, note that
again this holds on basis elements:
where J = (j1 , ..., jn ) runs over all n-tuples where each entry takes value between
1, ..., n. However, the previous theorem tells us that if J contains any repeated
indices, then vj1 ∧ · · · ∧ vjn = 0. Therefore we can in fact take J to run over all
permutations of (1, ..., n). Therefore the equation above actually becomes
X σ(1)
L(v1 ∧ · · · ∧ vn ) = (a1 · · · anσ(n) )(vσ(1) ∧ · · · vσ(n) ) (4.1)
σ∈Sn
X σ(1)
= sgn(σ)(a1 · · · anσ(n) )(v1 ∧ · · · ∧ vn ) (4.2)
σ∈Sn
Definition 4.10. Let V n be a vector space and let v1 , ..., vn and w1 , ..., wn be two
basis for V . Then these two basis are said to be equivalent if the linear map
A : V → V such that wj = Avj has positive determinant, i.e., det A > 0.
Corollary 4.11. The two basis v1 , ..., vn and w1 , ..., wn of V are equivalent iff
w1 ∧ · · · ∧ wn = λ(v1 ∧ · · · ∧ vn )
where ξφ is defined as follows: Let p ∈ U and dx1 , ..., dxm be the induced basis for
Tp∗ M via the chart map φ. Then any α ∈ ∧k Tp∗ M can be written as a summand
X
α= ai1 ,...,ik dxi1 ∧ · · · ∧ dxik
1≤i1 <···<ik ≤m
m
where the summand contains k
terms (some could be zero) by Theorem 4.4 and
Corollary 4.8. Then define
ξφ : (p, α) 7→ (φ(p), aI ) ∈ Rm × R( k )
m
where the first m coordinates are the coordinates of φ(p), and the last m k
coor-
dinates are the coefficients ai1 ,...,ik . Clearly this map is invertible by definition. To
check that this is a smooth atlas, i.e., the transition maps are smooth, let (V, ψ) be
an overlapping chart with (U, φ) and let dy 1 , ..., dy m be the induced basis for Tp∗ M
via the chart map ψ. Then
X i1 ik
∂x ∂x
α= ai1 ,...,ik j1
dy j1
∧ ··· ∧ jk
dy jk
1≤i1 <···<ik ≤m
∂y ∂y
X i1
∂x ∂xik
= ai1 ,...,ik j
· · · j dy j1 ∧ · · · ∧ dy jk ,
1≤i <···<i ≤m
∂y 1 ∂y k
1 k
where the indices j1 , ..., jk are summed over using the summation convention. Since
∂xis /∂y js are smooth functions because φ, ψ are smoothly compatible charts, we see
that from definition ξψ ◦ ξφ−1 is smooth. This concludes the construction.
Definition 4.12. A section of ∧k T ∗ M is called a differential k-form. The integer
k is called the degree of the form. We say α is a smooth k-form is as a section
α : M → ∧k T ∗ M is a smooth map between manifolds. We denote the space of
smooth k-forms by Ωk (M ).
Proposition 4.13. A k-form α can be written, locally in a chart (U, φ), as a sum-
mand X
α= ai1 ,...,ik dxi1 ∧ · · · ∧ dxik .
1≤i1 <···<ik ≤m
Then α is smooth iff the functions ai1 ,...,ik ∈ C ∞ (U ), for every open set U in the
atlas AM .
Proof. First, assume α : M → ∧k T ∗ M is a smooth section, then for every chart
(U, φ) in the atlas, the component functions are just the components of the com-
position of smooth maps ξφ ◦ α ◦ φ−1 , which are of course smooth. The converse
implies the smoothness of α : M → ∧k T ∗ M as a section follows immediately from
Definition 1.10.
Corollary 4.14. If (U, φ, xi ) and (V, ψ, y i ) are two overlapping smooth coordinate
charts on M n , then the following identity holds on U ∩ V :
i
∂y
dy ∧ · · · ∧ dy = det
1 n
dx1 ∧ · · · ∧ dxn = [J(ψ ◦ φ−1 ) ◦ φ]dx1 ∧ · · · ∧ dxn .
∂xj
Proof. This follows immediately from (4.3).
49 CHAPTER 4. DIFFERENTIAL FORMS
Proof. Let ω be a nowhere vanishing n-form on M . Then for each chart (U, φ, xi )
there exists a nowhere vanishing function f 6= 0 such that ω = f dx1 ∧ · · · ∧ dxn .
We can always take such a chart near each point so that f > 0, otherwise we can
replace x1 by −x1 . Now suppose (Uα , φα ) and (Uβ , φβ ) are two such charts. Then
suppose ω = f dx1α ∧ · · · ∧ dxnα = gdx1β ∧ · · · ∧ dxnβ where f, g > 0. Then
So J(φβ ◦ φ−1
α ) = g/f > 0.
Conversely, let A be an orientation. For each local chart Uα , we consider the
form ωα = dx1α ∧ · · · ∧ dxnα . Pick a partition of unity ρα suboridnate to the open
cover Uα . Then consider the n-form
X
ω= ρα ωα .
α
Theorem 4.18. Every connected orientable smooth manifold has exactly two ori-
entations.
(F ∗ ω)p (v1 , ..., vk ) = ωF (p) (dFp (v1 ), ..., dFp (vk )).
(b) F ∗ (ω ∧ η) = (F ∗ ω) ∧ F ∗ (η).
Then one can check that d defined in Rn satisfies the properties (i)-(iv). Moreover,
one can check that if F : U → V is a smooth map in Euclidean space, then we have
F ∗ (dω) = d(F ∗ ω) (4.6)
4.4. PULLBACK AND EXTERIOR DIFFERENTIATION 52
the detailed calculation is in [1, Proposition 14.23]. Now for a general manifold M ,
we first show the existence of d. Let ω ∈ Ωk (M ), and we define d on chart (U, φ) by
dω = φ∗ d(φ−1∗ ω).
This is well-defined, since for any other chart (V, ψ), the map φ ◦ ψ −1 is a diffeomor-
phism. Hence by (4.6) we have
Therefore, multiplying both sides by ψ ∗ , we get that φ∗ d(φ−1∗ ω) = ψ ∗ d(ψ −1∗ ω). It
follows quite easily from the calculation in Rn that this definition satisfies (i)-(iv).
Next, we prove uniqueness. Suppose d is any operator satisfying (i)-(iv). We
claim that if ω1 , ω2 ∈ Ωk (M ) such that ω1 = ω2 on an open set U ⊂ M , then
dω1 = dω2 on U . To see this, let p ∈ U and η = ω1 − ω2 . Let ψ ∈ C ∞ (M ) be a
bump function that is identically 1 on some neighborhood of p and supported in U .
Then ψη is identically zero. So (i)-(iv) implies that
0 = d(ψη) = dψ ∧ η + ψdη.
Evaluating this at p using ψ(p) = 1 and dψp = 0, we have that dω1 |p − dω2 |p =
dηp = 0.
P let ωI ∈ Ω (M ) be arbitrary, and (U, φ) be a smooth chart on M . Write
k
Now
ω = I ωI dx on U . For any p ∈ U , by means of a bump function we can construct
global smooth functions ω̂I , x̂i that agrees with ωI , xi in a neighborhood of p. By
virtual of (i)-(iv), we see that dω has to have the expression of (4.4) at p: More
detailed explanation can be found in [6, p.221]. Since p is arbitrary, this d must be
everywhere equal to the one we defined locally by (4.4).
F∗
Ωk (N ) Ωk (M )
d d
Ωk+1 (N ) F∗
Ωk+1 (M )
Proof. We first claim that for any f ∈ C ∞ (N ), we have F ∗ (df ) = d(f ◦ F ). This is
a simple calculation:
Fp∗ (dfF (p) )(v) = dfF (p) (dFp (v)) = (dFp (v))(f ) = v(f ◦ F ) = d(f ◦ F )p (v).
Next, since d is local, for any ω ∈ Ωk (N ), we can restrict to local charts (V, ψ, xi )
of M and (U, φ, y j ) of N such that F (V ) ⊂ U . Let
X
ω= ωj1 ···jk dxj1 ∧ · · · ∧ dxjk .
J
53 CHAPTER 4. DIFFERENTIAL FORMS
Then X
dω = dωj1 ···jk ∧ dxj1 ∧ · · · ∧ dxjk .
J
Hence
X
F ∗ (dω) = F ∗ (dωj1 ···jk ) ∧ F ∗ (dxj1 ) ∧ · · · ∧ F ∗ (dxjk )
J
X
= d(ωj1 ···jk ◦ F ) ∧ d(xj1 ◦ F ) ∧ · · · ∧ d(xjk ◦ F ).
J
Next,
X
F ∗ω = (ωj1 ···jk ◦ F ) ∧ F ∗ dxj1 ∧ · · · ∧ F ∗ dxjk
J
X
= (ωj1 ···jk ◦ F ) ∧ d(xj1 ◦ F ) ∧ · · · ∧ d(xjk ◦ F ).
J
Therefore,
X
d(F ∗ ω) = d(ωj1 ···jk ◦ F ) ∧ d(xj1 ◦ F ) ∧ · · · ∧ d(xjk ◦ F ) = F ∗ (dω).
J
Now how does the one form d(y j ◦ F ) relate to dxi ’s? By the way how 1-forms
transform (p.281 of Lee’s Smooth manifold), we have that
∂(y j ◦ ψ ◦ F ◦ φ−1 ) i ∂ F̂ j i
d(y j ◦ F ) = d(y j ◦ ψ ◦ F ) = dx = dx .
∂xi ∂xi
Therefore, we see that
!
1 n
∂ F̂ ∂ F̂
F ∗ (udy 1 ∧ · · · ∧ dy n ) = (u ◦ F ) · · · jn dxj1 ∧ · · · ∧ dxjn
∂xj1 ∂x
!
∂ F̂ j
= (u ◦ F ) det dx1 ∧ · · · ∧ dxn
∂xi
= (u ◦ F )(det dF )dx1 ∧ · · · ∧ dxn .
then
X X ∂fk
dω = dfk ∧ dxk = l
dxl ∧ dxk .
k k,l
∂x
Therefore,
X ∂fk
∂ ∂ ∂ ∂
dω i
, j = dx ∧ dx
l k
, j
∂x ∂x k,l
∂xl i
∂x ∂x
X ∂fk
= (δil δjk − δik δji )
k,l
∂xl
∂fj ∂fi
= i
− j.
∂x ∂x
One can check that this is equal to the right hand side of 4.7 applies to the vector
fields X = ∂i , Y = ∂j . This concludes the proof.
5 | Integration on Manifolds
Then, we can now write (5.1) more explicitly. Suppose that on the chart (U, φ, xi ),
we have that
ω = f dx1 ∧ · · · ∧ dxn
for some smooth function f : Supp ω ⊂ M → R. Then the pullback formula gives
(recall that on manifolds, the coordinate function xi is really xi ◦ φ):
(φ−1 )∗ ω = (φ−1 )∗ (f dx1 ∧ · · · ∧ dxn )
= (f ◦ φ−1 )d(x1 ◦ φ ◦ φ−1 ) ∧ · · · ∧ d(xn ◦ φ ◦ φ−1 )
= (f ◦ φ)dx1 ∧ · · · ∧ dxn
where the last expression is a differential form in a subset of Rn , and xi ’s are coor-
dinate functions in Rn . Therefore, using the definition of integral in Rn above, we
can rewrite (5.1) as
Z Z Z
−1 ∗
ω=± (φ ) (f dx ∧ · · · ∧ dx ) := ±
1 n
(f ◦ φ−1 )dx1 · · · dxn .
M M φ(U )
5.1. BASIC DEFINITIONS AND PROPERTIES 56
One needs to check that this definition is indepent of charts chosen. First, recall
the change of variable formula in Rn :
Lemma 5.1. Let F : U → F (U ) be a diffeomorphism between subsets of Rn . Let
J(F ) = (∂F i /∂xj ) denote the Jacobian of F . Let (y j ) denote the coordinate systems
in F (U ) and let (xi ) denote that of U . Then we have, for an integrable function f ,
Z Z
f dy · · · dy = (f ◦ F )|J(F )|dx1 · · · dxn .
1 n
F (U ) U
Proposition 5.2. The definition of integral above does not depend on which chart
we choose.
Proof. We only need to check the case for positively oriented charts, since the other
cases will follow exact same calculation, except we will get a minus sign when com-
puting Jacobian determinants. So suppose (U, φ, xi ) and (V, ψ, y j ) are two intersect-
ing charts such that Supp ω ⊂ U ∩ V. And suppose
ω = f dx1 ∧ · · · ∧ dxn = hdy 1 ∧ · · · ∧ dy n .
Then the transformation for dxi into dy j gives the relation
ω = hdy 1 ∧ · · · ∧ dy n
1
∂y ∂y n
=h · · · in ◦ φdxi1 ∧ · · · ∧ dxin
∂xi1 ∂x
i
∂y
= h det ◦ φdx1 ∧ · · · ∧ dxn
∂xj
where in the last step, we reorder every dxi1 ∧· · · dxin into dx1 ∧· · ·∧dxn , at a cost of a
sgn(σ) where σ(i1 , ..., in ) = (1, ..., n). But that with the partial derivatives combined
is exactly the determinant of the matrix (∂y i /∂xj ). Therefore, the expression above
gives i
∂y
f = det ◦ φ.
∂xj
Now, since ψ ◦ φ−1 is a positively oriented chart, we have that J(ψ ◦ φ−1 ) > 0.
Therefore, we have that
Z Z
−1
(h ◦ ψ )dy · · · dy =
1 n
h ◦ ψ −1 ◦ (ψ ◦ φ−1 )|J(ψ ◦ φ−1 )|dx1 · · · dxn
ψ(V ) φ(U )
Z
= h ◦ ψ −1 ◦ (ψ ◦ φ−1 )J(ψ ◦ φ−1 )dx1 · · · dxn
φ(U )
Z i
∂y
= h · det j
◦ φ ◦ φ−1 dx1 · · · dxn
∂x
Zφ(U )
= (f ◦ φ−1 )dx1 · · · dxn .
φ(U )
Now suppose that ω ∈ Ωnc (M ). Let (Ui , φi ) be positive charts and let (ρi ) be a
partition of unity suboridnate to the open cover (Ui ) of M . Then we may write
X X
ω= ρi ω =: ωi .
i i
Since Supp ω is compact, it intersects finitely many open cover (Ui )’s. Hence we
really have a finite sum
Xm
ω= ωi .
i=1
Then we define the integral of ω over M to be
Z Xm Z
ω= ωi .
M i=1 M
Note that we have used the fact that |J(ψ ◦ φ−1 )| = |J(T )| = 1. This proves
(a).
have
Z Z Z Z
∗ −1 ∗ ∗ −1 ∗
F ω=− [(φ ◦ F ) ] F ω = − (φ ) ω = − ω
N φ(U ) φ(U ) M
Figure 5.1: A manifold with boundary. An interior point and a boundary point
are shown.
5.2. MANIFOLDS WITH BOUNDARY AND BOUNDARY ORIENTATION 60
∂ 0 d(φ ◦ γ)i ∂
v = vi i
= γ (0) = (0) i .
∂x dt ∂x
We claim that since v ∈ / Tp (∂M ), then v n 6= 0. In fact, we will show that v ∈ Tp (∂M )
iff v n = 0: Suppose v ∈ Tp (∂M ). Then there exists a curve η : [0, ε] → M such
that η ∈ ∂M and η 0 (0) = v. Then under the chart (U, φ), we see that a portion of
η([0, ε]) near p is mapped to ∂H n . Hence (φ ◦ η)n (t) = 0 for sufficiently small t by
definition. Therefore we see that v n = d(φ ◦ η)n /dt|t=0 = 0. Conversely, if v n = 0,
then consider η̂ : [0, ε] → H n defined by η̂ i (t) = v i (t) + φn (p). Then it is easy to
check that the curve η = φ−1 ◦ η̂ : [0, ε] → M satisfies η(0) = p, η 0 (0) = v. Since
η̂ ⊂ ∂H n , we have η ⊂ ∂M . Therefore v ∈ Tp (∂M ).
61 CHAPTER 5. INTEGRATION ON MANIFOLDS
i∗∂M (N ¬ ω)p (v1 , ..., vn−1 ) = ω(di∂M (Np ), di∂M (v1 ), ..., di∂M (vn−1 ))
Proof. Denote σ = i∗∂M (N ¬ ω). Then σ ∈ Ωn−1 (∂M ). It will follow that σ is an
orientation form on ∂M if we can show it never vanishes. Given a basis (E1 , ..., En−1 )
of Tp (∂M ), the fact that Np ∈ / Tp (∂M ) implies (Np , E1 , ..., En−1 ) is a basis for Tp M .
Therefore
σp (E1 , ..., En−1 ) = ωp (Np , E1 , ..., En−1 ) 6= 0
since ωp is non-vanishing at every p, and if ωp is zero on a basis, then a multilinear-
argument would show that ωp ≡ 0. We also used the fact that i∗∂M is just restriction
to vectors tangent to ∂M . Since σp does not vanish and p is arbitrary, σ is an
orientation form. Moreover, since σ(E1 , ...En−1 ) > 0 iff ωp (Np , E1 , ..., En−1 ) > 0, the
orientation determined by σ is exactly as described in our informal statement.
It remains to show that this orientation is independent of the choice of N . Let
(x ) be coordinate of a boundary chart near p ∈ ∂M . If N and Ñ are two different
i
outward-pointing vector field on ∂M , then both (Np , ∂1 , ..., ∂n−1 ) and (Ñp , ∂1 , ..., ∂n−1 )
are bases for Tp M , and the transition matrix between them has determinant given
by N n (p)/Ñ n (p) > 0. Thus both bases determine the same orientation for Tp M .
Hence both N, Ñ determines the same orientation for Tp (∂M ).
Corollary 5.12. Let ∂H n−1 be equipped with the induced orientation when H n itself
has the standard orientation inherited from Rn . Then we can identify ∂H n with
Rn−1 . But the induced orientation agrees with the standard orientation on Rn−1
only when n is even. That is,
∂H n = (−1)n Rn−1 .
Proof. The standard orientation on Rn−1 is positively oriented for ∂H n iff the or-
dered basis [−∂n , ∂1 , ..., ∂n−1 ] is a positively oriented basis for Rn , where −∂n is an
outward-pointing vector field of H n . Note that
[−∂n , ∂1 , ..., ∂n−1 ] = −[∂n , ∂1 , ..., ∂n−1 ] = (−1)n [∂1 , ..., ∂n ].
This concludes the proof.
X
n−1 Z R Z R Z R
∂ωi
(−1) i−1
··· (x)dx1 · · · dxn
i=1 0 −R −R ∂xi
X
n−1 Z RZ R Z R
∂ωi ci · · · dxn
= (−1) i−1
··· (x)dxi dx1 · · · dx
i=1 0 −R −R ∂xi
X
n−1 Z RZ R Z R
···
i
ci
(−1)i−1 =−R dx · · · dx · · · dx = 0,
[ωi (x)]xxi =R 1 n
=
i=1 0 −R −R
because we have chosen R large enough such that ω = 0 when xi = ±R. The only
term that might not be zero is the one for which i = n. For that term we have
Z Z R Z RZ R
∂ωn
dω = (−1) n−1
··· n
(x)dxn dx1 · · · dxn−1
Hn −R −R 0 ∂x
Z R Z R
n
= (−1)n−1 ··· =0 dx · · · dx
[ωn (x)]xxn =R 1 n−1
−R −R
Z R Z R
= (−1)n ··· ωn x1 , . . . , xn−1 , 0 dx1 · · · dxn−1
−R −R
from the calculation above, where ∂H n has the induced orientation. Since φ∗ takes
outward-pointing vectors on ∂M to outward-pointing vectors on H n , it follows that
R preserving diffeomorphism on φ(U ) ∩ ∂H , and the expres-
n
φ|U ∩∂M is an orientation
sion above is equal to ∂M ω. For an interior chart, we get the same computations
with H n replaced by Rn . This proves the theorem in this case.
Finally, let ω ∈ Ωn−1
c (M ) be arbitrary. Choose a cover of Supp ω by finitely
many domains of positively oriented smooth charts (Ui )i∈I , and choose a subordinate
smooth partition of unity (ρi )i∈I . We can apply the argument to ρi ω for each i and
obtain
Z XZ XZ XZ
ω= ρi ω = d (ρi ω) = dρi ∧ ω + ρi dω
∂M ∂M M M
Z
i
! i
Z !i Z
X X
= d ρi ∧ω+ ρi dω = 0 + dω
M i M i M
P
because i ψ1 ≡ 1. This completes the proof of Stokes’ theorem.
oriented otherwise. Notice that ∂[a, b] has only two points a, b. Hence the integral
degenerates into a sum. Since the outward-pointing normal at a is a negatively
oriented basis for Ta [a, b], we assign a minus sign to a. Since the outward-pointing
normal at b is a positively oriented basis for Tb [a, b], we assign a plus sign to b, i.e.,
∂[a, b] = b − a.
Now take ω = f (x)dx. Then by Stokes’ theorem,
Z Z Z b
f (b) − f (a) = ω= dω = f 0 (x)dx.
∂[a,b] [a,b] a
while Z
ω = f (γ(b)) − f (γ(a))
∂γ([a,b])
as discussed.
Corollary 5.16 (Green’s theorem). Let D ⊂ R2 be a region and ∂D is smooth.
If P, Q are functions defined on an open region containing D and have continuous
partial derivatives there, then
Z Z
∂Q ∂P
P dx + Qdy = − dx ∧ dy.
∂D D ∂x ∂y
Proof. Choose ω = P dx + Qdy ∈ Ω1 (D), then the claim immediately follows from
Stokes’ theorem.
Corollary 5.17 (Divergence theorem). Let S ⊂ R3 be a region with smooth bound-
ary ∂S. Let F = (Fx , Fy , Fz ) : R3 → R3 be a continuously differentiable vector field.
Then Z Z
(∇ · F)dV = F · dS
S ∂S
where dV = dx∧dy ∧dz is the volume form on S, and dS = (dy ∧dz, dz ∧dx, dx∧dy)
is the area element. One can check that dS(v, w) = v × w for two vectors v, w in
R3 .
5.4. APPLICATIONS OF STOKES’ THEOREM 66
ω = F · dS = Fx dy ∧ dz + Fy dz ∧ dx + Fz dx ∧ dy.
where dS = (dy∧dz, dz ∧dx, dx∧dy) is again the area element, and dl = (dx, dy, dz)
is the curve element.
Proof. This again follows from straightforward calculation with suitable choice of
form ω = Fx dx + Fy dy + Fz dz.
Part II
• For each triple X, Y, Z ∈ Ob(C), a composition map HomC (Y, Z)×HomC (X, Y ) →
HomC (X, Z), denoted (f, g) 7→ f ◦ g.
These data must satisfy the following axioms:
(C1) For every X ∈ Ob(C), there exists a morphism idX ∈ HomC (X, X), such that
f ◦ idX = f and idX ◦g = g whenever these compositions make sense.
• Consider the category Top, where the objects are topological spaces, and for
X, Y ∈ Ob(Top), let HomTop (X, Y ) be the set of all continuous maps from
X to Y .
• Consider the category Vec, where the objects are vector spaces, and for V, W ∈
Ob(Vec), let HomVec (V, W ) be the set of all linear maps from V to W .
• Consider the category Grp, where the objects are groups, and for G, H ∈
Ob(Grp), let HomGrp (G, H) be the set of all group homomorphisms from G
to H.
• Consider the category Mfd, where the objects are smooth manifolds, and for
M, N ∈ Ob(Mfd), let HomMfd (M, N ) be the set of all smooth maps from M
to N .
Recall that when we first learned what a set is, we then went on to study functions
between sets; similarly, now we study functions between categories, called functors:
• For every pair X, Y ∈ Ob(A), a map F : HomA (X, Y ) → HomB (F (X), F (Y )),
denoted by ϕ 7→ F (ϕ).
These data should be compatible with identity and composition in the following
sense:
2. For every composable pair (f, g), we have F (f ◦ g) = F (f ) ◦ F (g). Note that
the composition f ◦ g is composition of morphisms in A, while F (f ) ◦ F (g) is
composition of morphisms in B.
f g
X−
→Y −
→Z
F (f ) F (g)
F (X) −−→ F (Y ) −−→ F (Z)
Example 6.4 (the tangent functor). The description above might seem very ab-
stract. But really this is just a fancier language that can be used to describe many
examples we have already seen. One might think of an example where, for any
smooth manifold M and a point p ∈ M , its tangent space Tp M is a vector space,
and for F : M → N smooth maps, we have a linear map dFp : Tp M → TF (p) N . And
what’s luck is that associativity is satisfied because d(G ◦ F )p = dGF (p) ◦ dFp . Does
this mean we have a functor Mfd → Vec? Almost, but not quite: This is base-point
dependent. However, we can introduce the category of pointed manifolds, de-
noted by Mfd∗ , where the objects are ordered pair (M, p), where M is a manifold
and p ∈ M is a point, and the morphisms are pointed maps F : (M, p) → (N, p0 ),
i.e., maps F : M → N with p0 = F (p). Then we get a functor Mfd∗ → Vec because
we can define the functor to take the sequence of pointed manifolds and morphisms
→ (N, p0 ) −
→ (Q, p00 )
F G
(M, p) −
But note that given a smooth map F : M → N , we can define its global differen-
tial dF : T M → T N , via (p, X) 7→ (F (p), dFp (X)). This is a smooth map between
tangent bundles. So we get a functor Mfd → Mfd, by sending the sequence
F G
M−
→N −
→Q
to the sequence
dF dG
T M −→ T N −→ T Q.
Now we define what is called a contravariant functor. It almost just behaves like
a functor, but it reverses the direction of the morphism:
Definition 6.5. Let A, B be two categories. A contravariant functor F : A → B
contains the following data:
• A map F : Ob(A) → Ob(B), denoted by A 7→ F (A).
Definition 6.9. We define the kth (de Rham) cohomology group with real
coefficients of M by the quotient of vector spaces (in particular it is a group under
vector addition)
ker d|Ωk (M ) Z k (M )
H k (M ) = = k .
im d|Ωk−1 (M ) B (M )
Therefore, this group is a quotient of closed k-forms Z k (M ) , i.e., dω = 0, by the
exact k-forms B k (M ), i.e., ω = dη for some η ∈ Ωk−1 (M ). Then
M
H(M ) := H k (M ).
0≤k≤dim M
Ωkc (M ) ∼
= Cc∞ (M ) ⊗ Ωk (M ),
In the remaining part of this chapter, we will first develop the basic theory of de
Rham cohomology. In the next chapter, we will see some beautiful applications of
cohomology and then we can appreciate how powerful it is: we can actually deduce
topological information about manifolds from cohomology.
F ∗ : H k (N ) → H k (M ).
This makes H k into a contravariant functor Top → Grp (in fact it is also a con-
travariant functor Top → Vec).
73 CHAPTER 6. COHOMOLOGY
Proof. By Proposition 4.22, we know that the pullback commutes with d. This
means that if α ∈ Ωk (N ) is a closed form, i.e., dα = 0, then F ∗ α ∈ Ωk (M ) is also
closed, because d(F ∗ α) = F ∗ (dα) = 0. Therefore F ∗ (Z k (N )) ⊂ Z k (M ). Similarly,
if β ∈ Ωk (N ) is an exact form, i.e., β = dη, then F ∗ β = F ∗ (dη) = d(F ∗ η) ∈ Ωk (M )
is also exact, hence F ∗ (B k (N )) ⊂ B k (M ).
Now, we define F ∗ : H k (N ) → H k (M ) as follows: For every [α] ∈ H k (N ), it is
represented by some closed form α ∈ Ωk (N ). We define
F ∗ ([α]) = [F ∗ α] ∈ H k (M ).
Proof. Here we will only prove the claim for differentiable paths. Piecewise differ-
entiable path can be handled similarly by doing each pieces at a time and adding
them up.
Let H : [0, 1] × [0, 1] → M be a path-homotopy from γ0 to γ1 . Denote I 2 :=
[0, 1] × [0, 1]. Since exterior derivative commutes with pullback by Theorem 4.22,
and the two vertical edges Γ2 , Γ3 of ∂I 2 are crushed into two points γ0 (0), γ0 (1) via
H (See Figure 6.1), Stokes’ theorem gives
Z Z Z
∗ ∗
H ω= d(H ω) = H ∗ (dω) = 0
∂I 2 I2 I2
since ω is closed. But on the other hand (See Figure 6.1 for the orientation of
integration),
Z 3 Z
X Z Z Z Z
∗ ∗ ∗ ∗
0= H ω= H ω= H ω− H ω= ω− ω.
∂I 2 i=0 Γi Γ1 Γ3 γ0 γ1
1 xdy − ydx
ω := .
2π x2 + y 2
Simple calculation suggests that dω = 0. Hence ω is a closed form. For each integer
k, let Ck : [0, 1] → R2 − 0 be the curve defined by Ck (t) = (cos 2πkt, sin 2πkt), i.e.,
a circle that warps around the origin k times. One can calculate and verify that
75 CHAPTER 6. COHOMOLOGY
Γ1
M
γ1
H
Γ3 I2 Γ2
γ0
Γ0
Figure 6.1: The homotopy H maps the square I 2 to the loops enclosed by γ0 , γ1 ,
maps Γ0 to γ0 and Γ1 to γ1 , collapses Γ2 to γ0 (1) = γ1 (1), and collapses Γ3 to
γ0 (0) = γ1 (0).
R
ω = k. Hence, the form ω, integrated along Ck , detects the number of times the
Ck
curve Ck goes around the origin.
Now, consider an arbitrary R piecewise differentiable loop γ : [0, 1] → M with
γ(0) = γ(1) = (1, 0). What is γ ω? It is a result in algebraic topology that γ ' Ck
for some unique integer k (which is exactlyR the number of times γ warps around the
origin). Hence by Theorem 6.14, we have γ ω = k for that k. Therefore, ω is called
the winding number form, because it detects the number of times any γ winds
around the origin.
This example is also very instructive, because of the following reason: We know
that d2 = 0. If ω is exact, i.e., ω = dη for some η ∈ Ω0 (R2 − 0), then dω = d2 η = 0,
as expected. So there are two related questions one can ask:
• Does there exists η ∈ Ω0 (R2 − 0) such that ω = dη?
Note that the use of Stokes’ theorem here is legit because η, although not necessarily
having compact support on R2 − 0, does have compact support on Ck because Ck
is compact. In fact, here we are really taking about i∗Ck η, where iCk : Ck → R2 − 0
is the inclusion map.
Therefore, immediately we know the answer to the second question is no, with
ω being a counterexample. In fact, we have H 1 (R2 − 0) ∼ = R, because R2 − 0
1 ∼
deformation retracts onto S and H (S ) = R, as we will see later.
1 1
Remark 6.18. From the previous proposition and Example 6.15, we can conclude
that R2 −0 is not simply connected. Indeed, the definition of being simply connected
means that one can continuously (even smoothly) shrink any loop γ to its basepoint
at γ(0). However, this clearly cannot be done for a loop in R2 − 0 containing
the missing origin, since there is no continuous shrinking possible because of the
missing hole. This is the reason why R2 − 0 is not simply connected, and the
reason why H 1 (R2 − 0) 6= 0. We have come to a very important observation which
shows the geometric meaning of cohomology H k (M ): the kth cohomology detects
k-dimensional holes in spaces. We will not go on and discuss why a missing point
is a 1-dimensional hole since it would be too much of a detour.
Example 6.24. In this example we compute Hc∗ (R). Since there are no k-forms for
k > 1 on R, we have Hck (R) = 0 for k > 2. So we only need to compute Hc0 (R) and
Hc1 (R).
77 CHAPTER 6. COHOMOLOGY
Thus Bc1 (R) ⊂ ker I. Conversely, if g(x)dx ∈ ker I ⊂ Ω1c (R), then it is easy to check
that Z x
f (x) = g(u)du ∈ Cc∞ (R)
−∞
and df = g(x)dx. Thus ker I ⊂ Bc1 (R). Hence we conclude that Bc1 (R) = ker I.
We also note that Zc1 (R) = Ω1c (R), i.e., every 1-form on R is closed. One inclusion
is obvious, and to see Ω1c (R) ⊂ Zc1 (R), let g(x)dx ∈ Ω1c (R). Then
d(g(x)dx) = g 0 (x)dx ∧ dx = 0.
As a final remark, note that I : Ω1c (R) → R descends to a map I : Hc1 (R) → R, still
defined by literal integration. The result ker I = Bc1 (R) suggests that the descended
integration map I : Hc1 (R) → R is an isomorphism.
such that dq ◦ dq−1 = 0 for all q. We denote a chain complex like this by A∗ .
dqA
··· Aq Aq+1 ···
Fq F q+1
Remark 6.31. It should be clear from definition that for an exact sequence A∗ , one
has H k (A∗ ) = 0 for all k.
g
2. The sequence B −
→ C → 0 is exact iff g is surjective.
f g
3. The sequence 0 → A −→B− → C → 0 is exact iff f is injective, g is surjective,
and im f = ker g. Such a sequence is called a short exact sequence.
fq gq
0 → Aq −→ B q −
→ Cq → 0
is a short exact sequence. That is, we have the following big commutative diagrams
with exact rows:
.. .. ..
. . .
dq+1
A dq+1
B dq+1
C
f q+1 g q+1
0 Aq+1 B q+1 C q+1 0
dqA dqB dqC
fq gq
0 Aq Bq Cq 0
dq−1
A dq−1
B dq−1
C
f q−1 g q−1
0 Aq−1 B q−1 C q−1 0
.. .. ..
. . .
f∗ g∗ d∗
· · · → H q (A) −→ H q (B) −
→ H q (C) −
→ H q+1 (A) → · · ·
Proof Sketch. Detailed proof (e.g., check exactness) can be found in [3], though quite
unreadable. We only define the maps f ∗ , g ∗ , d∗ here and leave the rest to the readers
who are interested. Here f ∗ : H q (A) → H q (B) is induced by the map f q : Aq → B q
and similarly g ∗ : H q (B) → H q (C) is induced by g q : B q → C q , according to Remark
6.29. We briefly review the construction of the map d∗ : H q (C) → H q+1 (A): Let
c ∈ C q with dqC (c) = 0, i.e., c represents a class in H q (C), then since g q is surjective,
there exists b ∈ B q such that g q (b) = c. Now
Therefore dqB (b) ∈ ker g q+1 = im f q+1 . Thus there exists a ∈ Aq+1 such that dqB (b) =
f q+1 (a). Thus we define
d∗ [c] = [a].
6.4. THE MAYER-VIETORIS SEQUENCE 80
Theorem 6.34 (The Five Lemma). Given a commutative diagram of Abelian groups
and group homomorphisms
f1 f2 f3 f4
··· A B C D E ···
α β γ δ ε
··· A0 f1′
B0 f2′
C0 f3′
D0 f4′
E0 ···
in which the rows are exact, if the maps α, β, δ and ε are isomorphisms, then so is
the middle one γ.
Proof. First, we prove that γ is injective. Suppose for c ∈ C we have γ(c) = 0.
Then f30 (γ(c)) = 0, and by commutativity of the diagram δ(f3 (c)) = 0. Since δ is
an isomorphism, f3 (c) = 0, and thus c ∈ ker f3 = im f2 . It follows that there exists
b ∈ B such that f2 (b) = c. Now, γ(f2 (b)) = γ(c) = 0, and by commutativity of
the diagram f20 (β(b)) = 0. So, β(b) ∈ ker f20 = im f10 , and thus there exists a0 ∈ A0
such that β(b) = f10 (a0 ). Since α is an isomorphism, there exists a ∈ A such that
α(a) = a0 . We then have β(b) = f10 (α(a)) = β(f1 (a)) by commutativity, and since
β is an isomorphism we have b = f1 (a). Now, we see that c = f2 (f1 (a)) = 0 by
exactness, so γ is indeed surjective.
Now, we prove that γ is surjective. Let c0 ∈ C 0 be arbitrary. By exact-
ness f40 (f30 (c0 )) = 0. Since δ, ε are isomorphisms, f40 = ε ◦ f4 ◦ δ −1 , so we have
ε(f4 (δ −1 (f30 (c0 )))) = 0. Since ε is an isomorphism, f4 (δ −1 (f30 (c0 ))) = 0, and thus
δ −1 (f30 (c0 )) ∈ ker f4 = im f3 , and thus there exists c ∈ C such that δ −1 (f30 (c0 )) =
f3 (c). Now, consider γ(c). We have f30 (γ(c)) = δ(f3 (c)) = f30 (c0 ), and thus c0 −γ(c) ∈
ker f30 = im f20 . Therefore, there exists b0 ∈ B 0 such that c0 − γ(c) = f20 (b0 ). Since β
is an isomorphism, there exists b ∈ B so that b0 = β(b). It follows by commutativity
of the diagram that c0 − γ(c) = f20 (β(b)) = γ(f2 (b)), and thus c0 = γ(c + f2 (b)). This
proves surjectivity, and we are done.
Here, the first map is given by i∗ (ω) = (i∗U ω, i∗V ω). Then we obtain the Mayer-
Vietoris sequence
i∗ φ
0 → Ω∗ (M ) −
→ Ω∗ (U ) ⊕ Ω∗ (V ) −
→ Ω∗ (U ∩ V ) → 0
81 CHAPTER 6. COHOMOLOGY
0 → Ω∗c (U ∩ V ) −
→ Ω∗c (U ) ⊕ Ω∗c (V ) −
→ Ω∗c (M ) → 0
δ +
where
δ(ω) = (−(jU )∗ ω, (jV )∗ ω).
6.4. THE MAYER-VIETORIS SEQUENCE 82
Proposition 6.37. The Mayer-Vietoris sequence for compact support above is ex-
act.
Proof. The map δ is injective because if δ(ω) = (−(jU )∗ ω, (jV )∗ ω) = (0, 0), then
(jU )∗ ω = 0, so ω ≡ 0 on U , in particular on U ∩ V .
Clearly, we have +(−(jU )∗ ω, (jV )∗ ω) = −(jU )∗ ω + (jV )∗ ω = 0. Thus im δ ⊂
ker +. Conversely, let (ω, τ ) ∈ Ω∗c (U ) ⊕ Ω∗c (V ) such that ω + τ = 0 on M . Then
ω = −τ on U ∩ V . Thus if we define η = ω on U ∩ V , then we have δ(η) = (ω, τ ) =
(−τ, τ ) = (−j∗ ω, j∗ ω). So ker + ⊂ im δ.
Finally, let ω ∈ Ω∗c (M ). Then ω = +(ρU ω, ρV ω). Moreover, since
which is a closed set (Supp ρU ) intersected with a compact set (Supp ω), which is
compact in a Hausdorff space, we have ρU ω ∈ Ω∗c (U ), and similarly ρV ω ∈ Ω∗c (V ).
So + is surjective. This concludes the proof.
Again by snake lemma (Theorem 6.33), we get a long exact sequence, also called
the Mayer-Vietoris sequence for compact supports, from the short exact sequence
of cochain complexes above:
∗ d
· · · → Hcq (U ∩ V ) → Hcq (U ) ⊕ Hcq (V ) → Hcq (M ) −
→ Hcq+1 (U ∩ V ) → · · ·
Remark 6.38 (Explicit description of d∗ ). Just like what we did in the ordinary
MV sequence, we give an explicit description of d∗ here. Let τ ∈ Ωqc (M ) be a closed
form representing [τ ] ∈ Hcq (M ). By exactness of the MV sequence for Ωqc ’s, there
exists an element in Ωqc (U ) ⊕ Ωqc (V ) that sums to τ , namely (ρU τ, ρV τ ). Note that
ρU τ, ρV τ indeed have compact supports (why?). Then under exterior differentiation
c (U ) ⊕ Ωc (V ). Since dτ = 0 in Ωc (U ∩ V ) by
it goes to (d(ρU τ ), d(ρV τ )) ∈ Ωq+1 q+1 q+1
6.4.2 Cohomology of S 1
We now calculate H ∗ (S 1 ), H ∗ (S 1 ) using the Mayer-Vietoris sequence. We do the
regular cohomology first.
By Example 6.22, and the fact that S 1 is connected, we see that H 0 (S 1 ) = R. By
Example 6.19, we know that H k (S 1 ) = 0 for k > 1. Hence it remains to calculate
H 1 (S 1 ). We cover S 1 by two open sets U, V as suggested by Figure 6.2:
Now the Mayer-Vietoris sequence reads
δ d∗
H 0 (S 1 ) → H 0 (U ) ⊕ H 0 (V ) −
→ H 0 (U ∩ V ) −
→ H 1 (S 1 ) → H 1 (U ) ⊕ H 1 (V ) → · · ·
`
Since U, V ∼ = R, U ∩ V ∼= R R, using H 0 (R) = R, H 1 (R) = 0, we get that the MV
sequence reads:
δ d∗
R→R⊕R− →R⊕R− → H 1 (S 1 ) → 0.
83 CHAPTER 6. COHOMOLOGY
U S1
V
Figure 6.2: Choose U to be an open subset of S 1 that contains the upper hemi-
sphere, and V an open subset that contains the lower hemisphere, such that U ∩ V
is homeomorphic to two disjoint copies of intervals.
R⊕R R⊕R ∼
H 1 (S 1 ) = ∗
= = R.
ker d im δ
Remark 6.39 (A generator for H 1 (S 1 )). We now find an explicit generator of the
group H 1 (S 1 ) ∼
= R. Since H 1 (S 1 ) = R2 / ker d∗ , anything that is not killed by d∗ in
H 0 (U ∩ V ) will represent a generator. Hence, let α ∈ Ω0 (U ∩ V ) be a closed 0-form
(i.e., locally constant function on U ∩ V ) that is not in im δ = ker d∗ , then d∗ α will
represent a generator of H 1 (S 1 ). For example, if we take α to be a locally constant
function that is 1 on one connected piece of U ∩ V , and 0 on the other, then it will
represent the element (1, 0) ∈ H 0 (U ∩ V ), which is clearly not in im δ.
Now by Remark 6.36, we have
(
−d(ρV α) on U
d∗ α =
d(ρU α) on V
Proof. By Example 6.24, for any open set WR homeomorphic to R, we have Hc1 (W ) ∼ =
R, with the isomorphism defined by α 7→ W α. In other words, while making R R the
identification H 1 (U ∩V ) ∼ = R2 , we are really using the isomorphism ω 7→ ( A ω, B ω),
where A, B are the two connected components
R Rof U ∩ V . Now Rthe image
R δ(ω) =
(−(jU )∗ ω, (jV )∗ ω) is identified with ( U −(jU )∗ ω, V (jV )∗ ω) = (− U ∩V ω, U ∩V ω). in
R2 . Now putting all the identification together, and viewing δ as a map R2 → R2 ,
the map is actually
δ : (λ, µ) 7→ (−λ − µ, λ + µ)
which makes it clear that dim im δ = 1.
Hc0 (S 1 ) ∼
= im ϕ = ker δ ∼
=R
by the previous proposition and the rank-nullity theorem. And since Hc1 (U ) ⊕
Hc1 (V ) → H 1 (S 1 ) is a surjection by exactness,
R⊕R ∼
Hc1 (S 1 ) = = im δ ∼
=R
ker δ
by the 1st isomorphism theorem and the previous proposition.
Let [α] ∈ H 1 (R2 − 0), which is represented by a closed form α ∈ Ω1 (R2 − 0). Take
any smooth closed loop γ in R2 − 0 and define
Z
ψ([α]) := α.
γ
Example 6.46. By the example above, we may consider another interesting appli-
cation of the Mayer-Vietoris sequence. The problem is to calculate H ∗ (Ps ), where
Ps := R2 − {p1 , ..., ps }
is the plane with s distinct points removed. Clearly H k (Ps ) = 0 for k > 2 because
of the dimension, and H 0 (Ps ) = R because Ps is connected. Hence the problem
is really about the calculation of H 1 (Ps ). We choose U, V in the Mayer-Vietoris
sequence as Figure 6.3 suggests:
U∼
= Ps−1 V ∼
= P1
p1 p2
ps
ps−1
Figure 6.3: The relative position of p1 , ..., ps does not matter since cohomology
is invariant under homeomorphisms. So we may choose U, V as suggested, with
U∼ = Ps−1 , V ∼
= P1 , U ∩ V ∼
= R2 .
H 1 (Ps )/ ker α ∼
= Rs
we conclude that dim im d∗ = 0. Hence dim ker α = 0. This implies that ker α = 0
and
H 1 (Ps ) ∼
= Rs
which proves the claim. Similar to the previous remark, we can construct an explicit
isomorphism ψ : H 1 (Ps ) → Rs , by choosing s smooth closed loops γ1 , ..., γs , each
only contains p1 , ..., ps Rin their Rinterior (ref. Jordan curve theorem), respectively.
Then define ψ([α]) = ( γ1 α, ..., γs α).
H n+1 0 0 0
Hn ? 0 0
H n−1 ? 0 R
H n−2 ? 0 0
.. .. .. ..
. . . .
H1 ? 0 0
H0 ? i
R⊕R δ
R
H n (S n ) = R,
H k (S n ) = 0 for k = 2, · · · , n − 1,
Such a family of operators is called a chain homotopy operator. Note that (*)
implies that π ∗ s∗ is equal to identity on H q (Rn+1 ), although it is not identity on
Ω∗ (Rn+1 ). Because any element [α] ∈ H q (Rn+1 ) is represented by a closed form α,
and exact forms are set to be zero in H q (Rn+1 ). Hence the two terms on the right
hand side of (*) will both be zero when we descend to H q (Rn+1 ). This will finish
the proof.
It remains to construct the family of operators K q : Ωq (Rn+1 ) → Ωq−1 (Rn+1 )
described above. If we let x1 , ..., xn , t be the coordinate on Rn+1 , the we note that
89 CHAPTER 6. COHOMOLOGY
for any form in Ωq (Rn+1 ), it either contains the differential of the last coordinate dt
in the wedge, or it does not. We define K q linearly as follows: If the input does not
contain dt, i.e., the input is of the form
then define
K q (α) := 0.
If the input contains dt, i.e., the input is of the form
then define Z t
K (β) := dx ∧ · · · ∧ dx
q i1 iq−1
· f (x, λ)dλ.
0
We now check that (*) holds for forms α, β respectively. Since elements in Ωq (Rn+1 )
are linear combinations of elements of the form α, β, we are done since K q ’s are
defined linearly. We first check (*) holds for α. By definition of π, we have s∗ (f ) =
f ◦ s, and π ∗ (dxa ) = dxa , s∗ (dxa ) = dxa , and so on. Therefore,
by the fundamental theorem of calculus. This proves that (*) holds on forms of type
α. Now we check that (*) holds on forms of type β. Note that since s∗ dt = d(t ◦ s) =
d(0) = 0, we have
(id −π ∗ s∗ )β = β.
Now, we have
∂f a
K q+1
dβ = K q+1
dx ∧ dx ∧ · · · ∧ dx
i1 iq−1
∧ dt
∂xa
Z t
∂f
= (−1) dx ∧ · · · ∧ dx
q−1 i1 iq−1
∧ dx ·
a
a
(x, λ)dλ
0 ∂x
6.6. POINCARÉ LEMMAS 90
and
Z t
dK β = d dx ∧ · · · ∧ dx
q i1 iq−1
· f (x, λ)dλ
0
Z t
∂
= f (x, λ)dλ · dxa ∧ dxi1 ∧ · · · ∧ dxiq−1
∂xa 0
Z
∂ t
+ f (x, λ)dλ · dt ∧ dxi1 ∧ · · · ∧ dxiq−1
∂t 0
Z t
∂f
= a
dt · (−1)q−1 dxi1 ∧ · · · ∧ dxiq−1 ∧ dxa + f (x, t)dt ∧ dxi1 ∧ · · · ∧ dxiq−1
∂x
Z0 t
∂f
= a
dt · (−1)q−1 dxi1 ∧ · · · ∧ dxiq−1 ∧ dxa + (−1)q−1 β.
0 ∂x
Therefore,
dK q β − K q+1 dβ = (−1)q β
which again proves (*).
Remark 6.50 (General Poincaré Lemma). In fact, let M be any smooth manifold,
and π : M ×R → M be the projection, and s : M → M ×R be the zero section. The
proof of the Poincaré lemma carries over and shows that π ∗ , s∗ gives an isomorphism
H ∗ (M × R) ∼
= H ∗ (M ).
The construction of the isomorphisms are more difficult, hence we will talk
through the proof for the remainder of this section.
To prove this theorem, again consider π : M × R → M , the projection map. But
π ω ∈ Ω∗ (M × R) clearly has noncompact support, for example for a compactly
∗
π∗ : Ω∗c (M × R) → Ω∗−1
c (M )
91 CHAPTER 6. COHOMOLOGY
by again only caring about forms with dt in it: Give M × R the product charts of M
and R, with t being the coordinate on R. Every ω ∈ Ωqc (M × R) is of the following
two types: Either
ω = (π ∗ ϕ)f (x, t) ϕ ∈ Ωqc (M ), f ∈ Cc∞ (M × R)
or
ω = (π ∗ ϕ) ∧ f (x, t)dt ∞
c (M ), f ∈ Cc (M × R).
ϕ ∈ Ωq−1
Then we define π∗ as follows, for two types of forms respectively:
π∗ : (π ∗ ϕ)f (x, t) 7→ 0
Z ∞
∗
π∗ : (π ϕ) ∧ f (x, t)dt 7→ ϕ f (x, t)dt.
−∞
We will need to show that π∗ d = dπ∗ . So that this gives a well-defined map
π∗ : Hc∗ (M × R) → Hc∗−1 (M ).
Lemma 6.52. π∗ d = dπ∗ .
Proof. To prove the equality, we just need to show that it holds for both types of
forms. First let ω ∈ Ωck (M × R) be of “type 1”, i.e. ω = (π ∗ ϕ)f for some ϕ ∈ Ωk (M )
and some compactly supported smooth f : M × R → R. Then
d(π∗ ω) = d(0) = 0,
and
π∗ (dω) = π∗ (d(π ∗ ϕ) ∧ f + (−1)k (π ∗ ϕ) ∧ df )
= π∗ (d(π ∗ ϕ) ∧ f ) + (−1)k π∗ ((π ∗ ϕ) ∧ df )
= π∗ (π ∗ (dϕ) ∧ f ) + (−1)k π∗ ((π ∗ ϕ) ∧ df )
= (−1)k π∗ ((π ∗ ϕ) ∧ df )
where the last equality holds by definition of π∗ . Therefore, we just need to show
that π∗ ((π ∗ ϕ) ∧ df ) = 0. In local coordinates (x1 , · · · , xn , t), we have
∗ ∗ ∂f 1 ∂f n ∂f
π ϕ ∧ df = π ϕ ∧ dx + · · · + n dx + dt
∂x1 ∂x ∂t
∗ ∗ ∂f
=π ϕ∧ π η+ dt
∂t
∂f
= π ∗ ϕ ∧ π ∗ (η) + π ∗ ϕ ∧ dt
∂t
∂f
= π ∗ (ϕ ∧ η) + π ∗ ϕ ∧ dt,
∂t
where η ∈ Ω1 (M ) is given by η = d(f ◦ s), with s : M → M × R being the zero
section. Clearly η is globally defined, and (∂f /∂t)dt is globally defined when a
product atlas is used. Noting that f is compactly supported, we then have
Z ∞
∗ ∂f
π∗ (π ϕ ∧ df ) = ϕ dt = ϕ[f (x, ∞) − f (x, −∞)] = 0
−∞ ∂t
6.6. POINCARÉ LEMMAS 92
as desired. This establishes the equality dπ∗ ω = π∗ dω in the case that ω is of type
1.
Now assume ω is of type 2, i.e. ω = (π ∗ ϕ)f dt for some ϕ ∈ Ωk−1 (M ) and
some compactly supported smooth f : M × R → R. Then, in local coordinates
(x1 , ..., xn , t), we have
Z ∞
d(π∗ ω) = d ϕ f dt
−∞
Z ∞ Z ∞
= dϕ ∧ k−1
f dt + (−1) ϕd f dt
−∞ −∞
Z ∞ Z ∞
∂f
= dϕ ∧ f dt + (−1) ϕ ∧ dx
k−1 a
a
dt
−∞ −∞ ∂x
while
∗ ∗ ∂f a
π∗ (dω) = π∗ d(π ϕ) ∧ f dt + (−1) (π ϕ) ∧
k−1
dx ∧ dt
∂xa
Z ∞
k−1 ∗ ∂f a
= dϕ ∧ f dt + π∗ (−1) π ϕ ∧ a dx ∧ dt
−∞ ∂x
Z ∞
k−1 ∗ a ∂f
= dϕ ∧ f dt + π∗ (−1) π (ϕ ∧ dx ) a dt
−∞ ∂x
Z ∞ Z ∞
∂f
= dϕ ∧ f dt + (−1)k−1 (ϕ ∧ dxa ) a
dt
−∞ −∞ ∂x
Z ∞ Z ∞
∂f
= dϕ ∧ f dt + (−1) ϕ ∧ dx
k−1 a
a
dt
−∞ −∞ ∂x
= d(π∗ ω).
Now we need to define another map e∗ in the reverse direction, and show that
they are mutual inverses in the cohomology level. We define, on forms,
e∗ : Ω∗c (M ) → Ω∗+1
c (M × R)
R
as follows: Let e = e(t)dt ∈ Ω1c (R) such that R e = 1. Then define e∗ : Ω∗c (M ) →
Ω∗+1
c (M × R) by
e∗ : ϕ 7→ (π ∗ ϕ) ∧ e.
It is easy to see that de∗ = e∗ d. Thus e∗ descends to a well-defined map on the
cohomology groups. It follows directly from definition that
π∗ e∗ = idΩ∗c (M ) .
K : (π ∗ ϕ)f (x, t) 7→ 0
93 CHAPTER 6. COHOMOLOGY
and
Z t Z ∞
∗ ∗
K : (π ϕ) ∧ f (x, t)dt 7→ π ϕ f (x, λ)dλ − A(t) f (x, λ)dλ
−∞ −∞
where Z Z t
A(t) = e= e(λ)dλ.
(−∞,t) −∞
Finally, we conclude the proof of Theorem 6.51 because this shows that e∗ , π∗
are isomorphisms that are inverses of each other on the level of cohomology groups,
proving the Poincaré lemma for compactly supported cohomology.
Proof. Again we check the statement on the two types of forms. First, let
because ∂f /∂t(x, ±∞) = 0 by the assumption that f ∈ Cc∞ (M × R). Therefore, the
claim holds for forms of type 1. Now let
Now we calculate (dK −Kd)ω, which is a bit tedious because it involves cancellation
of many terms. First, for simplicity, we put
Z t Z ∞
β(x, t) := f (x, λ)dλ − A(t) f (x, λ)dλ.
−∞ −∞
6.6. POINCARÉ LEMMAS 94
Then we have
dKω = d(π ∗ ϕβ(x, t))
∗ ∂β a ∂β
= π dϕ · β(x, t) + (−1) ϕ ∧ a dx + (−1) ϕ ∧
q−1 q−1
dt
∂x ∂t
Z t Z ∞
∗ ∗
= π dϕ f (x, λ)dλ − π dϕA(t) f (x, λ)dλ
−∞ −∞
Z t Z ∞
q−1 ∗ a ∂ q−1 ∗ a ∂
+ (−1) π ϕ ∧ dx f (x, λ)dλ − (−1) π ϕA(t)dx f (x, λ)dλ
∂xa −∞ ∂xa −∞
Z t Z ∞
q−1 ∗ ∂
+ (−1) π ϕ ∧ dt f (x, λ)dλ − A(t) f (x, λ)dλ .
∂t −∞ −∞
Furthermore,
∗ q ∗ ∂f a
Kdω = K π dϕ ∧ f (x, t)dt + (−1) π ϕ ∧ a dx ∧ dt
∂x
Z t Z ∞
∗
= π dϕ f (x, λ)dλ − A(t) f (x, λ)dλ
−∞ −∞
Z t Z ∞
q−1 ∗ ∂f ∂f
+ (−1) π ϕ ∧ dx a
a
(x, λ)dλ − A(t) a
(x, λ)dλ
−∞ ∂x −∞ ∂x
Z t Z ∞
∗ ∗
= π dϕ f (x, λ)dλ − π dϕA(t) f (x, λ)dλ
−∞ −∞
Z t Z ∞
q−1 ∗ a ∂ q−1 ∗ a ∂
+ (−1) π ϕ ∧ dx f (x, λ)dλ − (−1) π ϕA(t)dx f (x, λ)dλ.
∂xa −∞ ∂xa −∞
Hence we see that
Z t Z ∞
q−1 ∗ ∂
(dK − Kd)ω = (−1) π ϕ ∧ dt f (x, λ)dλ − A(t) f (x, λ)dλ
∂t −∞
Z−∞∞
= (−1)q−1 π ∗ ϕ ∧ f (x, t)dt − (−1)q−1 π ∗ ϕ ∧ dt · f (x, λ)dλ · e(t)
−∞
Z ∞
∗
= (−1) ω − (−1)
q−1 q−1
π ϕ· f (x, t)dt ∧ e = (−1)q−1 (id −e∗ π∗ )ω.
−∞
Theorem 6.55. Every manifold has a good cover. If the manifold is compact, then
the cover may be chosen to be finite.
Proof. Endow a smooth manifold with a Riemannian metric, which is possible for
any smooth manifold: see [1], Proposition 13.3. Then every point on the manifold
has a geodesically convex neighborhood, and intersection of geodesically convex
neighborhoods is again geodesically convex (see [4], p.72, Theorem 3.7 and p.76,
Proposition 4.2).
Let U = (Uα )α∈I , U0 = (Uβ0 )β∈J be two open covers of M . If for all β ∈ J,
Uβ0 ⊂ Uα for some α ∈ I, then we say U0 is a refinement of U, and we write U < U0 .
Corollary 6.56. Every open cover on a manifold has a refinement which is a good
cover.
Definition 6.57. A directed set is a set I with a relation < such that
(c) For any a, b ∈ I, there exists c ∈ I such that a < c and b < c.
Corollary 6.60. The good covers are cofinal in the set of all covers of a manifold
M.
The existence of good covers on manifolds gives a powerful tool of proving many
results by using the Mayer-Vietoris sequence and five lemma, inducting on the car-
dinality of a good cover. This proof method is called the Mayer-Vietoris argument.
We will prove several statements using the Mayer-Vietoris argument. But in the
proof we will always assume that our manifold is of finite type, i.e., admits a fi-
nite good cover (for example, take any compact manifolds). Many statements that
we will prove are much more general and holds for all smooth manifolds (and even
topological spaces), but then the proof will require much more work. Here, we are
willing to sacrifice generality for more concise and elegant proofs.
6.7. THE MAYER-VIETORIS ARGUMENT 96
Here, we are slightly abusing notation and use α instead of [α] to denote a cohomol-
ogy class in H q (M ). This map is well-defined: If α = α0 ∈ H q (M ), i.e., α − α0 = dη
for some η ∈ Ωq−1 (M ), then
Z Z Z Z
0
(α − α ) ∧ β = dη ∧ β = d(η ∧ β) − (−1) deg η
η ∧ dβ
M M Z M Z M
= η ∧ β − (−1) deg η
η ∧ dβ = 0.
∂M M
In the second to last equality, we used Stokes’ theorem in the first term, which we
can because η ∧ β ∈ Ωn−1
c (M ), and since ∂M = ∅, the first term is zero. The second
term is also zero because β is closed since it represents a cohomology class. Similarly
one can show that if β = β 0 ∈ Hcn−q (M ), the resulting pairing is also well-defined.
Theorem 6.65 (Poincaré
R Duality for Oriented Manifold). Let M n be an
R oriented
manifold. The paring : H (M ) ⊗ Hc (M ) → R defined by α ⊗ β 7→ M α ∧ β is
q n−q
r − d∗
· · · → H q (U ∪ V ) H q (U ) ⊕ H q (V ) H q (U ∩ V ) H q+1 (U ∪ V ) → · · ·
∫ ∫ ∫ ∫ ∫
α7→ U ∪V
α∧(•) (σ,τ )7→ U
σ∧(•)+ V
τ ∧(•) ω7→ U ∩V
ω∧(•) η7→ U ∪V
η∧(•)
+∗ δ∗ (d∗ )∗
· · · → Hcn−q (U ∪ V )∗ Hcn−q (U )∗ ⊕ Hcn−q (V )∗ Hcn−q (U ∩ V )∗ Hcn−q−1 (U ∪ V )∗ → · · ·
Here, r and − are restriction and difference maps defined in the MV sequence,
and d∗ is the boundary map defined in the MV sequence. The map δ : Hcn−q (U ∩
V ) → Hcn−q (U ) ⊕ Hcn−q (V ) is the signed inclusion δ(ω) = (−j∗ ω, j∗ ω), and + :
Hcn−q (U ) ⊕ Hcn−q (V ) → Hcn−q (U ∪ V ) is the summation map.
6.7. THE MAYER-VIETORIS ARGUMENT 98
• For the first square, let α ∈ RH q (U ∪ V ). We first go down and then go left.
Denote the linear functional U ∪V α ∧ (•) by α̂. By definition of dualization of
a linear map,
Z
∗
+ (α̂)(β1 , β2 ) = α̂(+(β1 , β2 )) = α ∧ (β1 + β2 )
U ∪V
• For the second square, we again first go down and then go left. Let (σ, τ ) ∈
H q (U ) ⊕ H q (V ). The linear functional we get in Hcn−q (U ∩ V )∗ , recalling the
definition of dualization, will be defined as follow: for any α0 ∈ Hcn−q (U ∩ V ),
it gets mapped to
Z Z Z Z
0 0 0
σ ∧ (−j∗ α ) + τ ∧ (j∗ α ) = − σ ∧ α + τ ∧ α0 .
U V U V
Now if we first go left and go down in the third square, we would get (note
that Supp(ω) ⊂ U ∩ V ):
Z Z
∗
d ω∧τ =− (dρV ) ∧ ω ∧ τ
U ∪V U ∩V
Hence we see that the result we get from two different paths only differ by a
sign, i.e., differ by (−1)deg ω+1 .
99 CHAPTER 6. COHOMOLOGY
Remark 6.67. There are still two quick remarks I want to make about the proof
of the previous lemma, and how it helps us to prove Poincaré duality:
• There is one small detail: We know MV sequence for Hc∗ is exact, but we do
not know yet when we dualize everything the sequence remain exact. Luckily
in this case, it is.
Reason: The functor HomR (−, M ) is exact iff M is an injective R-module.
Given a field k, every k-vector space Q is an injective k-module. See details
in the example section of Wikipedia. In particular, HomR (−, R) is exact.
• So we know from lemma 6.66 that the diagram there is sign commutative.
But to prove Poincaré duality we want to use the Five Lemma. But the
Five Lemma requires the diagram to be commutative instead of just sign
commutative. The trick here is to fix the vertical map a little bit so that
everything becomes commutative: Suppose we have a diagram with exact
rows and sign commutative squares, with vertical maps being isomorphisms
as below, but say the first square is only commutative up to a minus sign:
f g
··· A B C ···
α β γ
··· A0 f′
B0 g′
C0 ···
this map is Rsurjective: Take any φ ∈ Hcn (Rn )∗ and ω ∈ Hcn (Rn )Ra generator of R the
group, i.e., Rn ω = 1. Then say φ(ω) = λ ∈ R. Then the map (λ) : α 7→ λ Rn α
is also a linear functional on Hcn (Rn ) that agrees with φ on the generator ω, hence
they are equal. This shows the integration map is a surjective linear map between
vector spaces of same dimension, hence is an isomorphism.
Now induct on the cardinality of a good cover. Suppose Poincaré duality holds
for any manifold having a good cover with at most p. Let M be a manifold with
p + 1 good cover U0 , ..., Up . Now (U0 ∪ · · · ∪ Up−1 ) ∩ Up is a manifold with a good
6.7. THE MAYER-VIETORIS ARGUMENT 100
Künneth Formula
Theorem 6.69 (Künneth Formula). Let M, F be two manifolds, then for every
nonnegative integer n, one has
M
n
H (M × F ) ∼
n
= H p (M ) ⊗ H n−p (F ).
p=0
H ∗ (M × F ) ∼
= H ∗ (M ) ⊗ H ∗ (F ),
L
where H ∗ (M ) = ∞ p ∗
p=1 H (M ) denote the cohomology ring, and similarly for H (F )
and H ∗ (M × F ).
Proof. We again prove this theorem for manifolds with finite good cover, using the
beautiful MV argument. There are two obvious projection maps
π :M ×F →M
ρ : M × F → F.
ψ : H ∗ (M ) ⊗ H ∗ (F ) → H ∗ (M × F )
ω ⊗ ϕ 7→ π ∗ ω ∧ ρ∗ ϕ
The trick is to put this exact sequence on top, and the MV exact sequence of the
pair (U × F, V × F ) on the bottom, match them up using ψ componentwise, and
show that we have a commutative diagram
Ln Ln Ln
p=0 H p (U ∪ V ) ⊗ H n−p (F ) p=0 (H
p
(U ) ⊕ H p (V )) ⊗ H n−p (F ) p=0 H p (U ∩ V ) ⊗ H n−p (F )
ψ ψ ψ
H ((U ∪ V ) × F )
n
H (U × F ) ⊕ H (V × F )
n n
H ((U ∩ V ) × F )
n
L Ln
Note that here instead of writing p∈Z we write p=0 . This is because when
p < 0 or p > n, one of the two groups we are tensoring is the trivial group, so
the direct sum becomes a finite direct sum with only nontrivial cohomology groups
tensored together.
The commutativity of the first and the second square commutes is left as an
exercise. The first one is easy once all the relevant maps are written out and we
can just chase the diagram. The second square commutes because we are using the
isomorphism (See Remark 6.70)
M
n
∼
=
M
n
(H (U ) ⊕ H (V )) ⊗ H
p p n−p
(F ) −
→ (H p (U ) ⊗ H n−q (F )) ⊕ (H p (V ) ⊗ H n−q (F ))
p=0 p=0
n−p ∼
=
(αp , β p ) ⊗ τ −
→ (αp ⊗ τ n−p , β p ⊗ τ n−p )
for αp ∈ H p (U ), β p ∈ H p (V ), τ n−p ∈ H n−p (F ).
Accepting this, then the commutativity of the second square can be verified
easily. Now we verify commutativity of the third square
Ln ∗ Ln
p=0 H (U ∩ V ) ⊗ H (F ) d (U ∪ V ) ⊗ H n−p (F )
p n−p p+1
p=−1 H
ψ ψ
d∗
H n ((U ∩ V ) × F ) H n+1 ((U ∪ V ) × F )
6.7. THE MAYER-VIETORIS ARGUMENT 102
L L
Note that in the second term on the right, the sum p∈Z actually becomes np=−1 .
It suffices to verify the commutativity of the diagram for pure tensors (Remark
6.71). Let ω ⊗ ϕ ∈ H p (U ∩ V ) ⊗ H n−p (F ). Then
ψd∗ (ω ⊗ ϕ) = π ∗ (d∗ ω) ∧ ρ∗ ϕ
d∗ ψ(ω ⊗ ϕ) = d∗ (π ∗ ω ∧ ρ∗ ϕ).
d∗ (π ∗ ω ∧ ρ∗ ϕ) = d((π ∗ ρU )π ∗ ω ∧ ρ∗ ϕ)
= (dπ ∗ (ρU ω)) ∧ ρ∗ ϕ
= π ∗ (d∗ ω) ∧ ρ∗ ϕ.
Remark 6.70. tensor product distributes over direct sum is a standard result in
algebra. A detailed proof of this isomorphism can be found in this post. To get an
idea, we can do a simple case and prove (A ⊕ B) ⊗ C ∼ = (A ⊗ C) ⊕ (B ⊗ C). The
isomorphism can be constructed via φ : (A ⊕ B) ⊗ C → (A ⊗ C) ⊕ (B ⊗ C) where
φ : (a, b) ⊗ c 7→ (a ⊗ c, b ⊗ c).
ψ : (a ⊗ c1 , b ⊗ c2 ) 7→ (a, 0) ⊗ c1 + (0, b) ⊗ c2 .
One can verify that both φ, ψ are well-defined homomorphisms that are inverses of
each other. For example
φψ(a⊗c1 , b⊗c2 ) = φ((a, 0)⊗c1 +(0, b)⊗c2 ) = (a⊗c1 , 0)+(0, b⊗c2 ) = (a⊗c1 , b⊗c2 ).
The key here is to realize that (A ⊕ B) ⊗ c does not only consists of elements of the
form (a, b) ⊗ c, but linear combinations of elements like this.
Remark 6.71. Finally, I shall remark that for our purpose, it suffices to check the
diagram above commutes for elements with the same coefficient in H n−q (F ), i.e.,
elements of the form (αp ⊗ τ n−p , β p ⊗ τ n−p ). This is because for any random element,
it can be decomposed into elements with the same H n−q (F ) coefficients:
Therefore by linearity, it suffices to check the diagram commutes for elements with
the same H n−q (F ) coefficients.
103 CHAPTER 6. COHOMOLOGY
for all k. We prove this statement by induction on n. First, the claim obviously
holds for the base case where n = 1, by Section 6.4.2. Next, we assume that
H k (T n−1 ) = R( )
n−1
k
the term in the direct sum containing it will be the trivial group, which does not
contribute to H k (T n ). And since any finite tensor products of R is isomorphic to R,
the number of nontrivial factors of H k (T n ) in the direct sum above corresponds to
the number of distinct ways one
n
can write i1 + · · · + in = k, where i1 , ..., in ∈ {0, 1}.
And the answer of course is k . This is why the binomial coefficient appears in this
problem, because it is just a combinatorial problem in disguise.
Poincaré duality:
(Hck (M ))∗ ∼= H n−k (M ).
R
Recall that the isomorphism H n−k (M ) → (Hck (M ))∗ is given by η 7→
R M (•) ∧ ω up
to a sign. Therefore there exists a unique ηS ∈ H (M ) such that M (•) ∧ ηS = IS .
n−k
Proposition 6.73. When S is compact, its Poincaré dual may be taken to have
compact support.
Proof. Now Rsuppose M n has a finite good cover, and S k is compact. R Then since S
is compact, S i∗S ω < ∞ for any ω ∈ Ωk (M ). Thus this time ω 7→ S i∗S ω is a map
IS : H k (M ) → R. Now since M has a finite good cover, its de Rham cohomology
(with compact support) are finite-dimensional vector spaces, so Poincaré duality
H k (M ) ∼= (Hcn−k (M ))∗ implies that Hcn−k
R (M ) ∼
= (H k (M ))∗ . The isomorphism
Hcn−k (M ) → (H k (M ))∗ is given by η 7→ M (•) ∧ η Rup to a sign. Therefore, there
exists a unique ηS0 ∈ Hcn−k (M ) such that IS (ω) = M ω ∧ ηS0 , for all ω ∈ H k (M ).
That is, Z Z
i∗S ω = ω ∧ ηS0 .
S M
ϕα ◦ ϕ−1
β : (Uα ∩ Uβ ) × R → (Uα ∩ Uβ ) × R
n n
are vector space automorphisms of Rn in each fiber and hence give rise to maps
Proof. Let (Uα , ψα )α be an oriented atlas for M with transition functions hαβ =
ψα ◦ ψβ−1 . Let ϕα : π −1 (Uα ) → Uα × Rn be the family of local trivilizations for E
with transition functions gαβ . Then the composition
ϕα ψα ×id
π −1 (Uα ) −→ Uα × Rn −−−→ Rm × Rn
which is given by
σαβ (x, y) = (hαβ (x), gαβ (ψβ−1 (x))y).
Then its Jacobian is given by
J(hαβ ) ∗
J(σαβ ) =
0 gαβ (ψβ−1 (x))
Hck+n (E) ∼
= Hck (M ).
Proof. By Lemma 6.77, we know that E is also an oriented manifold, and by Corol-
lary 6.80, we know that E is of finite type. Then the statement can be proven by
using Poincaré duality twice, and H ∗ (E) ∼
= H ∗ (M )
Hck+n (E) ∼
= (H m+n−k (E))∗ ∼
= (H m+n−k (M ))∗ ∼
= Hck−n (M ).
One can easily see that exterior differentiation preserves forms with vertical compact
support. Hence we define the kth cohomology of E with compact vertical
support
k ker(d : Ωkcv (E) → Ωk+1
cv (E))
Hcv (E) := .
im(d : Ωcv (E) → Ωkcv (E))
k−1
Now we define a map, called integration along fiber, similar to what we did
in order to prove Poincaré Lemma.
cv (E) → Ω (M ) as follows:
Definition 6.84. Define π∗ : Ωk+n k
where ζα again contains less than n fiber differential in its wedges. Then define
−1
cv (π (Uα ) → Ω (Uα ) via
π∗ : Ωk+n k
Z
(π∗ ωα )(p) = ϕ(p) · f (x, t)dt1 ∧ · · · ∧ dtn
Ep
for ρ ∈ C ∞ (Uα ) and f has compact support on every fiber of E. Then on π −1 (Uα ) ∩
π −1 (Uβ ), note that since π ∗ dxi = dxi , one has
∂xi1 ∂xik p1
τ =ρ · · · dy ∧ · · · ∧ dy pn
∂y p1 ∂y pk
and
∂t1 ∂tn q1
gdu1 ∧ · · · ∧ dun = f · · · du ∧ · · · duqn
∂uq1 ∂u
qn
∂tµ
= f det du1 ∧ · · · ∧ dun
∂uν
where in the second equality, we used the argument in the proof of (4.3). Now for a
fixed point p on the intersection Uα ∩ Uβ , we have, by the change of variable formula
of integration,
Z
π∗ ωα (p) = ρdx ∧ · · · ∧ dx (p) ·
i1 ik
f dt1 ∧ · · · ∧ dtn
E
Z p
= ρdxi1 ∧ · · · ∧ dxik (p) · f ◦ ϕ−1α (p, t)dt · · · dt
1 n
R
µ
n
Z
∂xi1 ∂xik p1 −1 ∂t
= ρ p1 · · · p dy ∧ · · · ∧ dy (p) · pn
f ◦ ϕβ (p, u) det ν
du1 · · · dun
∂y ∂y k R ∂u
Z
n
= τ (p) · g ◦ ϕ−1
β (p, u)du · · · du
1 n
ZR
n
µ
∂t
det = det(gβα ) > 0
∂uν
by assumption, hence this gives the 4th equality. It remains to prove that π∗ is
independent of trivilizations. Let
ϕ0β : π −1 (Vβ ) → Vβ × Rn
Proof. Let (Uα , ϕα ) be a trivilization for E, and let ρα be a partition of unity sub-
ordinate to Uα (note that this time Uα ’s are actually on E instead of M ). Since for
109 CHAPTER 6. COHOMOLOGY
P
any ω ∈ Ωk+n
cv (E), we have ω = α ρα ω, and both π∗ and d are linear, it suffices to
prove the proposition for ρα ω ∈ Ωk+n
cv (Uα ). This means that we may as well assume
that E = M × R is trivial. We consider two situations: First, if ω is a top degree
n
∗ k ∗ ∂f
= π∗ (π dϕ)f dt ∧ · · · ∧ dt + (−1) π ϕ ∧ µ dx ∧ dt ∧ · · · ∧ dt
1 n µ 1 n
∂x
= π∗ dω.
Now suppose ω does not have top degree in the fiber coordinate:
which is zero if dtµ ∧ dti1 ∧ · · · ∧ dtir 6= ±dt1 ∧ · · · ∧ dtn . Otherwise, we note that
Z
∂f µ
µ
dt ∧ dti1 ∧ · · · ∧ dtir = 0
Rn ∂t
−1
cv (M ), hence for each fixed x ∈ M , we have π (x) ∩ Supp ω is
because ω ∈ Ωk+n
compact. Hence ω has compact support on each fiber for a fixed x. Thus
Z +∞
∂f µ
µ
dt = f (..., +∞, ...) − f (..., −∞, ...) = 0.
−∞ ∂t
by π∗ [ω] = [π∗ ω], where on the right hand side the π∗ is the integration along fiber
cv (E) → Ω (M ).
map π∗ : Ωk+n k
π∗ (π ∗ ω ∧ τ ) = ω ∧ π∗ τ.
Proof. (a). Since two forms are the same iff they are the same at every point, we
may treat this problem locally and assume E = M × Rn . If
Now suppose
τ = (π ∗ ϕ)f (x, t)dt1 ∧ · · · ∧ dtn (6.2)
a top degree form in fiber coordinate with ϕ ∈ Ωk (M ). Then
Z
∗
π∗ (π ω ∧ τ ) = ω ∧ ϕ f (x, t)dt1 · · · dtn = ω ∧ π∗ τ.
Rn
and
Z XZ XZ XZ
∗
ω ∧ π∗ τ = ω ∧ π∗ ((π ρα )τ ) = ω ∧ ρα π ∗ τ = ω ∧ π∗ ((π ∗ ρα )τ )
M α M α M α Uα
by part (a). Hence again we reduced the problem to a local trivilization. That is, it
R that E = M × R again. With this setup, first suppose τ is given
n
suffices to assume
by (6.1). Then M ω ∧ π∗ τ = 0 because π∗ τ = 0. On the other hand,
Z Z
∗
π ω∧τ = π ∗ (ω ∧ ϕ) ∧ f (x, t)dti1 ∧ · · · ∧ dtir = 0
E E
because
because
P r < n by assumption. P Now assume τ is given by (6.2). Assume that
ω = |I|=m−k ωI dxI and ϕ = |J|=k ϕJ dxJ with ωI , ϕJ functions on M . Then we
have
Z Z Z
ω ∧ π∗ τ = (ωI dx ∧ ϕ) ·
I
f (x, t)dt1 · · · dtn
ZM R
Z
M n
= ωI ϕJ dx ∧ dx ·
I J
f (x, t)dt1 · · · dtn
ZM Rn
= π ∗ ω ∧ τ.
E
Proof. This is exactly like the proof of the Poincaré Lemma for compactly supported
cohomology (Theorem 6.51). We have already constructed R π∗ and showed that it
commutes with d. Now we define e ∈ Ωc (R ) with Rn e = 1, and define e∗ :
n n
∗
Ωk (M ) → Ωk+n
cv (M × R ) via e∗ : ϕ 7→ (π ϕ) ∧ e. Then we construct the chain
n
homotopy operator just like we did previously, and proof of Theorem 6.53 works
identically.
Now we are ready to prove the Thom isomorphism theorem in its most general
form.
Theorem 6.89 (General Thom isomorphism). Let E be an orientable vector bundle
of rank n over an orientable manifold of finite type. Then integration along fiber
gives an isomorphism
n+k
π∗ : Hcv (E) → H k (M ).
Proof. Note that E is an orientable manifold of finite type by Theorem 6.77 and
Corollary 6.80. The proof of Thom isomorphism is another example of the Mayer-
Vietoris argument. Let U, V be open cover of M . Using a partition of unity from
the base M , we obtain a partition of unity on E, subordinate to U 0 = π −1 (U ), V 0 =
π −1 (V ). We then see that the sequence
d∗
··· k+n
Hcv (E|U ′ ∪V ′ ) k+n
Hcv (E|U ′ ) ⊕ Hcv
k+n
(E|V ′ ) k+n
Hcv (E|U ′ ∩V ′ ) k+n+1
Hcv (E|U ′ ∪V ′ ) ···
π∗ π∗ π∗ π∗
The first two squares of this diagram are clearly commutative. We check the com-
mutativity for the third square. Recalling from Remark 6.36, we have that
π∗ d∗ ω = π∗ d((π ∗ ρU )ω) = π∗ ((π ∗ dρU ) ∧ ω) = dρU ∧ π∗ ω = d∗ π∗ ω
where we used the projection formula, i.e., part (a) of Theorem 6.87, and the fact
that both ω and dω are closed since they represent cohomology classes. So the
diagram is commutative.
Note that if U, V are good covers, then E|U ′ , E|V ′ , E|U ′ ∩V ′ are trivial and U 0 , V 0
are good covers on E, by Lemma 6.79 and Corollary 6.80. So in these cases, π∗ are
isomorphisms by Proposition 6.88. By the Five Lemma if the Thom isomorphism
holds for U 0 , V 0 , U 0 ∩ V 0 , then it holds for U 0 ∪ V 0 . The proof now proceeds by
induction on the cardinality of a good cover for the base (which is also the cardinality
of the corresponding good cover for E).
Definition 6.90 (Thom class). Let E be an orientable rank n vector bundle over an
n
orientable manifold M of finite type. The Thom isomorphism π∗ : Hcv (E) → H 0 (M )
determines a cohomology class
τ (E) = (π∗ )−1 (1) ∈ Hcv
n
(E)
called the Thom class of E. It is characterized by the property that π∗ τ (E) = 1.
That is, the Thom class τ (E), when integrated on each fiber (with co-orientation)
gives unit.
Proposition 6.91. The inverse Thom isomorphism (π∗ )−1 : H k (M ) → Hcv
n+k
(E) is
given by
(π∗ )−1 (α) = π ∗ α ∧ τ (E).
Proof. Since π∗ τ (E) = 1, by the projection formula (Theorem 6.87), one has
π∗ (π ∗ α ∧ τ (E)) = α ∧ π∗ τ (E) = α ∧ 1 = α.
This proves the claim.
0 → T S → T M |S → N S → 0
Np S S
Tp S
p
Figure 6.4: One should think of the normal bundle N S of S literally as being the
bundle whose fiber at each point p ∈ S is tangent to S, i.e., orthogonal to Tp S.
At every p ∈ S we have an exact sequence of vector spaces 0 → Tp S → Tp M →
Np S → 0, which means that Np S ∼ = Tp M/Tp S, i.e., Np S is isomorphic to the vector
space that is the complement of Tp S in Tp M , which is isomorphic to the orthogonal
complement of Tp S in Tp M , if we have a Riemannian metric on M .
By the tubular neighborhood theorem, which we will not prove, every submani-
fold S has a tubular neighbohood of T (S), and in fact T (S) is diffeomorphic to an
open neighborhood the normal bundle N S. The reader is encourage to refer to [1,
Theorem 6.24] and [7, p.67] for a proof, but we will not dive into this rabbit hole.
The orientation of T (S) and N S, when viewed as subsets of M , is given by the
induced orientation of M . However, we may orient the fiber of these bundles again
using co-orientation. For example, the normal bundle is oriented in such a way
that T S ⊕ N S has the orientation of T M |S .
Now, we apply the Thom isomorphism theorem to the normal bundle N S. Then
we get a sequence of maps
(π∗ )−1 j∗
H 0 (S) −−−−→ Hcv
m−k
(N S) −
→ H m−k (M )
because ∂(N S) = ∅. Note that the dimension in the equation above works out:
dim ∂(N S) = dim(N S)−1 = m−1, and deg(α∧τ (N S)) = (k−1)+(m−k) = m−1.
Hence (6.4) reads
Z Z Z Z
∗ ∗ ∗
ω ∧ τ (N S) = π i ω ∧ τ (N S) = i ω ∧ π∗ τ (N S) = i∗ ω (6.5)
NS NS S S
by part (b) of the projection formula (Theorem 6.87), and by the definition of Thom
class that π∗ τ (N S) = 1. Putting (6.3), (6.4), and (6.5) together, we see that
Z Z
ω ∧ j∗ τ (N S) = i∗ ω.
M S
This is exactly the characterizing property of the Poincaré dual. Hence we conclude
that ηS = j∗ τ (N S).
Corollary 6.95. Suppose π : E → M is an oriented vector bundle over an oriented
compact manifold. Then the Poincaré dual η of the zero section of E can be chosen
to be the Thom class τ (E) of E.
Proof. Note that M embeds diffeomorphically as the zero section of E, hence there
is an exact sequence of vector bundles
0 → T M → T E|M → E → 0.
In other words, the normal bundle of M in E is E itself. Then the claim follows
immediatly from Theorem 6.94.
115 CHAPTER 6. COHOMOLOGY
The goal of this chapter is to prove some of the most important results in differential
topology using de Rham cohomology theory. The main tools that we will introduce
are degree theory and intersection theory.
The readers may recognize that degree and intersection theory are also the main
tools used to prove these results in standard differential topology textbooks like [7]
and [8]. But we want to stress that these tools arises naturally under the framework
of de Rham cohomology, whereas in the texts mentioned above it is not obvious
how this is true. Moreover, topological invariants like Euler characteristics and Lef-
schetz number are defined using intersection numbers directly in these texts without
involving (co)homology groups, thus readers who choose to continue their study in
algebraic topology may find it difficult to see why these definitions are equivalent
when they read a textbook like [3]. Hence the goal of this writing is to bridge the
gap between existing literature and hopefully provide a smooth transition between
the frameworks of differential and algebraic topology.
Unfortunately, the prerequisites needed to understand the proofs are by no ways
trivial. It would be ideal if the reader knows proofs of results in differential topol-
ogy like regular value theorem, Thom’s transversality theorem, Sard’s theorem, and
tubular neighborhood theorem. This is well introduced in Chapter 6 of [1] or in
standard texts like [7] and [8]. However, it is sufficient to just know the results
without knowing the proof. Still, readers might find it difficult to understand this
chapter because we quickly introduce relevant concept and omit the proofs of many
important results. This is completely normal, since our only goal here is to provide a
rapid introduction to this vast and deep subject. My goal here is not trying to make
the readers understand everything in this chapter, but to show them many interest-
ing and beautiful results. If you are fascinated by some of these results despite not
understanding a considerable amount of material, then I have successfully fulfilled
my duty. In that case, grab a book on differential topology and start reading! And
don’t worry: it will certainly be more well-written than anything you read here!
the category of manifolds and smooth maps. However, if f is proper, then since
Supp f ∗ ω = f −1 (Supp ω), then we see that f ∗ ω ∈ Ωkc (M ). Moreover, one can show
that
Proposition 7.2. Ω∗c (and also Hc∗ ) is a contravariant functor in the category of
proper smooth manifolds (where the objects are smooth manifolds and the mor-
phisms are proper smooth maps) to the category of vector spaces.
Note that we know from Section 6.4.1 that Ω∗c is a covariant functor under
inclusions of open sets.
Now suppose that M m is a connected oriented manifold. It was shown in Corol-
lary 6.68 that Hcn (M ) ∼
= R.
Definition 7.3. If both M m , N m are oriented and connected manifolds of the same
degree, and f : M → N is a proper map, then f ∗ : Hcm (N ) → Hcm (M ) will be
a linear map between R, i.e., is multiplication by a number. This real number is
denoted by deg(f ), called the degree of f .
One can also remove the proper restriction from the definition above, but then
we must assume that both M and N are compact manifolds. But of course in this
case, Hck (M ) = H k (M ) since M is compact.
Remark 7.4. Note that the Poincaré duality theorem states that the isomorphism
Hcn (M ) → R is given by integration. Therefore, we have that for any ω ∈ Ωnc (M ),
Z Z
∗
f ω = (deg f ) · ω.
M N
Proof. Let q ∈
/ f (M ) in N . By the lemma above, f (M )c is open, so we choose an
Ropen neighborhood V of q such that f (M ) ∩ V = ∅. Choose ω ∈ Ωm c (V ) such that
∗
N
ω = 1. But f ω = 0. Hence deg(f ) = 0.
Definition 7.7. Let F : M → N be a smooth map between manifolds. A point
q ∈ N is called a regular value of f if for every p ∈ F −1 (q) ⊂ M , the differential
dFp : Tp M → Tq N is surjective. If q is not a regular value of F , it is called a
singular value of F .
It is a deep theorem of Sard, proven in Chapter 6 of [1], that the set of critical
values in N has measure zero. Hence, almost every point in N is a regular value. In
particular, it is always possible to find a regular value of any smooth map F : M →
N on N .
Now we show how the definition of degree (usually defined in algebraic topology)
relates the degree in differential topology:
Theorem 7.8. Assume f : M m → N m a proper map between orientable manifolds,
and q ∈ N is a regular value of F , then we have
X
deg(f ) = sign(dfx ).
x∈f −1 (q)
B1 (0) ψ
B1 (0)
F : S 2n × [0, 1] → S 2n
(p, t) 7→ p cos(πt) + Xp sin(πt).
X
k
deg(fj ) = 0.
j=1
Proof. The proof is completely analogous to the proof of the previous lemma. Let
iRj : ∂j M → M be the inclusion. Then fj = g ◦ ij . Let ω ∈ Ωm−1 (X) such that
X
ω = 1. Then dω = 0 since ω is already a top degree form on X. Then
X
k X
k Z k Z
X
deg(fj ) = deg(fj ) ω= fj∗ ω
j=1 j=1 X j=1 ∂j M
k Z
X Z Z
= i∗j g ∗ ω = ∗
d(g ω) = g ∗ (dω) = 0
j=1 ∂j M M M
Theorem 7.13 (Brouwer Fixed Point Theorem). Let Dm ⊂ Rm be the closed unit
disk. Then every continuous map f : Dm → Dm has a fixed point.
F1 (p)
F (p) = .
1+r
Then F maps Dm to Dm by triangle inequality, and F is a smooth map from Dm
to itself. Moreover,
Thus |F (p) − p| ≤ |F (p) − F0 (p)| + |F0 (p) − p| < 2|F0 (p) − p|. Hence F : Dm → Dm
is smooth and does not have a fixed point.
Now consider the smooth map G : Dm → S m−1 defined by
p − F (p)
G(p) = .
|p − F (p)|
121 CHAPTER 7. DIFFERENTIAL TOPOLOGY
Lemma 7.14. Suppose f : S m → S m is a smooth odd map, which means for any p
we have f (−p) = −f (p), then deg(f ) is an odd number.
this, note that since N is a regular value of f , so for any x ∈ f −1 (±N ) ∩ S∧1 , we
have dfx : Tx S 1 → Tf (x) S 1 is surjective. Then for any vector v ∈ Tx S 1 , there exists
a curve γ such that γ 0 (0) = v and γ(0) = x. If f sends adjacent arcs separated by x
to the same hemisphere, then3 dfx (v) = (f ◦ γ)0 (0) = 0, since f ◦ γ changes direction
at f (x), i.e., at t = 0. Thus dfx is not surjective.
Therefore, adjacent arcs goes to different hemisphere S>1 or S<1 . Also note that
f ((1, 0)) and f (0, 1) belong to S>1 and S<1 (or the other way around) separately.
So we must start with arc mapping into S>1 , and then arc mapped into S<1 , and
alternate... and eventually end with arc mapping into S<1 . So there must be an even
number of arcs, i.e., odd number of endpoints. So |f −1 (±N ) ∩ S∧1 | = |f −1 (N )| is
odd. This proves the base case.
Now we prove the inductive case. To this end, suppose f : S m → S m is an odd
map. WLOG N = (0, ..., 0, 1) is a regular value of f . Denote S m−1 the equator of S m
by setting xm+1 = 0. By the same argument, we have f −1 (±N ) ∩ S∧m = {p1 , ..., pk }
and
1
|f −1 (N )| = |f −1 (±N )| = |f −1 (±N ) ∩ S∧m |.
2
By the inverse function theorem and possibly shrinking neighborhoods, we can
choose neighborhoods U of N , disjoint Ui of pi for i = 1, ..., k, such that f sends
each Ui diffeomorphically to U or −U , and such that Ui ⊂ S∧m for all i. Now put
a
k
M= S∧m − Ui
i=1
and
g = r ◦ π ◦ f |M
where π(x1 , ..., xm+1 ) = (x1 , ..., xm ) is the projection and r(x) = x/|x| retract from
B1 (0) \ {0} to S m−1 . We set g0 = g|S m−1 . Then by straightforward calculation, g0 is
an odd map. Now consider gi = g|∂Ui . Since we have a sequence of diffeomorphisms
f π r
∂Ui −
→ ±∂U −
→ π(∂U ) −
→ S m−1
Proof. For m = 1, define g(p) = f (−p) − f (p). We want to show g has a zero
on S 1 . Note g(1, 0) = f (−1, 0) − f (1, 0) and g(−1, 0) = f (1, 0) − f (−1, 0). Hence
g(1, 0) = −g(−1, 0). If g(1, 0) = 0 we are done, otherwise g(S 1 ) is connected and
contains both positive and negative numbers on the real line, so 0 ∈ g(S 1 ). This
proves the base case.
Now for m 6= 1, suppose f : S m → Rm is continuous, and suppose f (−p) 6= f (p)
for all p ∈ S m . Let ε = inf p∈S m |f (−p) − f (p)| > 0. Then take an ε/4-smooth
approximation f 0 of f by Whitney approximation theorem (Theorem 1.30). If there
exists a p such that f 0 (−p) − f 0 (p) = 0, then
ε ε
|f (−p) − f (p)| ≤ |f (p) − f 0 (p)| + |f 0 (p) − f 0 (−p)| + |f 0 (−p) − f (−p)| ≤ + <ε
4 4
which contradicts the choice of ε. Therefore we have found a smooth map f 0 : S m →
Rm , such that f 0 (−p) 6= f (p) for all p ∈ S m . Now let g : S m → S m−1 be defined by
f 0 (p) − f 0 (−p)
g(p) = .
|f 0 (p) − f 0 (−p)|
By definition, g̃ = g|S m−1 is a map from S m−1 to itself, and is an odd map by
construction. Therefore by the lemma above, deg(g̃) is odd. Hence in particular
deg(g̃) 6= 0. Now
g|S∧m : S∧m → S m−1
is clearly an extension of g̃, as ∂S∧m = S m−1 . Thus by Lemma 7.11, we have deg(g̃) =
0, which is the desired contradiction.
L
L
Tp L
p S p S
Tp S Tp L Tp S
Note that ϵ(p, S, L) = (−1)sℓ ϵ(p, L, S), and similarly I(S, L) = (−1)sℓ I(L, S).
Now we present the key result of this section:
Theorem 7.18. Suppose that S s , Lℓ are closed, oriented submanifolds of an oriented
manifold N n of finite type (i.e., has a finite good cover), such that
• S is compact;
• s + ℓ = n;
• L ⋔ S.
Now let ηS be the compactly supported closed form representing the Poincaré dual of
S (by Proposition 6.73) in N , and let ηL be a closed form representing the Poincare
dual of L in N . Then we have
Z
I(S, L) = ηS ∧ ηL .
M
Proof. For simplicity, we write ϵ(p) := ϵ(p, S, L). Since L ⋔ S, for all p ∈ S ∩ L,
we can find4 a local coordinate (x1 , ..., xn ) defined on an open neighborhood Up of p
such that
Up ∩ S = {xs+1 = · · · = xs+ℓ = 0}
4
Short proof: Since Tp L ⊕ Tp S = Tp N , we can choose v1 , ..., vℓ ∈ Tp L and vℓ+1 , ..., vℓ+s ∈ Tp S.
Then give N a Riemannian metric, and let the curve xi (λ) = expp (λvi ). Then the coordinates
x1 , ..., xs+ℓ will do the job. For a more detailed and rigorous proof, see Lemma 18 on p.119 of [12].
125 CHAPTER 7. DIFFERENTIAL TOPOLOGY
and
Up ∩ L = {x1 = · · · = xs = 0}.
The orientation from on S ∩ Up is given by dx1 ∧ · · · dxs , and the orientation form on
L ∩ Up is given by dxs+1 ∧ · · · ∧ dxn . Then the wedge dx1 ∧ · · · ∧ dxn will correspond
to direct sum orientation. Hence the orientation form on Up is given by
Np = (L ∩ Up ) ∩ N =: Lp
N S
Up
N
L
Lp
Tp N = ϵ(p)Tp S ⊕ Tp L
we have that
Lp = ϵ(p)Np .
7.4. INTERSECTION THEORY 126
Then we have, by the definition of Poincaré dual and discussion above, that
Z Z X Z X Z X
ηS ∧ ηL = ηS = ηS = ϵ(p) ηS = ϵ(p)
M L p∈S∩L Lp p∈S∩L Np p∈S∩L
ϕ
T2
φ
S
L
θ
2
[0, 2π)
Figure 7.3: Let S, L be two submanifolds of T 2 with orientations indicated as in
the figure, and let φ : [0, 2π)2 → T 2 be the diffeomorphism parameterizing the torus
as shown in the figure. It is easy to see that φ is orientation preserving, where we
define the orientation of T 2 as follows: View T 2 as a submanifold of R3 , and let
v, w ∈ Tp T 2 . We declear the ordered basis v, w ∈ Tp T 2 to be positive iff v, w, n is a
positive basis for R3 , where n is the normal to T 2 at p.
We see that I(S, L) = ϵ(p, S, L) = +1 from Figure 7.3 where x = (a, b). We
know that ηS satisfies Z
ηS = 1,
Np S
Again this is consistent with the result we get from counting I(S, L).
One can check that our choice of ηS , ηL are indeed the Poincaré dual of S and L.
To see this, let α ∈ Ω1 (T 2 ) be any closed 1-form. Write φ∗ α = F (θ, ϕ)dθ+G(θ, ϕ)dϕ.
Then we see that φ∗ (α ∧ ηS ) = F (θ, ϕ)ρ(ϕ)dθ ∧ dϕ. Since Supp ηS can be chosen to
be sufficiently small as long as it is contained in a thin tubular neighborhood of S,
and ρ has mass 1, we see that we can make ηS more and more concentrated on S:
think of it as a dirac delta function in the limit. If we denote ϕ0 the ϕ coordinate of
φ−1 (S) in [0, 2π)2 , we can make ρ(ϕ) → δ(ϕ0 ), then it is clear that
Z Z 2π Z 2π Z 2π Z
∗ 0
φ (α ∧ ηS ) = F (θ, ϕ)ρ(ϕ)dθdϕ = F (θ, ϕ )dθ = φ∗ α.
[0,2π)2 0 0 0 φ−1 (S)
Under the diffeomorphism φ, we see that on T 2 the integral equality above becomes
Z Z
α ∧ ηS = α,
T2 S
Then X
∗ j ∗ m−j
ηΓ f = j (π1 ωi ∧ π2 θk
(−1)j Aik ). (7.2)
i,j,k
Proof. Since all the cohomology groups are finite dimensional vector spaces, under
Poincaré duality and dual vector space isomorphism, we have
= Hcm−j (M )∗ ∼
H j (M ) ∼ = H m−j (M )∗ ∼
= H m−j (M ).
is a basis of H m (M × M ).
Now we begin our calculation:
Z
(π1∗ ωij ∧ π2∗ θkm−j ) ∧ (π1∗ θsm−u ∧ π2∗ ωtu )
M ×M
Z
= (−1) m(m−j)
π1∗ (ωij ∧ θsm−u ) ∧ π2∗ (ωtu ∧ θkm−j )
ZM ×M Z
= (−1) m(m−j)
ωi ∧ θs
j m−u
ωtu ∧ θkm−j
MZ MZ
= (−1) m(m−j)
δju ωi ∧ θs
j m−j
ωtj ∧ θkm−j
M M
m(m−j)
= (−1) δju δis δtk .
In the calculation above, the first equality can be seen by swapping the 2nd term
in the wedge to the last; the 3rd equality can be seen by considering integrating on
local coordinate of product M × M ; the 4th equality can be seen by the following
argument: Since all forms
R are compactly supported, if one coordinate or more is
missing, say x , then φ(U ) (· · · )dxn = 0 where φ is some local chart and U ⊂ M is
n
some open set containing support of the form in question (if it is not contain in one
chart, use a partition of unity argument).
Now we write ηΓf as a linear combination of basis elements:
X
∗ j ∗ m−j
ηΓ f = j (π1 ωi ∧ π2 θk
cik )
i,j,k
129 CHAPTER 7. DIFFERENTIAL TOPOLOGY
and denote the expression on the right hand side by ω. For fixed u, s, t, by the
definition of Poincaré dual we have
Z Z
∗ m−u ∗ u
(π1 θs ∧ π2 ωt ) ∧ ω = i∗ (π1∗ θsm−u ∧ π2∗ ωtu ) (7.3)
M ×M Γf
where i : Γf → M × M denote the inclusion map. We now calculate the left hand
side and the right hand side of (7.3) separately. The left hand side is
X Z
(−1) m2
cik
j (π1∗ ωij ∧ π2∗ θkm−j ) ∧ (π1∗ θsm−u ∧ π2∗ ωtu )
i,j,k M ×M
2
X
= (−1)m cik
j δju δis δtk (−1)
m(m−j)
i,j,k
m2 +m(m−u)
= (−1) cst
u
= (−1)−mu cst
= (−1)mu cst
u.
Equating the left hand side and right hand side of (7.3) gives
(−1)mu cst
u = (−1)
u(m−u) st
Au .
Hence
−u 2
cst
u = (−1) Ast u st
u = (−1) Au .
I(Γf , ∆) = L(f ).
Proof. By the definition of Poincaré dual, and the previous lemma, we have
Z Z
η Γ f ∧ η∆ = i∗∆ ηΓf = L(f ).
M ×M ∆
The second part of the theorem follows directly from Theorem 7.18. Finally, if f
has no fixed point, then Γf ∩ ∆ = ∅. Hence by the definition of intersection number,
I(Γf , ∆) = 0. Thus L(f ) = 0.
Remark 7.26. Therefore, we see an alternative definition of L(f ) for a smooth map
f : M → M between a compact oriented manifold, which is L(f ) := I(Γf , ∆). Note
that some books like [7] defined L(f ) = I(∆, Γf ). But we insist on defining L(f )
to be I(Γf , ∆) because it agrees with the cohomological definition, i.e., that L(f ) is
the alternating trace of f ∗ on cohomology groups, where the other convention is off
by a sign when dimension is odd.
We conclude this section with a final result, which relates Lefschetz number L(f )
with its local Lefschetz number Lx (f ) at every fixed point x.
Hence it suffices to prove that ϵ(p, Γf , ∆) = Lp (f ) for every fixed point p. Let
v1 , ..., vm be a positively oriented basis of Tp M . Then v1 ⊕ v1 , ..., vm ⊕ vm is a
positively oriented basis of T(p,p) ∆, and v1 ⊕ dfp (v1 ), ..., vm ⊕ dfp (vm ) a positively
oriented basis of T(p,p) Γf . Then a positively oriented basis of T(p,p) Γf ⊕ T(p,p) ∆ is
By the way, this section provides a full solution to Exercise 11.26 in [2], in case
you haven’t noticed (although the original problem has a typo, as the next remark
suggests).
Remark 7.30. One would note that the local Lefschetz degree in our definition
Lx = sign det(I − dfx ) is off by a factor (−1)m compared to literature like [7],
where they define Lx = sign det(dfx − I). P Thisj is because we
want to stick to
∗
the definition where L(f ) is definedPto be j (−1) Tr f |H j (M ) , and we would also
like to have the relation L(f ) = x∈Fix(f ) Lx (f ). These two requirements forces
133 CHAPTER 7. DIFFERENTIAL TOPOLOGY
Lx (f ) = sign det(I − dfx ), which disagrees with the convention like [7] when the di-
mension is odd. These different sign conventions can be quite annoying, and I think
Bott and Tu missed this in their Exercise 11.26 on p.129 of [2]. In fact, in [12], it is
calculated on p.123 that the local Lefschetz degree must be Lx (f ) = sign det(I − dfx )
for Theorem 7.29 to hold. See also the discussion in this MSE post.
Remark 7.32. Since the wedge product is alternating, one has dxi ∧ dxi = 0 for
all i. Therefore one may wonder why don’t we immediately get ηS ∧ ηS = 0. This
is not necessarily true. The reason is that ηS may not be a pure wedge, it can also
be linear combination of them. For example, take ηS = dx1 ∧ dx2 + dx3 ∧ dx4 , then
ηS ∧ ηS = 2dx1 ∧ dx2 ∧ dx3 ∧ dx4 .
• Ft ⋔ S, i.e., Ft (S) ⋔ S.
Thus we have Z Z
i∗S ω = i∗L ω
S L
7.6. EULER CHARACTERISTICS AND POINCARÉ-HOPF THEOREM 134
for every t ∈ [0, 1]. Hence our most important tool in intersection theory, Theorem
7.18, can be generalized to self-intersection number again.
Definition 7.34. Let M be a smooth manifold of finite type. Then the Euler
characteristic χ(M ) of M is defined to be the number
X X
χ(M ) = (−1)j rank H j (M ) = (−1)j βj .
j j
Proof. Take f : M → M to be idM in the Lefschetz fixed point theorem, and note
that
Z
I(∆, ∆) := η∆ ∧ η∆
M ×M
= L(idM )
X
= (−1)j Tr idM : H j (M ) → H j (M )
j
X
= (−1)j rank H j (M ).
j
where the 2nd equality is by the first part of the Lefschetz fixed point theorem. This
concludes the proof.
We immediately get
Proof. Just note that I(∆, ∆) = (−1)(2n+1) I(∆, ∆) = −I(∆, ∆) if we swap the two
2
Remark 7.38. Textbooks like [7] uses I(∆, ∆) as the definition of the Euler char-
acteristic χ(M ). The discussion above gives the reason why this definition is com-
patible with the definition using cohomology, which is more common in algebraic
topology literature.
We conclude this section by giving a proof of the Poincaré-Hopf theorem, which
requires almost no additional work other than what we have already developed.
Let M be a compact manifold, X a smooth vector field on M . For sufficiently
small t, let φt : M → M be the flow generated by X. Then it is not hard to see that
x ∈ M is a zero of the vector field X iff φt fixes x. Then from dφtx : Tx M → Tx M
we get a linear map
d
Ax = (dφtx ) .
dt t=0
Now if m is even, then this is exactly the Poincaré-Hopf theorem. If m is odd, then
we already know χ(M ) = 0 by Corollary 7.36. Hence multiplying both sides of the
equation above by (−1)m we again get Poincaré-Hopf theorem as desired.
Note that one can use Poincaré-Hopf theorem to give an alternative proof of the
hairy ball theorem:
Proof of Hairy Ball Theorem. Since χ(S 2n ) = 2, we conclude that smooth vector
fields on S 2n must have a zero.
7.7. EULER CLASS 136
H : M × [0, 1] → E
(p, t) 7→ (p, ts(p)).
M1 ψ0
M2
Then we have Z Z
e(E1 ) = e(E2 ).
M1 M2
Proof. We have already shown that the Euler class is independent of the choice of
sections. The commutative diagram implies s∗1 ψ ∗ = ψ0∗ s∗2 . Hence we have
But we claim that ψ ∗ τ (E2 ) = τ (E1 ). This is because the Thom class of E1 is
characterized by the property that fiber integration map applied to it gives unit:
π∗ τ (E1 ) = 1 by definition. But since ψ is an orientation preserving bundle isomor-
phism, it preserves fiber integration. Or in other words
Z Z Z
∗
ψ τ (E2 ) = τ (E2 ) = τ (E2 ) = 1
(E1 )p ψ((E1 )p ) (E2 )ψ0 (p)
137 CHAPTER 7. DIFFERENTIAL TOPOLOGY
This implies
Z Z Z Z
e(E1 ) = ψ0∗ e(E2 ) = e(E2 ) = e(E2 ).
M1 M1 ψ0 (M1 ) M2
ψ1 ψ2 ψ3
TM T∆ N∆ N (∆)
s1 s2 s3 s4
M ∆ id∆
∆ id∆
∆
where ψ1 (x, v) = (x, x, v, v), ψ2 (x, x, v, v) = (x, x, −v, v), and ψ3 : N (∆) → N (∆)
is an orientation preserving bundle isomorphism, whose existence is guaranteed by
the tubular neighborhood theorem. It is easy to verify that each of the blocks in the
diagram above satisfies the conditions in the previous lemma. Hence by repeatedly
applying the lemma above, we get
Z Z Z Z
e(T M ) = e(T ∆) = e(N ∆) = e(N (∆)).
M ∆ ∆ ∆
Now, s4 : ∆ → N (∆) is homotopic to the zero section (inclusion), and τ (N (∆)) may
also be taken as the Poincaré dual η∆ of ∆ in N (∆) by Corollary 6.95. Therefore,
by the definition of the Poincaré dual, we get that
Z Z Z Z
∗ ∗
e(N (∆)) = i∆ τ (N (∆)) = i∆ η∆ = η∆ ∧ η∆ = χ(N (∆)) = χ(M )
∆ ∆ ∆ N (∆)
where the last equality is because N (∆) deformation retracts onto ∆, which is
diffeomorphic to M , and the Euler characteristic is a homotopy invariant.
This concludes the proof.
7.7. EULER CLASS 138
We conclude this writing with a story of René Thom (1923-2002), who won the
Fields medal in 1958 for his contribution to topology. We hope that this serves as a
belated celebration of his 100th birth anniversary:
Once Thom was joined by two anthropologists in an interview. One of the ques-
tions they discussed was why human ancestors wanted to preserve tinders (Small dry
sticks or finely-divided fibrous matter used to help light a fire). One anthropologist
said that because fire can help ancient human to stay warm, another said because
fire can help cook delicious food. Thom, instead, said that when the night comes,
the waving firelight is breathtakingly beautiful.
Beauty is the eternal pursuit of mathematicians.
covariant, 69 form, 49
preserving, 57
good cover, 95 reversing, 57
gradient, 35 oriented
hairy ball theorem, 119, 135 manifold, 49
Hausdorff, 13 positively, 55
homeomorphism, 8 trivilization, 104
homotopic, 73 outward-pointing, 60
path, 73
partition of unity, 19
homotopy, 73
path-connected, 11
relative to A, 73
Poincaré
equivalence, 73
dual, 103
equivalent, 73
duality, 97
rel A, 73
lemma, 88
index lemma for compactly supported co-
of a vector field, 135 homology, 90
integral, 55 Poincaré-Hopf theorem, 135
of top degree form, 57 point
integration boundary, 59
along fiber, 90, 106 interior, 59
intersection number, 124 pointed
local, 123 manifolds, 70
inward-pointing, 60 maps, 70
preimage, 8
Lefschetz proper map, 116
number, 130 pullback, 50
fixed point theorem, 130 pushforward, 30
map, 130
local Lefschetz degree, 131 refinement, 95
regular value, 118
manifold
retraction, 73
smooth, 15
topological, 13 Sard’s theorem, 118
with boundary, 59 second-countable, 13
maximal atlas, 15 section, 37
Mayer-Vietoris self-intersection number, 133
sequence for compact suppports, 82 short exact sequence, 79
argument, 95 simply connected, 75
sequence, 80, 81 singular value, 118
nondegenerate, 96 smooth
on a subset, 24
open, 9 maps, 16
orientable, 49 structure, 15
orientation, 49 snake lemma, 79
atlas, 49 Stokes’ theorem, 62
boundary, 61 subspace topology, 9
141 INDEX
support
of a function, 19
compact, 72
of a form, 72
symmetrization, 43
tangent space, 26
tensor, 33
alternating, 43
components, 34
symmetric, 43
Thom
isomorphism, 111
transversality theorem, 133
class, 112, 114
René, 138
topological space, 8
topology, 8
transversal intersection, 123
transversality, 124
trivialization, 37
trivilization
equivalent, 104
oriented, 104
tubular neighborhood theorem, 113
upper-half space, 59
velocity, 26
wedge product, 45
Whitney approximation theorem, 24
winding number, 75
form, 75
Bibliography
[2] Raoul Bott, Loring W. Tu, Differential Forms in Algebraic Topology, Springer
(1982).
[8] John Milnor. Topology from the Differentiable Viewpoint. Princeton University
Press, 1997.