The de Rham Cohomology and Its Applications
The de Rham Cohomology and Its Applications
The de Rham Cohomology and Its Applications
SHIBO DAI
Abstract. This paper is devoted to several commonly used homology and cohomology
theories, as well as an important result which links them together – the de Rham theorem.
We introduce singular homology, singular cohomology as well as de Rham cohomology
in the first few sections. Then we state and prove the de Rham theorem. Finally, an
application of the de Rham theorem will be provided.
1. Introduction
The primary idea behind (co)homology is trying to identify “n-dimensional holes” in a
topological space. Common examples are the hole in the center of a torus and the inside of
a hollow sphere. Holes are interesting to measure because the mechanisms used to measure
them, their homology and cohomology groups, are algebraic topological invariants.
There are two different ways to do this: singular homology and de Rham cohomology.
Singular homology groups are the quotients of the groups of “closed singular chains” and
“boundary singular chains”. Closed chains are chains sent to zero by the boundary operator.
Boundary chains are singular chains which are obtained by applying the boundary operator
to other chains, and are a subgroup of closed chains. De Rham cohomology groups are the
quotient of the groups of “closed forms” and “exact forms”. Closed forms are differential
forms whose exterior derivative is zero. Exact forms are differential forms which are obtained
by applying the exterior derivative to other forms and are a subgroup of closed forms.
While the methods are different, they have a similar logic behind them. The de Rham
Theorem proves an isomorphism between singular cohomology groups, the dual of the sin-
gular homology groups, and the de Rham cohomology groups.
Finally, in section 6 an application of the de Rham theorem is given.
2. Singular homology
This first section will be concerned with developing the theory of singular homology. To
begin, we define the p-simplex.
Definition 2.1. A (geometric) p-simplex is a subset of Rn of the form
∑p ∑
p
{ ti vi : 0 ≤ ti ≤ 1, ti = 1}.
i=0 i=0
Definition 2.4. The free abelian group generated by all singular p-simplices in X, denoted
as Cp (X), is called the singular chain group of X in degree ∑
p. An element in Cp (X) is called
a singular p-chain. It is a finite formal linear combination i ai σi , where ai ∈ Z.
We now turn our attention to the boundary operator, which decomposes p-simplices into
a sum of its (p − 1)-faces.
∑
p
∂p σ = (−1)i σ|[v0 ,··· ,vi−1 ,vi+1 ,··· ,vp ] .
i=0
Where σ|[v0 ,··· ,vi−1 ,vi+1 ,··· ,vp ] is the restriction of σ to [v0 , · · · , vi−1 , vi+1 , · · · , vp ].
Proposition 2.6. For any p-chain, the boundary of its boundary is trivial, or ∂(∂σ) = 0.
Zp (M )
Hp (M ) = .
Bp (M )
is a complex, called the singular chain complex, and Hp (M ) is the pth homology group of
this complex. The equivalence class in Hp (M ) of a singular p-cycle c is called its homology
class, and is denoted by [c]. We say that two p-cycles are homologous if they differ by a
boundary.
A continuous map F : M → N induces a homomorphism F# : Cp (M ) → Cp (N ) on each
singular chain group, defined by F# (σ) = F ◦ σ for any singular simplex σ and extended
linearly to chains. An easy computation shows that F# ◦ ∂ = ∂ ◦ F# , so F# is a chain
map, and therefore induces a homomorphism on the singular homology groups, denoted by
F∗ : Hp (M ) → Hp (N ). It is immediate that (G ◦ F )∗ = G∗ ◦ F∗ and (IdM )∗ = IdHp (M ) , so
pth singular homology defines a covariant functor from the category of topological spaces
and continuous maps to the category of abelian groups and homomorphisms. In particular,
homeomorphic spaces have isomorphic singular homology groups.
We will now give some important properties of the singular homology group.
In addition to the properties above, singular homology satisfies the following version of
the Mayer-Vietoris theorem. Suppose M is a topological space and U, V ⊆ M are open
THE DE RHAM COHOMOLOGY AND ITS APPLICATIONS 3
Hp (U ∩ V ) Hp (M )
j∗ l∗
Hp (V )
Theorem 2.9. (Mayer-Vietoris for Singular Homology). Let M be a topological space
and let U, V be open subsets of M whose union is M . For each p there is a connecting
homomorphism ∂∗ Hp (M ) → Hp−1 (U ∩ V ) such that the following sequence is exact:
∂∗ α β
··· Hp (U ∩ V ) Hp (U ) ⊕ Hp (V ) Hp (M )
∂∗ α
Hp−1 (U ∩ V ) ···
where
α[c] = (i∗ [c], −j∗ [c]), β([c], [c′ ]) = k∗ [c] + l∗ [c′ ],
and ∂∗ [e] = [c], provided there exist f ∈ Cp (U ) and f ′ ∈ Cp (V ) such that k# f + l# f ′ is
homologous to e and (i# c, −j# c) = (∂f, ∂f ′ ).
∂∗ k∗ ⊕l∗
H p+1 (M ; R) ···
where the maps k ∗ ⊕ l∗ and i∗ − j ∗ are defined by
(k ∗ ⊕ l∗ )ω = (k ∗ ω, l∗ ω),
4 SHIBO DAI
(i∗ − j ∗ )(ω, η) = i∗ ω − j ∗ η,
and ∂ ∗ is defined by ∂ ∗ (γ) = γ ◦ ∂∗ with ∂∗ as in Theorem 2.7.
We will now begin to define smooth singular homology, which will correspond closely with
our original definitions of singular homology.
Definition 3.4. For a smooth manifold M , a smooth p-simplex in M is defined to be a
smooth map σ : ∆p → M .
Definition 3.5. The smooth chain group in degree p is denoted as Cp∞ (M ) ⊆ Cp (M ).
Elements in the group, which are called smooth chains, are finite formal linear combinations
of smooth p-simplices.
Note that the boundary of a smooth simplex is a smooth chain, we now define
Definition 3.6. The pth smooth singular homology group of M is defined as
Ker(∂p : Cp∞ (M ) → Cp−1
∞
(M ))
Hp∞ (M ) = ∞ .
Im(∂p+1 : Cp+1 (M ) → Cp∞ (M ))
Where the boundary operator is the same as defined in Definition 2.5 as Cp∞ (M ) ⊆ Cp (M ).
Theorem 3.7. For a smooth manifold M , the map ℓ∗ : Hp∞ (M ) → Hp (M ), which is induced
by inclusion, is an isomorphism.
Theorem 4.8. Let M be a smooth manifold, with or without boundary. Let Ωk (M ) denote
the vector space of smooth k-forms for M . There exists a R-linear map d : Ωk (M ) →
Ωk+1 (M ) such that the following properties hold:
(1) d(ω ∧ η) = dω ∧ η + (−1)k ω ∧ dη.
(2) d2 = 0.
This is the exterior derivative on smooth manifolds.
One can also prove that the exterior derivative on manifolds is unique.
Definition 4.9. Let M be a smooth manifold with or without boundary, and let k ∈ Z≥0 .
The exterior derivative d : Ωk (M ) → Ωk+1 (M ) is R-linear, so its subspaces are also R-linear.
Thus, define
Z k (M ) = Ker(d : Ωk (M ) → Ωk+1 (M )),
B k (M ) = Im(d : Ωk−1 (M ) → Ωk (M )).
If k > dim(M ) or k < 0, then Ωk is the zero vector space.
We now define the de Rham cohmology groups.
Definition 4.10. The kth de Rham cohomology group of M is defined to be
k
HdR (M ) = Z k (M )/B k (M ).
We now prove an important result about de Rham cohomology groups.
Proposition 4.11.
⨿ For a countable collection of smooth k-manifolds
∏ {Mj } with or without
k k
boundary, M = j Mj . There exists an isomorphism from j HdR (Mj ) to HdR (M ), which
is induced by the inclusion maps ℓj : Mj → M .
Proof. The pullback of the inclusion maps ℓ∗j : Ωk (M ) → Ωk (Mj ) induce an isomorphism
∏
from Ωk (M ) to J Ωk (Mj ), sending ω to (ℓ∗1 ω, ℓ∗2 ω, · · · ) = (ω|M1 , ωM2 , · · · ). If a smooth k-
form is zero on each Mj , then it is zero on M , so the map is injective. The map is surjective
because defining a k-form on each Mj defines a k-form on M . By passing to quotients we
obtain an isomorphism on de Rham cohomology groups. This must be countable because
smooth manifolds are second-countable. Thus, in order for M to be smooth, it must be
second countable, which is only accomplished when Mj is countable. □
Definition 5.3. For a smooth manifold M , a smooth differential k-form ω, and a smooth
k-simplex σ, the integral of ω over σ is defined as
∫ ∫
ω= σ ∗ ω.
σ ∆k
6 SHIBO DAI
∑
Additionally, for a smooth k-chain c = i c i σi ,the integral of ω over c is defined to be
∫ ∑ ∫
ω= ci ω.
c i σi
Theorem 5.4. (Stokes’ Theorem for Chains). For a smooth k-chain c in a smooth manifold
M and a smooth (k − 1)-form ω on M ,
∫ ∫
dω = ω.
c ∂c
Finally, we will develop the core construction in Bredon’s proof of the de Rham theorem,
which bridges the de Rham cohomology and singular homology explicitly.
Definition 5.5. A linear map ℓ : HdR k
(M ) → H k (M ; R) is defined such that for any
[ω] ∈ HdR (M ) and [c] ∈ Hk (M ) ∼
k ∞
= Hk (M ),
∫
ℓ[ω][c] = ω.
c̃
Where c̃ is a smooth k-cycle that represents the homology class [c]. This is called the de
Rham homomorphism.
Proposition 5.6. The de Rham homomorphism is well-defined.
Proposition 5.7. (Naturality of the de Rham homomorphism). Let M, N be smooth man-
ifolds and k be a non-negative integer.
(1) For a smooth map F : M → N , the diagram commutes:
p F∗ p
HdR (N ) HdR (M )
ℓ ℓ
F∗
H (N ; R)
p
H (M ; R)
p
η ∈ Ωp−1 (U ) and η ′ ∈ Ωp−1 (V ) such that ω = η|U ∩V − η ′ |U ∩V , and then let σ be the
p-form that is equal to dη on U and to dη ′ on V .Then, because ∂f + ∂f ′ = ∂e = 0
and dη|U ∩V − dη ′ |U ∩V = dω = 0, we have
∫ ∫ ∫ ∫ ∫ ∫
′
ω= ω= η− η = η+ η′
c ∂f ∂f ∂f ∂f ∂f ′
∫ ∫ ∫ ∫ ∫
′
= dη + dη = σ+ σ = σ.
f f′ f f′ e
Theorem 5.8. (The de Rham Theorem). For any smooth manifold M and nonnegative
k
integer k, the de Rham homomorphism ℓ : HdR (M ) → H k (M ; R) is an isomorphism.
Proof. Let us say that a smooth manifold M is a de Rham manifold if the homomorphism
ℓ : HdRk
(M ) → H k (M ; R) is an isomorphism for each p. Since ℓ commutes with the co-
homology maps induced by smooth maps by the proposition above, any manifold that is
diffeomorphic to a de Rham manifold is also de Rham. The theorem will be proved once we
show that every smooth manifold is de Rham.
If M is any smooth manifold, let us call an open cover {Ui } of M a de Rham cover if each
subset Ui is a de Rham manifold, and every finite intersection Ui1 ∩ · · · Uik is de Rham. A
de Rham cover that is also a basis for the topology of M is called a de Rham ⨿ basis for M .
Step 1. If {Mj } is a countable collection of de Rham manifolds, then j Mj is de Rham.
By Propositions
⨿ 3.11 and 4.2, for both de Rham and singular cohomology, the inclusions
ℓj : Mj → j Mj induce isomorphisms between the cohomology groups of the disjoint union
and the direct product of the cohomology groups of the manifolds Mj . As a⨿result of Prop
5.7, these isomorphisms commute with the de Rham homomorphism for j Mj . So the
latter is itself an isomorphism.
Step 2. Every convex open subset of Rn is de Rham. By the Poincaré Lemma of differential
forms, if k ̸= 0, then HdR k
(U ) = 0. By Proposition 4.2, the singular cohomology groups of U
are also trivial when k ̸= 0 because of the homotopy equivalence of U and a one-point space.
0
If k = 0, then HdR (U ) is the space consisting of constant functions and is one-dimensional.
Because H0 (U ) is generated by a singular 0-simplex, H 0 (U ; R) = Hom(H0 (U ); R) is one
dimensional. Let f be the constant function 1 and σ : ∆0 → M be a singular, smooth
0-simplex. Then,
∫
ℓ[f ][σ] = σ ∗ f = (f ◦ σ)(0) = 1.
∆0
Since the de Rham homomorphism is not the zero map for k = 0, it is an isomorphism.
Step 3. If M has a finite de Rham cover, then M is de Rham. Let M = ∪ki=1 Ui where
{Ui } is a finite de Rham cover. We prove the result by induction on k. For k = 1, the
result is obvious. Suppose next that M has a de Rham cover consisting of two sets {U, V }.
Putting together the Mayer-Vietoris sequences for de Rham and singular cohomology, we
obtain the following commutative diagram, in which the horizontal rows are exact and the
vertical maps are all de Rham homomorphisms:
p−1 p−1 p−1 p
HdR (U ) ⊕ HdR (V ) HdR (U ∩ V ) HdR (M )
a b c
H p−1
(U ; R) ⊕ H p−1
(V ; R) H p−1
(U ∩ V ; R) H (M ; R)
p
p p p
HdR (U ) ⊕ HdR (V ) HdR (U ∩ V )
d e
H (U ; R) ⊕ H (V ; R)
p p
H (U ∩ V ; R)
p
8 SHIBO DAI
then U and V are disjoint unions of de Rham manifolds, and so they are both de Rham by
Step 1. Finally, U ∩ V is de Rham because it is the disjoint union of the sets Bm ∩ Bm+1
for m ∈ Z, each of which has a finite de Rham cover consisting of sets of the form Uα ∩ Uβ ,
where Uα and Uβ are basis sets used to define Bm and Bm+1 , respectively. Thus M = U ∪ V
is de Rham by Step 3.
Step 5. Every open subset of Rn is de Rham. If U ⊂ Rn is such a subset, then U has a
basis consisting of Euclidean balls. Because each ball is convex, it is de Rham, and because
any finite intersection of balls is again convex, finite intersections are also de Rham. Thus,
U has a de Rham basis, so it is de Rham by Step 4.
Step 6. Every smooth manifold is de Rham. Any smooth manifold has a basis of smooth
coordinate domains. Since every smooth coordinate domain is diffeomorphic to an open
subset of Rn , as are their finite intersections, this is a de Rham basis. The claim therefore
follows from Step 4. □
References
1. John M. Lee Introduction to smooth manifolds, Springer Graguate Texts in Mathematics 218.