Alg Top
Alg Top
Pingbang Hu
This is a graduate-level course on algebraic topology taught by Jennifer Wilson at University of Michigan.
It is self-contained enough that only requires a background in abstract algebra and some point set
topology. We will use Hatcher [HPM02] as the main text, but the order may differ here and there. Enjoy
this fun course! In particular, I added some extra content that is not covered in lectures, things like
groupoid, fibered coproduct, feel free to skip these contents.
This course is taken in Winter 2022, and the date on the cover page is the last updated time.
Contents
2 Category Theory 16
2.1 Category Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 Free Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4 Covering Spaces 55
4.1 Lifting Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.2 Deck Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5 Homology 74
5.1 Motivation for Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.2 Simplicial Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.3 Singular Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.4 Functoriality and Homotopy Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.5 Relative Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.6 Degree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.7 Cellular Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.8 Eilenberg-Steenrod Axioms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7 Epilogue 117
7.1 Proof of Seifert-Van-Kampen Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.2 Review Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
1
B Algebra 129
B.1 Abelian Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
B.2 Free Abelian Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
B.3 Finitely Generated Abelian Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
B.4 Homological Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
CONTENTS 2
Chapter 1
1.1 Homotopy
We start with the most important and fundamental concept, homotopy.
Ft
Definition 1.1.2 (Homotopic). If a homotopy exists between f and g, we say they are homotopic
and write
f ≃ g.
Remark. Later, we’ll not state that a map is continuous explicitly since we almost always assume
this in this context.
3
Lecture 1: Homotopies of Maps
Example (Straight line homotopy). Any two (continuous) maps with specification
f, g : X → Rn
Example. Let S 1 denotes the unit circle in R2 , and D2 denotes the unit disk in R2 . Then the
inclusion f : S 1 ,→ D2 is nullhomotopic.
1
t=0 t= 2
t=1
We see that there is a homotopy from f (x) to 0 (the zero map which maps everything to 0),
and since 0 is a constant map, hence it’s actually a nullhomotopy. ⊛
Remark. It will essentially flip the orientation, hence we can’t deform one to another contin-
uously.
x0
Answer. Consider
Ft (x) := (1 − t)x + tx0 ,
Exercise. Suppose
f1 g1
X Y Z
f0 g0
where
f0 ≃ f1 , g0 ≃ g1 .
Ft Gt
Show
g0 ◦ f0 ≃ g1 ◦ f1 .
X ×I → Y ×I → Z
(x, t) 7→ (Ft (x), t) 7→ Gt (Ft (x)).
Remark. Noting that if one wants to be precise, you need to check the continuity of this
construction.
Ft (x) : X × I → Y
Remark. The continuity of Ft is an even stronger condition for the continuity of Ft for a fixed
t.
Definition 1.1.4 (Homotopy relative). Given two spaces X, Y , and let B ⊆ X. Then a homotopy
Ft (x) : X → Y is called homotopy relative B (denotes rel B) if Ft (b) is independent of t for all b ∈ B.
Example. Given X and B = {0, 1}. Then the homotopy of paths from [0, 1] → X is rel{0, 1}.
F0 (t)
F0 (0) F0 (1)
Definition 1.2.4 (Homotopy type). If X, Y are homotopy equivalent, then we say that they
have the same homotopy type.
∗
n
g
D
which is homotopic to idDn by straight line homotopy Ft (x) = tx. Specifically, we see that this
holds for any convex set. ⊛
(a) X is contractible.
(b) ∀x ∈ X, idX ≃ cx .
(c) ∃x ∈ X, idX ≃ cx .
Proof. We see that 2. ⇒ 3. is obvious. We consider 3. ⇒ 2. This follows from the following general
lemma.
f = idX ◦f ≃ cx ◦ f = cx ◦ g ≃ idX ◦g = g.
■
Then, from this Lemma 1.2.1, we see that assuming x0 ∈ X such that idX ≃ cx0 , then consider cx
for all x ∈ X, then from Lemma 1.2.1, we see that cx ≃ idX .
To show 3. ⇒ 1., we let x0 ∈ X such that idX ≃ cx0 .
f
X {∗}
g
g◦f: X →X
x 7→ x0 ,
g ◦ f = cx0 ≃ idX .
i r
B X B
r◦i
Answer. Suppose X is path-connected. Then we see that given two points x1 and x2 in X, there
exists a path γ(t) with
γ : [0, 1] → X, γ(0) = x1 , γ(1) = x2 .
Since X ≃ Y , then there exists a pair of f and g such that f : X → Y and g : Y → X with
f ◦ g ≃ idY , g ◦ f ≃ idX .
F G
Firstly, we let g(y1 ) =: x1 and g(y2 ) =: x2 . From the argument above, we know there exists such
a γ starting at x1 = g(y1 ) ending at x2 = g(y2 ). Now, consider f (γ(t)) = (f ◦ γ)(t) such that
we immediately see that y1′ and y2′ is path connected. Now, we claim that y1 and y1′ are path
connected in Y , hence so are y2 and y2′ . To see this, note that
f ◦ g ≃ idY ,
F
Since F is continuous in I, we see that there must exist a path connects y1 and y1′ . The same
argument applies to y2 and y2′ . Now, we see that the path
y1 → y1′ → y2′ → y2
I I
0 1 0 1
f
γ y1
Ft (y1 ) γ′
g(y1 ) = γ(0) =: x1 y2
y1′
f Ft (y2 )
y2′
g(y2 ) = γ(1) =: x2
g
X Y
⊛
One can further show that the connectedness is also preserved by any homotopy equivalence.
Figure 1.5: The deformation retraction of D2 \ {0} is just to enlarge that hole and push all the
interior of D2 to the boundary, which is S 1 .
Proof. For a cylinder, consider X × I → X. Define homotopy on a closed rectangle, then verify it
induces map on quotient.
For a Möbius band, we define a homotopy on a closed rectangle, then verify that it respect the
equivalence relation.
Finally, we use the universal property of quotient topology to argue that we get a homotopy on
Möbius band.
Figure 1.6: The deformation retraction for Cylinder and Möbius band
Remark. We see that Möbius band ≃ S 1 ≃ cylinder, hence the orientability is not homotopy
invariant.
• homotopy equivalence
• homotopy invariants
– path-connectedness
• not invariant
– dimension
– orientability
– compactness
1.3 CW Complexes
Example (Constructing spheres). We now see how to construct S 1 and S 2 from ground up.
• S 1 (up to homeomorphisma )
• S2
– glue boundary of 2-disk to a point
– glue 2 disks onto a circle
Figure 1.7: Left: Glue a 2-disk to a point along its boundary. Right: Glue 2 disks to S 1 .
The gluing instruction to construct S 2 in the right-hand side can be demonstrated as follows.
top hemi-sphere
bottom hemi-sphere
Notation. Let Dn denotes a closed n-disk (or n-ball) Dn ≃ {x ∈ Rn : ∥x∥ ≤ 1}, and let S n denotes
an n-sphere S n ≃ x ∈ Rn+1 : ∥x∥ = 1 .
Definition 1.3.2 (Cell). We call a point as a 0-cell, and the interior of Dn Int(Dn ) for n ≥ 1
as an n-cell.
Definition 1.3.3 (Skeleton). The 0-skeleton X 0 is a set of discrete points, i.e, 0-cell.
We inductively construct the n-skeleton X n from X n−1 by attaching n-cells enα , where α is the
index. Specifically, we follow the gluing instructions called attaching maps defined below and
obtain X n from X n−1 by
!
a
n−1 n .
Xn = X Dα
x ∼ φ (x)
α α
Remark. We write X (n) for n-skeleton if we need to distinguish from the Cartesian prod-
uct.
Definition 1.3.4 (Attaching map). The attaching map for every cell enα is a continuous map φα
such that
φα : ∂Dαn → X n−1 .
Finally, we let X be defined as [
X= X n,
n=0
Definition 1.3.5 (Weak topology). The weak topology, denoted as w, contains open set u such
that
u ⊆ X is open ⇔ ∀n u ∩ X n is open .
If all cells have dimension less than N and there exists an N -cell, then X = X N and we call it
N -dimensional CW complex.
Notation. Notice that we call the real projection space as RP , and we also have so-called
complex projection space, denote as CP .
α × eα : eα is an m-cell on X, eα is an n-cell on Y } .
{em n m n
Remark. The product topology may not agree with the weak topology on the X × Y . However,
they do agree if X or Y is locally compact or if X and Y both have at most countably many cells.
Remark. X ∨ Y is a CW complex.
1.4.3 Quotients
Let X be a CW complex, and A ⊆ X subcomplex (closed union of cells), then X / A is a quotient space
collapse A to one point and inherits a CW complex structure.
Remark. X / A is a CW complex.
ϕα quotient
Sn Xn X n / An
Example. We can take the sphere and squish the equator down to form a wedge of two spheres.
X X /A
A
A
Example. We can take the torus and squish down a ring around the hole.
X X /A
≃ ≃
A A
We see that X / A is homotopy equivalent to a 2-sphere wedged with a 1-sphere via extending the
orange point into a line, and then sliding the left point to the line along the 2-sphere towards the
other points, forming a circle.
Category Theory
f ◦ idX = f, idX ◦g = g
16
Lecture 6: A Foray into Category Theory
C Ob(C) Mor(C)
set Sets X All maps of sets
fset Finite sets All maps
Gp Groups Group Homomorphisms
Ab Abelian groups Group Homomorphisms
k−vect Vector spaces over k k-linear maps
Rng Rings Ring Homomorphisms
Top Topological spaces Continuous maps
Haus Hausdorff Spaces Continuous maps
hTop Topological spaces Homotopy classes of continuous maps
Top∗ Based topological spacesa Based mapsb
a Topological spaces with a distinguished base point x0 ∈ X
b Continuous maps that presence base point f : (X, x0 ) → (Y, y0 ) such that f : X → Y, f (x0 ) = y0 is continuous.
Remark. From Definition 2.1.1, we see that a category is just any directed diagram plus compo-
sition law and identities.
idA A B idB .
∀g1 , g2 f ◦ g1 = f ◦ g2 ⇒ g1 = g2 .
g1
f
A M N
g2
∀g1 , g2 g1 ◦ f = g2 ◦ f ⇒ g1 = g2 .
g1
f
M N B
g2
Lemma 2.1.1. In set, Ab, Top, Gp, a map is monic if and only if f is injective, and epic if and only
if f is surjective.
Proof. In set, we prove that f is monic if and only if f is injective. Suppose f ◦ g1 = f ◦ g2 and f
is injective, then for any a,
hence g1 = g2 .
Now we prove another direction, with contrapositive. Namely, we assume that f is not injective
and show that f is not monic. Suppose f (a) = f (b) and a ̸= b, we want to show such gi exists.
This is easy by considering
g1 : ∗ 7→ a, g2 : ∗ 7→ b.
■
2.1.1 Functor
After introducing the category, we then see the most important concept we’ll use, a functor. Again, we
start with the definition.
F : Ob(C ) → Ob(D)
X 7→ F (X);
such that
• F (idX ) = idF (x) ;
• F (f ◦ g) = F (f ) ◦ F (g).
Lecture 7: Functors
Example (Applying a covariant functor). Assume that we initially have a commutative diagram in C 21 Jan. 10:00
as
f
X Y
g
g◦f
Z
and a functor F : C → D. After applying F , we’ll have
F (f )
F (X) F (Y )
F (g)
F (g◦f )=F (g)◦F (f )
F (Z)
F : Ob(C ) → Ob(D)
X 7→ F (X).
such that
Example (Applying a contravariant functor). Then, we see that in this case, when we apply a con-
travariant functor F , the diagram becomes
F (f )
F (X) F (Y )
F (g)
F (g◦f )=F (f )◦F (g)
F (Z)
Example (Identity functor). Define I as I : C → C such that it just send an object C ∈ C to itself.
[f : G → H] 7→ [f : G → H] .
[f : X → Y ] 7→ [f : X → Y ] .
aG is now just the underlying set of the group G.
bX is now just the underlying set of the topological space X.
set → k−vect
s 7→ "free" k-vector space on s
T : k−vect → k−vect
V 7→ V ∗ = Homk (V, k).
If we are working on a basis, we can then represent T as a matrix A, and we further have
A 7→ A⊤ .
π1 : Top∗ → Gp
(X, x0 ) 7→ π1 (X, x0 ),
Hp : Top → Ab
X 7→ Hp (X),
S FS
∃!f : group homorphism
f
G
where U (G) is the forgetful functor from the category of groups to the category of sets. This is
the statement that the free functor and the forgetful functor are adjoint; specifically that the free
functor is the left adjoint (appears on the left in the Hom above).
Remark. Whenever we state a universal property for an object (plus a map), an object (plus a map)
may or may not exist. If such object exists, then it defines the object uniquely up to unique
isomorphism, so we can use the universal property as the definition of the object (plus a map).
Lemma 2.2.1. Universal property defines FS (plus a map S → F (S) ) uniquely up to unique
isomorphism.
Proof. Fix S. Suppose
S → FS , S → FeS
both satisfy the unique property. By universal property, there exist maps such that
S FeS S FS
∃!φ ∃!ψ
f f
FS FeS
We’ll show φ and ψ are inverses (and the unique isomorphism making above commute). Since we
FS FeS
f f
S idFS S idFe
S
f f
FS FeS
FS FeS
f ψ f
φ
φ◦ψ=idFS S FS ψ◦φ=idFe
S FeS S
φ ψ
f f
FS FeS
where the identity makes these outer triangles commute, then by the uniqueness in universal prop-
erty, we must have
φ ◦ ψ = idFS , ψ ◦ φ = idFeS ,
so φ and ψ are inverses (thus group isomorphism). ■
Definition 2.2.3 (Word). Fix a set S, and we define a word as a finite sequence (possibly ∅)
in the formal symbols −1
s, s | s ∈ S .
Then we see that elements in FS are equivalence classes of words with the equivalence relation
being
• deleted ss−1 or s−1 s. i.e.,
vs−1 sw ∼ vw
vss−1 w ∼ vw
(c) Check that FS satisfies the universal property with respect to the map
S → FS , s 7→ s.
The fundamental group is the first and the simplest homotopy group, which records information about
the basic shape, or holes, of the topological space. Not surprisingly, this is already interesting enough
to study and can help us distinguish between different shapes just by finding the spaces’ corresponding
fundamental groups.
3.1 Path
We start with the definition.
γ: I → X
Definition 3.1.2 (Homotopy path). A homotopy of paths γ0 , γ1 is a homotopy from γ0 to γ1 rel{0, 1}.
γt
Definition 3.1.3 (Path composition). For paths α, β in X with α(1) = β(0), the composition a α · β
is
1
α(2t), if t ∈ 0,
2
(α · β)(t) :=
1
β(2t − 1), if t ∈
,1 .
2
23
Lecture 8: The Fundamental Group π1
Remark. By the pasting lemma, this is continuous, hence α · β is actually a path from α(0) to β(1).
Answer. We show that γ and γ ◦ ϕ are homotopic rel{0, 1} by showing that there exists a continuous
Ft such that
F0 = γ, F1 = γ ◦ ϕ.
Notice that since ϕ is continuous, so we define
We see that
F0 (x) = γ(x), F1 (x) = γ ◦ ϕ(x),
and also, we have
Ft (x) ∈ X
for all x, t ∈ I.
Now, we check that Ft really gives us a homotopic rel{0, 1}. We have
which shows that 0 and 1 are independent of t, hence γ and γ ◦ ϕ are homotopic rel{0, 1}. ⊛
Exercise. Fix x1 , x1 ∈ X. Then homotopy of paths (relative {0, 1}) is an equivalence relation on
paths from x0 to x1 .
Definition 3.2.1 (Fundamental group). Let X denotes the space and let x0 ∈ X be the base point.
The fundamental group of X based at x0 , denoted by π1 (X, x0 ), is a group such that
• Elements: Homotopy classes rel{0, 1} of paths [γ] where γ is a loop with γ(0) = γ(1) = x0 a
x0
γ
γ : I → X, t 7→ x0
• Inverses: The inverse [γ]−1 of [γ] is represented by the loop γ such that
x0 x0
γ(t) γ(1 − t)
a We say γ is based at x0 .
Proof. We actually need to prove that the defined π1 actually is a group, hence, we prove that
and
1
γ1 (4t), t ∈ 0, ;
1 4
(γ · γ )(2t), t ∈ 0, ;
1 2
2
1 1
(γ1 · γ2 ) · γ3 (t) = = γ2 (4t − 1), t ∈ , ;
1 4 2
γ3 (2t − 1),
t∈ ,1
2 1
γ3 (2t − 1), t ∈ ,1 .
2
Then, we define ϕ : I → I such that
1 1
2t ∈ 0, , t ∈ 0, ;
2 4
1 1 3 1 1
ϕ(t) = t + ∈ , , t∈ , ;
4 2 4 4 2
t+1 3 1
∈ ,1 , t ∈ ,1 .
2 4 2
Inverses. We want to show that γ · γ ≃ c, where γ(t) = γ(1 − t). Firstly, we have
1
γ(2t), t ∈ 0, ;
2
(γ · γ)(t) =
1
γ(1 − 2t), t ∈
,1 .
2
We consider Ft given by
1
γ(2xt),
x ∈ 0, ;
2
Ft (x) =
1
γ(1 − 2xt), x ∈
,1 .
2
If t = 0, we have
1
γ(0),
x ∈ 0, ;
2
F0 (x) = = x0
1
γ(1), x ∈
,1
2
for all x ∈ I, namely F0 = c, while when t = 1, we have
1
γ(2x),
x ∈ 0, ;
2
F1 (x) = = (γ · γ)(x),
1
γ(1 − 2x), x ∈
,1
2
γ
x0 x0 x0
γ
γ γ
t
γ
γ γ
Figure 3.1: Illustration of Ft . Intuitively, the path γ · γ is x0 → x0 → x0 . But now, Ft is
γ γ
x0 → t → x0 . We can think of this homotopy is pulling back the turning point along the original
path.
∀x0 , x1 ∈ X π1 (X, x0 ) ∼
= π1 (X, x1 ).
Proof. To show that the change-of-basepoint map is isomorphism, we show that it’s one-to-one and
onto.
• One-to-one. Consider that if [h · γ · h] = [h · γ ′ · h], then since we know that h−1 = h, hence
in the fundamental group π1 (X, x0 ), we see that
h · h · γ · h · h = h · h · γ ′ · h · h. ⇒ γ = γ ′
as we desired.
• Onto. We see that for every α ∈ π1 (X, x0 ), there exists a γ ∈ π1 (X, x0 ) such that
γ = h · α · h ∈ π1 (X, x1 )
since h · γ · h = α.a
We then see that the fundamental group of X does not depend on the choice of basepoint, only
on the choice of the path component of the basepoint. If X is path-connected, it now makes sense to
refer to the fundamental group of X and write π1 (X) for the abstract group (up to isomorphism).
■
a Notice that this is indeed the case, one can verify this by the fact that h : x0 → x1 and h : x1 → x0 .
Remark. We see that we can write π1 (X) up to isomorphism if X is path-connected from Theo-
rem 3.2.1.
Lemma 3.2.1. Given x0 , x1 , x2 ∈ X, α, α′ are two paths from x0 to x1 , and β, β ′ are two paths
from x1 to x2 . If ⟨α⟩ = ⟨α′ ⟩, ⟨β⟩ = ⟨β ′ ⟩, then ⟨α · β⟩ = ⟨α′ · β ′ ⟩.
α · β ≃ α′ · β ′ rel{0, 1}.
This is done by using homotopy H : I × I → X such that it combines F (2s, t) and G(2s − 1, t).
α β
x0 x1 x2
α′ β′
α′ β′
F (2s, t) G(2s − 1, t)
α β
α′ β′ γ′
α β γ
Lemma 3.2.3. Let X be a topological space, and x0 ∈ X. Then for every path homotopy ⟨α⟩ from
x1 to x2 , we have
⟨cx1 · α⟩ = ⟨α⟩ = ⟨α · cx2 ⟩.
Proof. We only need to prove cx1 · α ≃ α rel{0, 1}. The homotopy can be written out explicitly by
the following diagram.
α
1
0
cx1 α
cx1
1
cα−1 (1−t)
t From α−1 (0) to α−1 (1 − t)
cx : [0, 1] → X
t 7→ x
as a constant loop.
Remark. We’ll soon see that for any topological space x, Definition 3.2.1 defines a groupoid, denoted
by Π(X).
Definition 3.2.3 (Fundamental groupoid). Let X denotes the space, then the category Π(X) is a
fundamental groupoid of X such that
• Ob(Π(X)) := X
• Identity: For every p ∈ Ob(Π(X)) = X, we define 1p := ⟨cp ⟩ ∈ HomΠ(X) (p, p) be the constant
loop based at p such that for every ⟨α⟩ ∈ HomΠ(X) (p, q),
⟨α⟩ ⟨γ⟩
p q ⟨β⟩ r s
Then
⟨γ⟩ ◦ (⟨β⟩ ◦ ⟨α⟩) = (⟨γ⟩ ◦ ⟨β⟩) ◦ ⟨α⟩.
Proof. Note that in Definition 3.2.3, we need to show some definitions are indeed well-defined, and
we also need to show that Π(X) is actually a groupoid.
• Composition: Since if α ≃ α′ , β ≃ β ′ , we have
α · β ≃ α′ · β ′
Additionally, from Lemma 3.2.4, we see that given α is a path from p to q, then
( −1
⟨α · α⟩ = ⟨cq ⟩ =: idq
⟨α · α−1 ⟩ = ⟨cp ⟩ =: idp .
Furthermore, since ⟨α−1 · α⟩ = ⟨α⟩ ◦ ⟨α−1 ⟩ and ⟨α · α−1 ⟩ = ⟨α−1 ⟩ ◦ ⟨α⟩, hence this means Π(X) is
indeed a groupoid. ■
such that
(f, g) 7→ f · g := g ◦ f.
We can prove that
(HomC (x, x), ·)
defines a group AutC (x) called the isotropy group of C at x.
Exercise. For every x, y ∈ Ob(C ), if there exists f ∈ HomC (x, y), then f induces
≃
f∗ : AutC (x) → AutC (y),
Hence, we only need to verify their group composition agrees. But this is trivial, since for every
two ⟨α⟩, ⟨β⟩ ∈ AutΠ(X) (p),
π1 (S 1 ) ∼
= Z,
p : R → S1 φ : Z → π1 (S 1 , 1)
and
x 7→ e2πix n 7→ ⟨p ◦ γn ⟩
where p defined above is a covering map. We need to show that this is well-defined.
From the definition of φ, we see that it’s a homomorphism. But we also need to show
Remark. Intuitively, this winds around S 1 n times. The key to this proof was to understand S 1 via
the covering space R → S 1 . We will talk about covering spaces much later.
π(X × Y, (x0 , y0 )) ∼
= π1 (X, x0 ) × π1 (Y, y0 )
f f
Proof. Let Z → X × Y with z 7→ (fX (z), fY (z)). Then we have
• Paths I → X × Y .
• Homotopies of paths I × I → X × Y .
S 1 × S 1 × · · · × S 1 = (S 1 )k ,
| {z }
k times
π1 (S 1 )k ∼= Zk .
b a
∼
=
b
Remark. One way to think of the k-torus is as a k-dimensional cube with opposite (k−1)-dimensional
faces identified by translation.
π1 (Y, f (x0 ))
f∗
π1 (X, x0 ) γ∗
∼
=
g∗
π1 (Y, g(x0 ))
γ∗ ◦ f∗ (⟨α⟩) = g∗ (⟨α⟩).
F1 = g
x0
X ×I α
g(x0 ) Y
g◦α
(x0 , t) F F (x0 , t)
f ◦α
x0 α f (x0 )
F0 = f
We see that we can obtain a homotopy G : I × I → Y such that
G := F ◦ (α × id),
where we define α × id by
G1 g(x0 ) Y
1
g◦α
G
t γ
0 f ◦α
G0 f (x0 )
g(x0 )
g(x0 )
g(x0 )
g(x0 )
g◦α
1
γ −1
G g(x0 ) → γ(t)
t
γ
γ(t) → g(x0 )
0
γ −1 f ◦ α γ
Theorem 3.3.3 (Fundamental group is a homotopy invariant). If X, Y are homotopy equivalent, then
their fundamental groups are isomorphic.
Proof. ■ HW.
Example. π1 (S ∞ × S 1 ) ∼
= Z.
R2 \ {0} ∼
= S 1 × R,
which means that the generators are just loops around the hole intuitively.
π1 : Top∗ → Gp
(X, x0 ) 7→ π1 (X, x0 ).
f∗ : π1 (X, x0 ) → π1 (Y, y0 )
[γ] 7→ [f ◦ γ]
i.e.,
[f : X → Y ] 7→ [f∗ : π1 (X, x0 ) → π1 (Y, f (x0 ))] .
Proof. We need to check
• well-defined on path homotopy classes.
• f∗ is a group homomorphism.
1
f (α(2s)),
if s ∈ 0,
2
f∗ (α · β) = f∗ (α) · f∗ (β) =
1
if s ∈
f (β(1 − 2s)),
,1 .
2
• id(X,x0 )
∗
= idπ1 (X,x0 )
• (f∗ ◦ g∗ ) = (f ◦ g)∗
(X, x0 ) π1 (X, x0 )
f f∗
(Y, y0 ) π1 (Y, y0 )
■
Remark. We usually write f∗ if it’s a covarant functor, while writing f ∗ if it’s a contravariant
functor.
Remark. We see that the construction of fundamental group is actually constructing a functor.
Specifically,
π1 : Top∗ → Gp
such that
• on objects:
• on morphisms:
Definition 3.4.1 (Category of groupoid). The category of groupoid, denoted as Gpd, contains the
following data.
• Ob(Gpd): groupoids.
• Hom(Gpd): functors between groupoids.
• Composition: For every X, Y, Z ∈ Ob(Gpd),
F G
X Y Z
– on objects: ∀X ∈ Ob(X),
G ◦ F (X) := G(F (X)).
– on morphisms: ∀X, Y ∈ Ob(X) and f : X → Y ,
G ◦ F (f ) := G(F (f )).
• Associativity. Since the composition is defined based on two functors,a this holds trivially.
a For F G
example, given X → Y → Z.
Proof. We need to show that the composition is well-defined. Specifically, we need to check
f g
X1 X2 X3
G ◦ F (g) ◦ G ◦ F (f ) = G(F (g ◦ f )) = G ◦ F (g ◦ f ).
Π : Top → Gpd,
where
• on objects: For every X ∈ Ob(Top),
X 7→ Π(X);
such that
– on objects: For every p ∈ Ob(Π(X)) = X, Π(f )(p) = f (p). i.e.,
⟨α⟩ ⟨β⟩
p q r
we want to show Π(f ) (⟨β⟩ ◦ ⟨α⟩) = Π(f )(⟨β⟩) ◦ Π(f )(⟨α⟩). Indeed, since
and
Π(f )(⟨β⟩) ◦ Π(f )(⟨α⟩) = ⟨f ◦ β⟩ ◦ ⟨f ◦ α⟩ = ⟨(f ◦ α) · (f ◦ β)⟩.
Since ⟨f ◦ (α · β)⟩ = ⟨(f ◦ α) · (f ◦ β)⟩, hence Π(f ) is well-defined.
and
Π(f ) Π(g)
Π(X) Π(Y ) Π(Z)
Definition 3.5.1 (Free product). Given some collections of groups {Gα }α , the free product, denoted
by ∗Gα is a group such that
α
wgi gj v ∼ w(gi gj )v
when both gi , gj ∈ Gα . Also, for the identity element id = eα ∈ Gα for any α such that
weα v ∼ wv.
Specifically,
∗α Gα := {words in {Gα }α } ∼.
.
Remark. In particular, we have the following universal property of ∗α Gα . For every α, there is a
ια such that
ια : Gα → ∗α Gα , g 7→ g,
where ια is a group homomorphism, obviously. Further, (∗α Gα , ια ) satisfies the following property:
For every group H and a group homomorphism φα : Gα → G for all α, there exists a unique group
homomorphism φ : ∗α Gα → H such that φ ◦ ια = φα , i.e., the following diagram commutes.
Gα1
ια1
Gα2 ια2 ∗α Gα
φα1
.. φα2
∃!φ
.
• φ is a group homomorphism.
• φ ◦ ια = φα .
• Such φ is unique. Suppose there exists another ψ : ∗α Gα → H, then
ψ ◦ ι α = φα ⇒ ∀ ψ(g) = ψα (g),
g∈Gα
which is just φ.
Remark. We further claim that this universal property determines such free product uniquely. i.e.,
assume there are another group Ge and eια : Gα → G.e Assume (G, ια ) also satisfies the following
e e
property: For every group H and group homomorphism φα : Gα → H, then there exists a unique
group homomorphism φ : G e → H such that the following diagram commutes.
Gα1
ια1
e
Gα2 ια2
e
∗α G
eα
φα1
.. φα2
∃!φ
.
e∼
Then, G = ∗α Gα .
Proof. Assume (G, ια ) satisfies the universal property mentioned above. Then from the universal
e e
property and viewing Ge and ∗α Gα as H separately, we obtain the following diagram.
∃!f
G
e ∗α Gα
ια1
e ∃!g ια1
ια2
e ια2
Gα1
Gα2
..
.
We claim that
g ◦ f = id, f ◦ g = id .
To see this, we simply apply the same observation, for example,
Gα1
ια1
e
Gα2 ια2
e
G
e
ια1
e
∃!g◦f
..
. ια2
e
G
e
where g ◦ f comes from the previous diagram. But notice that id let the diagram commutes also,
and since it’s unique, hence g ◦ f = id. Similarly, we have f ◦ g = id. ⊛
If you’re careful enough, you may find out that all we’re doing is just writing out a specific example
of Lemma 2.2.1! Indeed, this is exactly the construction of a free group.
Y p2 W
Y W
is a Cocartesian diagram.
Remark. If we reverse all the directions of morphism, then we have so-called fibered product.
Example. Let C = Top, and let X ∈ Ob(Top). Given X0 , X1 ∈ X, and Int(X0 ) ∪ Int(X1 ) = X, if
we have
i0 : X0 ,→ X, i1 : X1 ,→ X
j0 : X0 ∩ X1 ,→ X0 , j1 : X0 ∩ X1 ,→ X1 ,
then
j0
X0 ∩ X1 X0
j1 i0
X1 i1
X
is a cocartesian diagram.
Proof. All we need to show is that given a topological space Y ∈ Top and f : X0 → Y , g : X1 → Y
in Top, we have
f ◦ j0 = g ◦ j1 .
j0
X0 ∩ X1 X0
j1 i0
f
X1 i1
X
∃!h
g
Y
We simply define h : X → Y , x 7→ h(x) such that
(
f (x), if x ∈ X0 ;
h(x) =
g(x), if x ∈ X1 .
h is clearly well-defined since the diagram commutes, so if x ∈ X0 ∩ X1 , then f (x) = g(x). The only
thing we need to show is that h is continuous. But this is obvious too since X = Int(X0 ) ∪ Int(X1 ),
and
h|Int(X0 ) = f |Int(X0 ) , h|Int(X1 ) = g|Int(X1 ) .
The uniqueness is trivial, hence this is indeed a cocartesian diagram. ⊛
Example. Let C = Top∗ . Given p ∈ X0 ∩ X1 , where all other data are the same with the above
example, we see that
j0
(X0 ∩ X1 , p) (X0 , p)
j1 i0
(X1 , p) i1
(X, p)
is a cocartesian diagram.
Example. Let C = Gp. Given P, G, H ∈ Ob(Gp), we claim that the fibered coproduct of i and j
exists.
i
P G
j
H
Consider G ∗ H be the free product between G and H, with two inclusions
ι1 : G ,→ G ∗ H, ι2 : H ,→ G ∗ H.
i
P G
j ι1
H ι2 G∗H
Let
N := ⟨ ι1 ◦ i(x) · (ι2 ◦ j(x))−1 | x ∈ P ⟩,
we define
G ∗p H = G ∗ H N .
.
i
P G
j ι1
τ
H ι2 G∗H
π
ν
G ∗p H
We claim that
i
P G
j τ
H ν G ∗p H
Proof. Firstly, since it’s just an outer diagram from above, hence it commutes. So we only need
to prove this diagram satisfies the second diagram. Given any group K, for every f : G → K,
g : H → K such that the following diagram commutes.
i
P G
j τ
f
H ν G ∗p H
h
g
K
We want to prove that there exists a unique h : G ∗p H → K such that this diagram still commutes.
The idea is simple, from the universal property of G∗H, we see that there exists a unique e
h : G∗H →
K such that
h ◦ ι1 = f, e
e h ◦ ι2 = g.
i
P G
ι1
j G∗H τ
ι2 π f
∃!e
h
H ν G∗H
h
g
G ∗p H
G ∗p H
We then see that there exists a unique h : G ∗p H → K such that the above diagram commutes. ⊛
Definition 3.5.3 (Free product with amalgamation). If two groups Gα and Gβ have a common sub-
group S{α,β} ,a given two inclusion mapsb iαβ : S{α,β} → Gα and iβα : S{α,β} → Gβ , the free product
with amalgamation α ∗S Gα is defined as ∗Gα modulo the normal subgroup generated by
α
Namely,c
∗Gα
.
α∗S Gα = α ⟨iαβ (s{α,β} )iβα (s{α,β} )−1 ⟩
and satisfies the universal property described by the following commutative diagram.
iαβ
S Gα
iβα
Gβ Gα ∗S Gβ
∃!
X
a In general, we don’t need S{α,β} to be a subgroup.
b We don’t actually need iαβ , iβα to be inclusive as well.
c i.e., i
αβ (s) and iβα (s) will be identified in the quotient.
iα
A∩B A
iβ
B A∪B
∃!
Theorem 3.6.1 (Seifert-Van Kampen Theorem). Given (X, x0 ) such that X = Aα with
S
α
∗ : π1 (Aα , x0 ) → π1 (X, x0 ).
α
If we additionally have Aα ∩ Aβ ∩ Aγ where they are all path-connected for every α, β, γ, then
π1 (X, x0 ) ∼
=α∗π1 (Aα ∩Aβ ,x0 ) π1 (Aα , x0 )
associated to all maps πa (Aα ∩Aβ ) → π1 (Aα ), π1 (Aβ ) induced by inclusions of spaces. i.e., π1 (X, x0 )
is a quotient of the free product ∗α π1 (Aα ) where we have
which is induced by the inclusion iαβ : Aα ∩ Aβ → Aα . We then take the quotient by the normal
subgroup generated by
{(iαβ )∗ (γ)(iβα )∗ | γ ∈ π1 (Aα ∩ Aβ )} .
We’ll defer the proof of Theorem 3.6.1 until we get familiar with this theorem.
A∩B X =A∪B
γ
x0
α
β
A
B
Intuitively we see the fundamental group of X, which is built by gluing A and B along their
intersection. As the fundamental group of A and B glued along the fundamental group of their
intersection. In essence, π1 (X, x0 ) is the quotient of π1 (A) ∗ π1 (B) by relations to impose the
condition that loops like γ lying in A ∩ B can be viewed as elements of either π1 (A) or π1 (B).
Remark. We canSuse a more abstract way to describe Theorem 3.6.1. Specifically, in the case that
2
n = 2, i.e., X = i=1 Ai , we let Ai =: Xi , then we have the following. The functor π1 : Top∗ → Gp
maps the cocartesian diagram in Top∗ to a cocartesian diagram in Gp as follows.
j0 (j0 )∗
(X0 ∩ X1 , x0 ) (X0 , x0 ) π1 (X0 ∩ X1 , x0 ) π1 (X0 , x0 )
π1
j1 i0 7−→ (j1 )∗ (i0 )∗
(X1 , x0 ) i1
(X, x0 ) π1 (X1 , x0 ) π1 (X, x0 )
(i1 )∗
π1 (X, x0 ) ∼
= π1 (X0 , x0 ) ∗π1 (X0 ∩X1 ,x0 ) π1 (X1 , x0 ).
Additionally, there is a more general version of Theorem 3.6.1, which is defined on groupoid. The
theorem is stated in Appendix A.1 with the proof.
With this more general version and the proof of which, we can apply it to Theorem 3.6.1. But one
question is that, the above proof works in Gpd rather than in Gp. We now see how to generalize a group
to a groupoid.
For any group G, we can define a groupoid, denoted as G also, as follows.
• Ob(G) = {∗}, a one point set.
• Hom(G) = {g ∈ G}.
• Composition: We define g ◦ h := h · g.
We see that the associativity of group elements implies the associativity of composition defined above,
and since there is an identity element in G, hence we also have an identity morphism, these two facts
ensure that G is a category.
Furthermore, since for every g ∈ G, there is a g −1 ∈ G, hence every morphism is an isomorphism,
which implies G is a groupoid.
With this, we see that we can view the following diagram in the category of groupoid Gpd.
(j0 )∗
π1 (X0 ∩ X1 , x0 ) π1 (X0 , x0 )
(j1 )∗ (i0 )∗
π1 (X1 , x0 ) π1 (X, x0 )
(i1 )∗
And to prove Theorem 3.6.1, we only need to show this diagram is cocartesian. This version of proof
is given in Appendix A.2.
We see that π1 (S 2 ) must be a quotient of π1 (A) ∗ π1 (B), but since A, B ≃ D2 , we know that
π1 (A) and π1 (B) are both zero groups, thus π1 (A) ∗ π1 (B) is the zero group, and π1 (S 2 ) is also the
zero group.
Remark. Note that the inclusion of A ∩ B ,→ A induces the zero map π1 (A ∩ B) → π1 (A),
which cannot be an injection. In fact, we know that π1 (A ∩ B) ∼
= Z since A ∩ B ≃ S 1 .
Example (Fundamental group of T ). We can use Seifert Van Kampen Theorem to compute the
fundamental group of a torus T .
Now note that A ≃ D2 and B ≃ S 1 ∨ S 1 , and since it’s a thickening of the two loops around the
torus in both ways, this suggests the question of how do we find π1 (B)? We grab a bit of knowledge
from Seifert Van Kampen Theorem before we continue.
Exercise. Suppose we have path-connected spaces (Xα , xα ), and we take their wedge sum
by identifying the points xα to a single point x. We also suppose a mild condition for
W
α X α
all α, the point xα is a deformation retract of some neighborhood of xα .
For example, this doesn’t work if we choose the bad point on the Hawaiian earring. Then
we can use Seifert Van Kampen Theorem to show that
_
π1 Xα , x ∼= ∗π1 (Xα , xα ) .
α α
X1 U3
U2
X1 ∪ U2 ∪ U3 ≃ X1
• π1 (B) = π1 (S 1 ∨ S 1 ) = Z ∗ Z = F2
• π1 (A ∩ B) = π1 (S 1 ) = Z
Further, we know that π1 (A ∩ B) → π1 (A) is the zero map. We need to understand π1 (A ∩ B) →
π1 (B). To do so we need to understand how we’re able to identify π1 (S 1 ∨ S1 ) with F2 and how we
identify π1 (S 1 ) with Z. We update our Figure 3.4 to talk about this.
B b
γ
a A a−1
b−1
π1 (A ∩ B) → π1 (B) ∼
= Fa,b
γ 7→ aba−1 b−1 .
By Seifert Van Kampen Theorem, we identify the image of γ in π1 (B) as [aba−1 b−1 ] with its image
in π1 (A), which is just trivial. Therefore, we have
.
π1 (T 2 ) = Fa,b ⟨aba−1 b−1 ⟩ ∼
= Z2 .
Example. Let’s see the last example which illustrate the power of Seifert Van Kampen Theorem.
Start with a torus, and we glue in two disks into the hollow inside.
Glue in 2 disks
Proof. We can place a CW complex structure on this space so that each disk is a subcomplex.
Then, we take quotient of each disk to a point without changing the homotopy type, hence X is
homotopy to
By the same property, we can expand one of those points into an interval, and then contract the
red path as follows.
π1 (X) = π1 (S 2 ∨ S 2 ∨ S 1 ) = 0 ∗ 0 ∗ Z ∼
= Z.
Exercise. Consider R2 \ {x1 , . . . , xn }, that is the plane punctured at n points. Show that
π1 (X) ≃ Fn .
Answer. Observe that X ≃ n S 1 , so one way to do this is to convince yourself that you can do a
W
deformation retract the plane onto the following wedge.
Definition 3.7.3 (Relater). R is the set of relaters (words in a generator and inverses).
Definition 3.7.4 (Finite presentation). If S and R are both finite, then G = ⟨S | R⟩ is a finite
presentation if S, R are, and we say that G is finitely presented.
Note. One way to think about whether G is finitely presented is that if r is a word in R then r = 1,
where 1 is the identity of G.
■
a https://en.wikipedia.org/wiki/Isomorphism_theorems
Remark. The advantages of using group presentation are that given G = ⟨S | R⟩, it’s now easy to
define a homomorphism ψ : G → H given a map φ : S → H, ψ extends to a group homomorphism
G → H if and only if ψ vanishes on R, i.e., ψ(r) = 0 for all r ∈ R. We see an example to illustrate
this.
Remark. The disadvantages are that this is computationally very difficult. Let’s see an example
to illustrate this.
ψ : {a, b} → H
ψ(a)ψ(b)ψ(a)−1 ψ(b)−1 = 1H ∈ H.
Namely, this is a presentation of the trivial group, but this is entirely unclear.
(b) Gluing a 2-disk by its boundary along a word w in the generators kills w in π1 . We then get a
presentation for π1 (X 2 ) given by
(c) Gluing in any higher dimensional cells along their boundary will not change π1 . That is, in a CW
complex, we have π1 (X) = π1 (X 2 ).
(b) For each 2-disk Dα2 , write attaching map as word wα in ai . i.e., π1 (X 2 ) = ⟨ai | wα ⟩.
(c) π1 (X) = π1 (X 2 ).
Remark. Every group is π1 of some space. Specifically, given a group G, we work with its presen-
tation ⟨S | R⟩.
Proof. We see that we can simply take a loop and then wind a 2-disk around the loop a for n
times.
a
a
⊛
We then see that given a group G with presentation ⟨S | R⟩, one can construct a 2-dimensional
CW complex with π1 = G by
• Set X 1 = s∈S S 1
W
• For each relation r ∈ R, glue in a 2-disk along loops specified by the word r.
Proof. ■ HW
Theorem 3.7.3. Every connected graph has a maximal tree. Every tree is contained in a maximal
tree.
Proof. ■ DIY
Corollary 3.7.1. Suppose X is a connected graph with basepoint x0 . Then π1 (X, x0 ) is a free group.
Furthermore, we can give a presentation for π1 (X, x0 ) by finding a spanning tree T in X. The
generators of π1 will be indexed by cells eα ∈ X − T , and eα will correspond to a loop that passes
through T , traverses eα once, then returns to the basepoint x0 through T .
Proof. The idea is simple. X is homotopy equivalent to X / T via previous work on the homework,
T contains all the vertices, so the quotient has a single vertex. Thus, it is a wedge of circles, and
each eα projects to a loop in X / T .
T X
T
X
T a
b q
x0
X X /T
We now prove that the maximal trees exist. Recall that X is a quotient of X 0 α Iα . Since
`
each subset U is open if and only if it intersects each edge eα in an open subset. A map X → Y if
and only if its restriction to each edge eα is continuous. Now, take X0 to be a subgraph. Our goal
is to construct a subgraph Y with
• X0 ⊂ Y ⊂ X
• Y deformation retracts to X0
• Y contains all vertices of X.
We then see that X = Xi .a Now, let Y0 = X0 . By induction, we’ll assume that Yi is a subgraph
S
i Check.
of Xi such that
• Yi contains all vertices of Xi .
• Yi deformation retracts to Yi−1 .
We can then construct Yi+1 by taking Yi and adding to it one edge to adjoin every vertex of Xi+1 ,
namely [
Yi+1 := Yi one edge to adjoint every vertex of Xi ,
which is possible if we assume Axiom of Choice.
We then see that Yi+1 deformation retracts to Yi by just smashing down each edge. Now, we
can show that Y deformation retracts to Y0 = X0 by performing the deformation retraction from
Yi to Yi−1 during the time interval [1/2i , 1/2i−1 ]. ■
a Hatcher [HPM02] do this by arguing the union on the right is both open and closed.
Proof. We see that A1 ∩ A2 ≃ S n−1 , where we need n ≥ 2 to let S n−1 be connected. We then have
π1 (S n ) ∼
=0 ∗ 0 = 0.
π1 (A1 ∩A2 )
Covering Spaces
p|uα : ux → ux
..
.
X
e
u1
u2
u3
X
p
ux
x
Although we already investigate into covering spaces quite a lot in homework, but some terminologies
are still worth mentioning.
p|Vi : Vi → U
is a homeomorphism.
55
Lecture 14: Covering Spaces Theory
` (Slice).−1Given a covering space X and the relating map p, we call the parts Vi of
Definition 4.1.4 e
the partition i Vi of p (U ) slices.
Remark. We see that p is a covering map if and only if every point x ∈ X has a neighborhood which
is evenly covered.
We immediately have the following proposition.
Proposition 4.1.1 (Homotopy lifting property). The covering spaces satisfy the homotopy lifting
property such that the following diagram commutes.
F
e0
X × {0} Ye
∃!F
et
p
X ×I Ft
Y
γ
e
p
γ
x0 x
e0
Note. Though we can directly use Proposition 4.1.1 to prove this, but we can see some insight
by directly proving this.
X
e
γ
e
x
e0
∃!e
γ
p
I
0 1 γ X
γ(0)
γ
is an open cover of [0, 1]. Note that since [0, 1] is a compact metric space, from Lebesgue Lemmaa ,
there exists a partition of [0, 1] such that
such that for every i, [ti , ti+1 ] ⊂ γ −1 (Ux ) for some x. Without loss of generality, we assume that
[ti , ti+1 ] ⊂ γ −1 (Uxi ), i.e.,
γ([ti , ti+1 ]) ⊂ Uxi .
..
.
Ux2 α3
Ux2 α2
Ux2 α1
.. ..
. .
Ux1 α3 Ux3 α3
Ux1 α2 Ux3 α2
Ux1 α1 p Ux3 α1
p
p
I X
γ(t1 )
γ(0)
0 t1 t2 1 γ γ
Ux1 γ(t2 )
Ux2 Ux3
x
e0
Ux1 α1
p|Ux
1 α1
I X
γ(t1 )
γ(0)
0 t1 t2 1 γ|[0,t1 ]
Ux1
−1
Specifically, let γ
e1 (t) = p|Ux ◦ γ(t) for 0 ≤ t ≤ t1 , we see that
1 α1
e1 : [0, t1 ] → X
γ e
• γ
e1 (0) = x
e e0
• γ
e1 is continuous
e
• x
e0 ∈ Ux1 α1 ,
we see that γ
e1 (0, t1 ) ⊂ Ux1 α1 , which implies
e
γ
e
e1
[0, t1 ] Ux1 α1 −1
p|Ux ⇒γ
e1 = p|Ux
e ◦ γ|[0,t1 ] = γ
e1 ,
1 α1 1 α1
γ|[0,t
1]
Ux1
hence this lift is unique. Now, we see that we can simply repeat this argument, namely replacing ti
by ti+1 , γ
ei (ti ) by γ
ei+1 (ti+1 ) and so on. Since this partition is finite, hence in finitely many steps,
we obtain a unique path homotopy γ e by concatenating all γ
ei starting at xe0 .
γ1 ≃ γ2 rel{0, 1}
F
we’ll show that γ e2 rel{0, 1} where p ◦ Fe = F . Firstly, we denote x0 := γ1 (0) = γ2 (0), such
e1 ≃ γ
F
e
that
?
γ
e1 (1) = γ
e2 (1)
γ
e1
X
e
γ
e2
x
e0
Fe
p
F1 = γ2
[0, 1] × [0, 1] γ2 (1)
X
x0
Ft F (s, t) γ1
t
γ2
γ1 (1) = γ2 (1)
x0 γ1 (1)
F0 = γ1
We claim that it’s sufficient to show that there exists a continuous Fe : I × I → X such that
p ◦ Fe = F , and Fe({0} × I) = x0 . It’s because
p ◦ Fe0 = F0 = γ1 , p ◦ Fe1 = F1 = γ2
Fe : I × I → X
(s, t) 7→ Fet (s),
..
.
Ux0 α3
Ux0 α2
Ux0 α1
Fe x
e0
p
Since F is continuous, we see that there exists an open neighborhood Ux0 of x0 such that
p−1 (Ux0 ) = α Ux0 α , where
`
∼
=
p|Ux : Ux0 α → Ux0 .
0α
Since F −1 (Ux0 ) is an open set contain {0} × I, there exists a ϵ0 > 0 such that [0, ϵ0 ] × I ⊂
F −1 (Ux0 ),b which implies
F ([0, ϵ0 ] × I) ⊂ Ux0 .
−1
Note that x0 ∈ Ux0 and p(e x0 ) = x0 , we may assume x e0 ∈ Ux0 α1 . Consider p|Ux α ◦
0 1
F |[0,ϵ0 ]×I , which is a lift of F |[0,ϵ0 ]×I . We claim that
−1
p|Ux ◦ F |[0,ϵ0 ]×I = Fe .
0 α1 [0,ϵ0 ]×I
is a lift starting at x
e0 ; also, for every t ∈ I,
s 7→ Fe (s, t)
[0,ϵ0 ]×I
is a lift of Ft starting at x
e0 . From the uniqueness of the lift of paths, we see that they’re equal.
Note that this implies Fe is now continuous at [0, ϵ0 ] × I, since F is continuous and p|Ux α is
0 1
a homeomorphism, hence continuous, then from
−1
Fe = p|Ux α ◦ F |[0,ϵ0 ]×I ,
[0,ϵ0 ]×I 0 1
| {z } | {z }
continuous continuous
(b) We now prove that Fe : I × I → X e is continuous. Assume there exists (s0 , t0 ) ∈ I × I such
that Fe is discontinuous at (s0 , t0 ). Then consider
n o
0 < ϵ0 ≤ inf s | Fe is discontinuous at s, t0 =: s1 ,
| {z }
∋s0 ⇏=∅
X
e
x
e1
Fe
p
[0, 1] × [0, 1]
γ2 (1)
X
F (s, t)
(s1 , t0 ) x1
Ux1
x0 γ1 (1)
0 ϵ0
Let x1 := F (s1 , t` e1 := Fe(s1 , t0 ), then there exists an open neighborhood Ux1 in X such
0 ), x
that x1 ∈ Ux1 = α Ux1 α , where
∼
=
p|Ux : Ux1 α → Ux1 .
1α
Ux1 α3
Ux1 α2
Ux1 α1
Fe
p
[0, 1] × [0, 1]
X
t0 + δ1 x1
F (s, t)
(s1 , t0 )
t0 − δ1
Ux1
s1 − ϵ1 s1 + ϵ1
We may assume x e1 ∈ Ux1 α1 . Then, we see that Fet0 is a lift of Ft0 , which means Fet0 is
continuous, hence there exists an s2 such that s1 − ϵ1 < s2 < s1 such that
Fe(s2 , t0 ) ∈ Ux1 α1 .
Fe(s2 , t0 ) =: x
e2
x
e1 Ux1 α1
Fe
p|Ux
1 α1
[0, 1] × [0, 1]
X
t0 + δ1
(s2 , t0 ) F (s, t)
t0
(s1 , t0 ) x1
t0 − δ1
Ux1
s1 − ϵ1 s1 + ϵ1
We see that Fe is continuous at (s2 , t0 ), hence there exists a δ2 > 0 such that
Note that here we can also consider a closed interval, which matches what we’re going to do.
Namely, we’re going to construct a closed box B. But this is just a technical detail.
Fe(L)
x
e2 x
e1 Ux1 α1
Fe
B p|Ux
1 α1
s1 − ϵ1 s1 + ϵ1
Now, observe that Fe(B) ⊂ Ux1 α1 . To see this, consider a fixed t ∈ (t0 + δ2 , t0 − δ2 ), then the
map Fe is
[s1 − ϵ1 , s1 + ϵ1 ] → X,
e s 7→ Fe(s, t) = Fet (s).
Specifically, a
Fet ([s1 − ϵ1 , s1 + ϵ1 ]) ⊂ p−1 (Ux1 ) = Ux1 α ,
α
with the fact that Fet ([s1 − ϵ1 , s1 + ϵ1 ]) is connected, and Fet (s2 ) ∈ Ux1 α1 with Fet is a lift of Ft ,
hence continuous, so
Fet ([s1 − ϵ1 , s1 + ϵ1 ]) ⊂ Ux1 α1 .
p|Ux ◦ Fe = F |B ,
1 α1 B
and −1
p|Ux ◦ F |B : B → Ux1 α1 ,
1 α1
so −1
p|Ux ◦ p|Ux ◦ F |B = F |B
1 α1 1 α1
−1
Fe = p|Ux α ◦ F |B ,
B 1 1
| {z } | {z }
continuous continuous
■
a https://en.wikipedia.org/wiki/Lebesgue%27s_number_lemma
b Notice that we’re working on product topology here.
c There is a tricky situation, namely while s1 = 1. But this can be considered also.
X
e1 X
e2 X
e3
a b b b
b a
b
a b
b b a a a
a
p1 p2 p3
x0
a X = S1 ∨ S1
b
Note that in each cover (those three on the top), the black dot is the preimage of {x0 }, namely
p−1
i ({x0 }).
(a) p∗ : π1 (X,
e xe0 ) → π1 (X, x0 ) is injective.
{[γ] | lift γ
e starting at x
e0 is a loop.} .
Proof. We prove this one by one.
(a) Suppose γ
e ∈ π1 (X,
e xe0 ) is in ker(p∗ ). Then
γ ]) = [p ◦ γ
[γ] = p∗ ([e e] .
Let γt be a nullhomotopy from γ to the constant loop cx0 rel{0, 1}. We can then lift γt to γ
et
where γ e. Now, we claim that
e0 = γ
• γ
e is a homotopy rel{0, 1}.
• γ
e1 is the constant loop cxe0 .
X
e X
e
γ γ
et
e p p
I γ X I ×I γt X
We see that the above diagrams prove the first claim, since we know that the left and right
edge of I × I maps to x0 under γt , and cxe0 lifts this, so by uniqueness t 7→ γ
et (0) and t 7→ γ
et (1)
must be constant paths at xe0 as desired.
Then the lift γ
et is a homotopy of paths to the constant loop, so [e
γ ] = 1.
Example. Given
max tree
X
e b
b x
e0 a
x0
X a b
Then
p∗ π1 = ⟨b, a2 , aba⟩ ⊆ π1 (X) = ⟨a, b | ⟩.
Proposition 4.1.3 (Lifting criterion). Let p : (Ye , ye0 ) → (Y, y0 ) be a covering map. Given
• f : (X, x0 ) → (Y, y0 );
• X is path-connected, locally path-connected,
then a lift
fe: (X, x0 ) → (Ye , y0 )
exists if and only if
f∗ (π1 (X, x0 )) ⊆ p∗ (π1 (Ye , ye0 )).
In diagram, we have
(X, x0 ) f
(Y, y0 ) π1 (X, x0 ) f∗
π1 (Y, y0 )
Proof. If we can show that there is a lift fe: RP 2 → R of f , then we’re done since we can apply the
straight line nullhomotopy on R, i.e.,
R
fe
p
RP 2 f
S1
f∗ (π1 (RP 2 )) = 0
f∗ = p∗ ◦ fe∗
fe(x) = ffγ(1).
γ ′ (1) = ffγ(1).
fg
Ye
fe(x)
◦γ
f]
ye0
fe
y0
x0
γ
f ◦γ
x
γ ′ f
f ◦ γ′
f (x)
X Y
V
α
x′
γ x f fα
x0 f (V )
fγ
fe(x′ ) = f γ · f α(1) ∈ U
e,
which implies
fe(V ) ⊆ U
e.
Ye
fe1
p
fe2
X f
Y
fe1 (x) U
e1
fe2 (x) U
e2
fe1
p
fe2
U
f (x)
x
Y
We see that U
e1 and U
e2 are slices of p−1 (U ), where U is an evenly covered neighborhood of f (x).
(a) If fe1 (x) ̸= fe2 (x). Then U e2 are disjoint. Since fe1 , fe2 are continuous, there exists a neigh-
e1 , U
borhood N of x with
fe1 (N ) ⊆ U
e1 , fe2 (N ) ⊆ U e2 ,
hence x ∈ Int(S).
e1 → X,
p1 : X e2 → X,
p2 : X
f
X
e1 X
e2
p1 p2
X
Note. Note that we’ve suppressed the data of p in the notation, but this data is essential to what
a deck transformation is, when this is unclear we write G(X,
e p).
• G(S 1 , pn ) ∼
= Z / nZ
∼
= S1
S1
Exercise (Deck Transformation is determined by the image of one point). Given X, X e are path-
connected, locally path-connected, deck map is determined by the image of any one point.
Answer.
X
e
f
p
X
e
p X
⊛
Corollary 4.2.1. If a deck transformation has a fixed point, it is the identity transformation.
e0 7→ x
x e1 .
Example (Regular and non-regular cover of S 1 ∨ S 1 ). Given the following covers of S 1 ∨ S 1 , determine
which cover is regular.
x
e1
x
e2
x
e0
Proof. The left one is regular, while the right one is not since there is no automorphism from x
e0 to
e1 or x
x e2 . ⊛
Definition 4.2.5 (Normal subgroup). A subgroup N of G is called a normal subgroup if it’s invariant
under conjugation, and we denote this relation as N ◁ G.
N (H) := {g ∈ G | gH = Hg} .
(b) H ≤ N (H).
(c) H is normal in N (H).
(d) If H ≤ G is normal, N (H) = G.
(e) N (H) is the largest subgroup (under containment) of G containing H as normal subgroup.
where G(X)
e are deck maps, and N (H) is the normalizer of H in π1 (X, x0 ).
p∗ (π1 (X,
e xe0 )) = [γ]p∗ (π1 (X,
e xe1 ))[γ −1 ].
γ
e
x
e0 x
e1
γ
γ =p◦γ
x0
e
p∗ (π1 (X,
e xe1 )) = [γ] · p∗ (π1 (X,
e xe0 )) · [γ]−1
Φ : N (H) → G(X)[γ],
e · 7→ τ
(a) Φ is surjective.
(b) ker(Φ) = H.
(c) Φ is a group homomorphism.
If we can prove all the above, then the result follows directly from the first isomorphism theorem.
Φ Φ
(c) Suppose we have loops [γ1 ] 7→ τ1 and [γ2 ] 7→ τ2 . We claim that γ1 · γ2 lifts to γ
e1 · τ (e
γ2 ).
· γ′
γ]
γ
e x
x
e0 x
e2 e3
p
lifts of γ ′ starting at x
e2
γ γ′
x0
x
e0 x
e2
τ γ′)
τ (e
e′
γ
e ∼
Corollary 4.2.2. If p is a normal covering, then G(X) = π1 (X, x0 ) / H.
Corollary 4.2.3. If X e ∼
e is the universal cover, then G(X) = π1 (X, x0 ).
Exercise. Let Σg be the genus g surface. Prove that Σg has a normal n-sheeted path-connected
cover for every n.
Homology
X X
γ1
γ
x0 x0
γ2
We see that it’s extremely hard to compute higher fundamental group. Hence, instead, we will study
the higher dimensional structure of X via homology.
• Cons.
– The definition is more opaque at first encounter.
• Pros.
74
Lecture 19: Simplex and Homology
– No basepoints
– Can compute using CW structure.
– Good properties. For example, Hn = 0 if n > dim X
△-Complex structure
CW Complex structure
• 2-simplex. Triangle.
• 3-simplex. Tetrahedron.
• n-simplex. The convex hull of (n + 1)-points position in Rn .
v2 v3
v2
v0 v0 v1 v0 v1 v0 v1
CHAPTER 5. HOMOLOGY 75
Lecture 20: Simplicial Complex
• The top of which is the 3-disk and cells and the orientation.
• We can view simplices as both combinatorial and topological objects.
An alternative definition can be made.
Definition 5.2.2 (Standard simplex). We say that an n-dimensional standard simplex, denoted by
∆n is ( )
X
n n+1
∆ = (t0 , . . . , tn ) ∈ R | ti ≥ 0, ti = 1 .
i
v1 v2
v0 v0 v1
v0
∆0 ∆1 ∆2
Remark. The standard simplices will implicitly come with a choice of ordering of the vertices as
∆n = [v0 , v1 , . . . , vn ]
such that the convex hull of these points is taken with this ordering.
v2
[v0 , v2 ] [v1 , v2 ]
[v0 , v1 , v2 ]
v0 [v0 , v1 ] v1
Definition 5.2.5 (Boundary). The boundary ∂σ of a simplex σ is the union of its faces.
CHAPTER 5. HOMOLOGY 76
Lecture 20: Simplicial Complex
σα : ∆ n → X
such that
(a) σα |∆
˚ n injective, each point of X is in the image of exactly one such map.
σβ : ∆n−1 → X.
Note. We see that the second condition of Definition 5.2.7 implies that attaching maps injective on
interior of faces.
Example (Difference between simplicial and ∆-complex structure of S 1 ). With Definition 5.2.7 and
Definition 5.2.8, we see the followings.
Example (Difference between simplicial and ∆-complex structure of a torus). The torus with the
following edges, a, b, c and the gluing in triangles A and B can be seen as follows.
CHAPTER 5. HOMOLOGY 77
Lecture 21: Simplicial Homology
A
a c a
Proof. For this ∆-complex, notice that we’ve glued down a triangle whose vertices are all identified.
This is not allowed in a simplicial complex/triangulation.
σ2
σ1 σ3
Then, the n-th homology group will be a subquotient of Cn (X), where the heuristic/imprecise idea
is
• Take subgroup of Cn of cycles. These are sums of simplices satisfying a combinatorial condition
on the boundary gluing maps to ensure that they close up, i.e., they have no boundary.
CHAPTER 5. HOMOLOGY 78
Lecture 21: Simplicial Homology
σ1
σ
σ
σ2
σ3
σ1 + σ2 + σ3 + σ4 cycles σ1 + σ2 + σ3 not a cycles
• To take the quotient, we consider two cycles to be equivalent if their difference is a boundary.
For example, in the case of torus, a is homologous to b since a − b is the boundary of the shaded
subsurface S on of the torus below.
b S
In fact, a and b are homotopic (which will imply they’re homologous essentially), but two loops do
not need to be homotopic to be homologous. For example, in the figure below, a + b is homologous
to c, since a + b − c is the boundary of S (a + b1 and c are not homotopic).
S a
b
c
Definition 5.2.9 (Simplicial chain group). We define the simplicial chain group Cn (X) of order n to
be the free Abelian group on the n-simplices of X such that
( )
X
Cn (X) := finite sums n
mα σα | mα ∈ Z, σα : ∆ → X .
Definition 5.2.10 (Cycle). Given any chain group Cn (X), a cycle of Cn (X) is those chains
P
mα σα
with no boundaries.
CHAPTER 5. HOMOLOGY 79
Lecture 21: Simplicial Homology
Definition 5.2.11 (Boundary homomorphism). A map ∂n : Cn (X) → Cn−1 (X) is called a boundary
homomorphism such that
Remark. We see that the definition of boundary homomorphism indeed coincides with the definition
of boundary when considering either ∆-complex or simplicial complex structure.
Example. We give some lower dimensions examples of Definition 5.2.11 to motivate the general
definition.
• For n = 1, ∂1 : C1 (X) → C0 (X) such that
Lemma 5.2.1. Let Cn being the simplicial chain group and ∂n being the boundary homomorphism,
for any n ≥ 2, we have
∂n ∂n−1
Cn (X) Cn−1 (X) Cn−2 (X)
∂n−1 ◦∂n =0
Proof. Since all Ci are free Abelian group, hence we only need to consider ∂n−1 ◦ ∂n (σ) = 0 for a
generator σ. Given a generator σ, the result follows from directly applying the definition and with
some calculation. ■
Definition 5.2.12 (Chain complex). A chain complex (C∗ , d∗ ) is a collection of maps dn between
groups Cn such that
dn+1 dn dn−1
... Cn+1 Cn Cn−1 ...
dn−1 ◦ dn = 0.
We see that we can certainly think Cn and dn as the simplicial chain group and the boundary
homomorphism as defined before since we already showed this combination satisfies Definition 5.2.12 in
Lemma 5.2.1. But we note that the definition of Cn and dn can be further abstract as what we have
defined.
CHAPTER 5. HOMOLOGY 80
Lecture 22: Calculation of Homology
Note. As just discussed, we can put different chain group structure on Cn . We’ll see what this
means later.a But for now, we think Cn be equipped with the definition we gave for simplicial chain
group.
a Spoiler: It just means we can give different definition about the map σ.
Definition 5.2.15 (Exact). We say that the sequence is exact at Cn provided that ker(dn ) =
Im(dn+1 ). A chain complex is exact if it is exact at each point.
Definition 5.2.16 (Homology group). The nth homology group of a chain complex (C∗ , d∗ ), denoted
as Hn or Hn (C∗ ), is the quotient
.
Hn := ker(dn ) Im(d ). n+1
Remark. The homology group measures how far the chain complex is from being exact at Cn .
With what we have just defined, it’s natural to define homology groups of space X with a ∆-complex
structure.
Definition 5.2.17 (Homology class). We say ker(∂n ) is the subgroup of cycles is Cn (X), and Im(∂n+1 )
is the subgroup of boundaries in Cn (X). We then set
{cycles}
. .
Hn (X) := ker(∂n ) Im(∂ ) = {boundaries}.
n+1
Remark. Definition 5.2.17 is saying that we should call the element of a homology group whose
chain group is some kinds of geometric subjects. In this case, we’re just considering ∆-complex
structure, but the definition of homology class is in fact more general.
Definition 5.2.18 (Simplicial homology group). By considering the chain complex being the simplicial
chain groups and boundary homomorphisms, we have so-called simplicial homology groups induced
by Definition 5.2.16.
CHAPTER 5. HOMOLOGY 81
Lecture 22: Calculation of Homology
Example (Homology group of RP 2 ). Calculate the homology group by Definition 5.2.16 with the
chain complex being the simplicial chain complex.
Proof. Let X = RP 2 .
w b v v1 v2 u2
A A
a c a
B
B
v b w v0 u0 u1
• C2 = Z⟨A, B⟩ = ZA ⊕ ZB
The chain complex is then
∂3 ∂2 ∂1 ∂0
0 C2 C1 C0 0
C2 : Im ∂3 = 0, ker ∂2 = 0,
C1 : Im ∂2 = ⟨2c, b − c + a⟩, ker ∂1 = ⟨b + a, c⟩,
C0 : Im ∂1 = ⟨v − w⟩, ker ∂0 = ⟨v, w⟩.
Hence,
.
= Z⟨v, w⟩ Z⟨v − w⟩ ∼
H0 ∼ =Z
= Z⟨b + a − c, c⟩ Z⟨2c, b + a − c⟩ ∼
. . .
= Z⟨b + a, c⟩ Z⟨2c, b + a − c⟩ ∼
H1 ∼ = Z 2Z
H2 = 0
Remark. Given a basis for a free Abelian group ⟨b1 , . . . , bn ⟩ we can replace bi with
bi ± m 1 b1 ± · · · ± m
di bi ± · · · ± mn bn .
Remark. Care is needed when doing change of bases over Z. For example, if b1 , b2 is a basis for
A ⊆ Zn , then b1 − b2 , b1 + b2 is not a basis, it is an index-2 subgroup. The key to this is that
CHAPTER 5. HOMOLOGY 82
Lecture 23: Singular Homology
1 1
has determinant −2 (not unit in Z).
1 −1
We can transform a basis for a free group into a different basis by applying a matrix of determinant
±1. If we apply a matrix of determinant D we will obtain generators for a subgroup of index |D|.
1 0 ··· 0 ±m1 0 ··· 0
0 1 · · · 0 ±m2 0 ··· 0
.. .. .. . . . .. .
. . . .. .. .. . ..
0 0 · · · 1 ±m i−1 0 · · · 0
0 0 · · · 0 1 0 · · · 0
0 0 · · · 0 ±m i+1 1 · · · 0
. . . . . . . .
.. .. .. .. .. .. . . ..
σ : ∆n → X.
Definition 5.3.2 (Singular chain complex). The chain complex defined with singular chain group and
singular boundary map defined as follows is called singular chain complex.
Definition 5.3.3 (Singular chain group). Let Cn (X) be the free group on singular n-simplices
in X, which we call it the singular n-chain group.
Definition 5.3.4 (Singular boundary map). With C∗ being the singular chain group, we defined
so-called singular boundary map ∂n as
CHAPTER 5. HOMOLOGY 83
Lecture 23: Singular Homology
Definition 5.3.5 (Singular homology group). The singular homology groups are the homology groups
of this singular chain complex given as
.
Hn (X) = ker ∂n Im ∂ .
n+1
Remark. We now see that from the definition of homology group, we can put different structure on
which. But the idea is the same, namely we are taking Hn (X) being
.
Hn (X) := ker ∂n Im ∂ ,
n+1
where the difference is what structure we put on X which induces different chain complex (C∗ (X), ∂∗ ).
In this case, we have singular homology group since we are considering singular chain complex, while
we can also have simplicial homology group.
Since the generating sets for Cn (X) when considering singular chain complex are almost always hugely
uncountable from its definition, it’s almost impossible to compute with these. However, it does give us
a definition that does not depend on any other structure than the topology of X, making it useful for
developing theory.
σ1 σ2 X
where we’ve glued [v1 , v2 ] of σ1 to [v0 , v2 ] of σ2 if σ1 |[v1 ,v2 ] and σ[v0 ,v2 ] are the same singular
n-chain with opposite signs.
Hn (X) ∼
M
= Hn (Xα ).
α
CHAPTER 5. HOMOLOGY 84
Lecture 24: Chain Homotopy
Definition 5.4.1 (Induced map on singular chains). For a given continuous map f : X → Y , we can
consider the map f# induced by singular chains as
f# : Cn (X) → Cn (Y )
[σ : ∆n → X] 7→ [f ◦ σ : ∆n → Y ].
Note. Note that we’re considering singular chain groups specifically in this case.
Remark. We see that the functoriality doesn’t depend on any kind of ∆-complex structure.
Definition 5.4.2 (Chain map). Given two chain complexes (C∗ , ∂∗ ) and (D∗ , δ∗ ), a chain map be-
tween them is a collection of group homomorphisms fn : Cn → Dn such that the following diagram
commutes.
∂n+2 ∂n+1 ∂n ∂n−1
... Cn+1 Cn Cn−1 ...
fn+1 fn fn−1
δn+2 δn+1 δn δn−1
... Dn+1 Dn Dn−1 ...
i.e. we have that δn ◦ fn = fn−1 ◦ ∂n .
Exercise. We have that f# ∂ = ∂f# . In other words, f# is a chain map. Thus, by the homework f#
induces a group homomorphism on the homology groups. We write this as f∗ : Hn (X) → Hn (Y )
for all n.
Theorem 5.4.1 (Homology group defines a functor). The n-th homology group Hn : X 7→ Hn (X)
gives a functor from Top to Ab.
Theorem 5.4.2 (Functoriality is homotopy invariant). If f, g : X → Y are homotopic, then they will
induce the same map on homology
f∗ = g∗ : Hn (X) → Hn (Y ).
The proof of Theorem 5.4.2 can be found here.
Exercise. Theorem 5.4.1 and Theorem 5.4.2 imply that Hn is a homotopy invariant.
I’m not sure whether the above discussion holds only for singular homology group or can be ex-
tended to general homology group. I link them to general homology group anyway.
Definition 5.4.3 (Chain homotopy). Given chain complexes (A∗ , ∂∗A ) and (B∗ , ∂∗B ) and chain maps
f# , g# : A∗ → B∗ . A chain homotopy from f# to g# is a sequence of group homomorphisms
ψn : An → Bn+1 such that
B
fn − gn = ∂n+1 ◦ ψn + ψn−1 ◦ ∂nA .
CHAPTER 5. HOMOLOGY 85
Lecture 24: Chain Homotopy
B B B B
∂n+2 ∂n+1 ∂n ∂n−1
... Bn+1 Bn Bn−1 ...
This diagram does not commute, however, the red map is the sum of the blue maps composed up,
so it shows everything that is going on.
Theorem 5.4.3. If there is a chain homotopy ψ from f# to g# , then the induced maps f∗ , g∗ on
homology are equal.
Proof. Let σ ∈ An be an n-cycle, i.e. ∂nA σ = 0. Then we compute that:
B
(fn − gn )(σ) = ∂n+1 (ψn (σ)) + ψn−1 (∂nA (σ)) = ∂n+1
B B
(ψn (σ)) ∈ Im(∂n+1 ).
This tells us that (fn − gn )(σ) is a boundary, and so (fn − gn )(σ) = 0 when considered as an
element of the homology group (with degree n). Thus, fn (σ) = gn (σ) in the homology group, and
so f, g induce the same map as desired. ■
We now sketch the proof of Theorem 5.4.2 given in Hatcher [HPM02]. From this point in the course
many of the theorems require much more algebraic work than we are interested in. We instead want to
learn how to use the computational tools.
Proof sketch of Theorem 5.4.2. Suppose we have some homotopy F : I × X → Y from f to g.
The most difficulty in this proof is the combinatorial difficulty involved in the fact that the product
of a simplex in X and I is not a simplex.
We now consider
∆′
Pn : Cn (X) → Cn+1 (Y )
alternating sums of restrictions
σ×id F
[σ : ∆n → X] 7→ ∆n × I −−−→ X × I − →Y
to each simplex in our subdivision
Y
∂n+1 Pn = g# − f# − Pn−1 ∂nX .
CHAPTER 5. HOMOLOGY 86
Lecture 25: Relative Homology
g#
∆n × {1}
P∂
∆n × {0}
f#
■
a We want to do this since the product between two simplices is not a simplex, as we just note.
Definition 5.5.2 (Good pair). Let X be a space, and A ⊆ X. Then (X, A) is a good pair if A is
closed and nonempty, and also it is a deformation retract of a neighborhood in X.
Theorem 5.5.1 (Long exact sequence of a good pair). If (X, A) is a good pair, then there exists a long
exact sequence (exact at every n) on reduced homology groups given by the following commutative
CHAPTER 5. HOMOLOGY 87
Lecture 25: Relative Homology
diagram.
i∗ j∗
... H
e n (A) H
e n (X) H
e n (X / A)
δ
i∗ j∗
H
e n−1 (A) H
e n−1 (X) H
e n−1 (X / A)
δ
i∗ j∗
... H
e 0 (X) H
e 0 (X / A) 0
Remark. The fact that this sequence is exact often means that if we know the homology groups of
two of the spaces we can compute the homology of the remaining space.
Before we see the proof of Theorem 5.5.1, we see one application.
• H
e n (Dn ) = 0 for all n since Dn is contractible.
• ∂Dn ∼
= S n−1 .
We then proceed by induction on n. To start with, we need to verify the following.
... e n (∂Dn )
H i∗ e n (Dn )
H j∗ e n (S n )
H
δ
e n−1 (∂Dn )
H i∗ e n−1 (Dn )
H j∗ e n−1 (S n )
H
... i∗ e 0 (Dn )
H j∗ e 0 (S n )
H 0
By induction, we have H e n−1 (S n−1 ) = Z, hence we can fill in some of these groups
e n−1 (∂Dn ) = H
as follows.
... 0 i∗ 0 j∗ He n (S n )
Z i∗ 0 j∗ e n−1 (S n )
H
... i∗ 0 j∗ e 0 (S n )
H 0
CHAPTER 5. HOMOLOGY 88
Lecture 25: Relative Homology
e n (S n ) δ
0 H Z 0
e i (Dn )
H e i (S n )
H e i−1 (∂Dn )
H
0 e i (S n )
H 0
Theorem 5.5.2 (Brouwer’s fixed point theorem). ∂Dn is not a retract of Dn . Hence, every continuous
map f : Dn → Dn has a fixed point.
Proof. If r : Dn → ∂Dn were a retraction, then by definition this would give us
i r
∂Dn Dn ∂Dn
id∂Dn
So then:
i∗ r∗
Z 0 Z
id
Lemma 5.5.1 (The short five lemma). Suppose we have a commutative diagram
ψ φ
0 A B C 0
α β γ
′ ψ′ φ′
0 A B C′ 0
CHAPTER 5. HOMOLOGY 89
Lecture 26: Continue on Relative Homology
And thus by injectivity of γ we know φ(b) = 0. By exactness, b ∈ Im ψ. We then may write for
some a ∈ A such that the following diagram commutes.
ψ φ
0 a b 0 0
α β γ
ψ′ φ′
0 α(a) 0 0 0
Definition 5.5.3 (Relative chain group). Let X be a space and let A ⊆ X be a subspace. Then we
define the relative chain group as
.
Cn (X, A) = Cn (X) C (A),
n
which is a quotient of Abelian groups of the singular chain groups between X and A.
Definition 5.5.4 (Relative chain complex). The relative chain complex (C∗ , ∂∗ ) consists of relative
chain group and the usual differential associated with the singular chain groups which induces our
relative chain group.
Remark. We can indeed adapt Definition 5.5.4 by either singular chain complex structure or sim-
plicial chain complex structure.
It’s not entirely clear that whether Definition 5.5.4 is well-defined, hence we have the following
exercise.
Exercise. Since ∂n∗ (Cn (A)) ⊆ Cn−1 (A), hence there exists a well-defined map
. .
∂n : Cn (X) C (A) → Cn−1 (X) C (A).
n n−1
We can verify that ∂ 2 = 0. Then, since ∂ 2 = 0 we can conclude that these groups will in fact form
a chain complex (C∗ (X, A), ∂).
Definition 5.5.5 (Relative homology group). The homology groups of the relative chain complex
(C∗ (X, A), ∂) are denoted by Hn (X, A), and they are called relative homology groups.
We see that there are something interesting going on in relative chain group. Indeed, we can further
classify the cycles in which as follows.
CHAPTER 5. HOMOLOGY 90
Lecture 26: Continue on Relative Homology
Definition 5.5.6 (Relative cycle). Elements in ker ∂n are called relative n-cycles. These are
elements α ∈ Cn (X) such that ∂n α ∈ Cn−1 (A).
relative cycle
relative 2-cycle
X X
Definition 5.5.7 (Relative boundary). Elements α in Im ∂n+1 are called relative n-boundaries.
This means that α = ∂β + γ where β ∈ Cn (X) and γ ∈ Cn−1 (A).
X
β
A
Figure 5.2: We see that we have α + γ = ∂β, where α is a relative boundary, and γ ∈ Cn−1 (A).
Theorem 5.5.3 (Long exact sequence of a pair). Let A ⊆ X be spaces, then there exists a long exact
sequence
... H
e n (A) i∗
He n (X) q H e n (X, A)
∂
i∗ q
H
e n−1 (A) ... H
e 0 (X, A) 0
Remark. A short exact sequence of chain complexes gives rise to a long exact sequence of homology
groups. Namely, given a short exact sequence of chain complexes (A∗ , ∂ A ), (B∗ , ∂ B ), (C∗ , ∂ C ) such
that
ι q
0 A∗ B∗ C∗ 0
where ι, q are chain maps such that
ιn qn
0 An Bn Cn 0
CHAPTER 5. HOMOLOGY 91
Lecture 27: Excision
is exact for all n. Then Theorem 5.5.1 will follow from a short exact sequence
0 C
e∗ (A) C
e∗ (X) C
e∗ (X, A) 0
where C
e∗ denotes the augmented chain complex (the one with Z after it, as in Definition 5.5.1).
Theorem 5.5.4 (Excision). Suppose we have subspace Z ⊆ A ⊆ X such that Z ⊆ Int(A). Then the
inclusion
(X − Z, A − Z) ,→ (X, A)
induces isomorphisms
∼
=
Hn (X − Z, A − Z) −
→ Hn (X, A).
Proof Sketch. We first see an equivalent formulation of Theorem 5.5.4.
(B, A ∩ B) ,→ (X, A)
induces an isomorphism
∼
=
Hn (B, A ∩ B) −
→ Hn (X, A)
B := X \ Z, Z = X \ B,
then we see that A ∩ B = A − Z and the condition requires from Theorem 5.5.4, Z ⊆ Int(A)
is then equivalent to
X = Int(A) ∪ Int(B)
since X \ Int(B) = Z.
A A
X Z X Z
Hn (X, A) Hn (X / Z, A / Z)
⊛
We now sketch the proof of the above equivalent form of Theorem 5.5.4, which is notorious for
CHAPTER 5. HOMOLOGY 92
Lecture 28: Singular Homology v.s. Simplicial Homology
being hairy.
• Given a relative cycle x in (X, A), subdivide the simplices to make x a linear combination of
chains on smaller simplices, each contained in Int(A) or X \ Z.
Z A X Z A X
Theorem 5.5.5. For good pairs (X, A), the quotient map q : (X, A) → (X / A, A / A) induces iso-
morphisms
∼
q∗ : Hn (X, A) = Hn (X / A, A / A) = ∼H e n (X / A)
for all n.
Proof Sketch. Let A ⊆ V ⊆ X where V is a neighborhood of A that deformation retracts onto A.
Using excision, we obtain a commutative diagram
∼
= ∼
=
Hn (X, A) Hn (X, V ) Hn (X − A, V − A)
q∗ q∗ ∼
= q∗
∼
= ∼
=
Hn (X / A, A / A) Hn (X / A, V / A) Hn (X / A − A / A, V / A − A / A)
• ∼
= is an isomorphism by excision.
• ∼
= is an isomorphism by direct calculation (since q is a homeomorphism on the complement
of A).
• ∼
= on Homework, since V deformation retracts to A.
Remark. The last equality is from the above exercise with A / A = {∗}.
CHAPTER 5. HOMOLOGY 93
Lecture 29: Proof of Theorem 5.5.6
Example. With pairs like (Rn+1 , S n ), you can just assert that this is a good pair.
Proof. If we want to be rigorous, we need to show that S n is a smooth submanifold of Rn+1 , but
we simply assert this in our course. ⊛
Theorem 5.5.6 (Singular homology agrees with simplicial homology). Let X be a ∆-complex. We
use ∆n (X) to represent the simplicial chain groups on X, and Cn (X) to denote the singular chain
groups. Likewise, we denote .
∆n (X, A) = ∆n (X) ∆ (A)
n
and .
Cn (X, A) = Cn (X) C (A).
n
[σ : ∆n → X] 7→ [σ : ∆n → X]
Hn∆ (X, A) ∼
= Hn (X, A).
Hn∆ (X) ∼
= Hn (X).
The proof of Theorem 5.5.6 uses the following lemma.
Lemma 5.5.2 (The five lemma). If we have a commutative diagram with exact rows as following,
i j k ℓ
A B C D E
α β γ δ ϵ
A′ B′ C′ D′ E′
i′ j′ k′ ℓ′
CHAPTER 5. HOMOLOGY 94
Lecture 29: Proof of Theorem 5.5.6
to reduce the proof to the finite skeleton X k of X, namely we can use induction.
From the long exact sequence of a pair we get
∆
Hn+1 (X k , X k−1 ) Hn∆ (X k−1 ) Hn∆ (X k ) Hn∆ (X k , X k−1 ) ∆
Hn−1 (X k−1 )
α β γ δ ϵ
We claim that Hn (X k , X k−1 ) are also free Abelian on the singular k-simplices defined by the
characteristic maps ∆k → X k when n = k, and 0 otherwise. Consider the map
a
Φ: (∆kα , ∂∆kα ) → (X k , X k−1 )
α
Corollary 5.5.1. If X has a ∆-complex structure (or is homotopy equivalent to one), then we have
the followings.
(a) If the dimension is ≤ d, then Hn (X) = 0 for all n > d.
(b) If X has no cells of dimension p, then Hp (X) = 0.
(c) If X has no cells of dimension p, then Hp−1 (X) is free Abelian.
Corollary 5.5.2. Given a singular homology class on X, without loss of generality we can choose
a ∆-complex structure on X, and we then we can assume the class is represented by a simplicial
n-cycle.
Remark. Recall the definition of homology class, as we noted before, this means we can view
singular chain complex as some kind of geometric subjects. The construction can be found in
Hatcher [HPM02].
5.6 Degree
CHAPTER 5. HOMOLOGY 95
Lecture 29: Proof of Theorem 5.5.6
f∗ : Z ∼
= Hn (S n ) → Hn (S n ) ∼
=Z
(b) If f : S n → S n , n ≥ 0 is not surjective, then deg(f ) = 0. To see this, we know that f∗ factors
as
Hn (S n ) Hn (S n − {∗}) = 0 Hn (S n )
f∗
And since the middle group is zero because S n \ {∗} is contractible, f∗ = 0, so is its degree.
(c) If f ≃ g, then f∗ = g∗ , so deg(f ) = deg(g).
so deg f = ±1.
(e) If f is a reflection fixing the equator, and swapping the 2-cells, then deg f = −1.
Hn (S n ) = ⟨∆1 − ∆2 ⟩.
∆1
f , reflection
∆2
Answer. By doing so, we see that the reflection f interchanges ∆1 and ∆2 , hence the
generator is now its negative. ⊛
CHAPTER 5. HOMOLOGY 96
Lecture 30: Degree
Theorem 5.6.1 (Hairy ball theorem). The sphere S n admits a nonvanishing continuous tangent
vector field if and only if n is odd.
Proof. Recall that a tangent vector field to the unit sphere S n ⊆ Rn+1 is a continuous map
v : S n → Rn+1
such that v(x) is tangent to S n at x, i.e., v(x) is perpendicular to the vector x for each x. Let v(x)
be a nonvanishing tangent vector field on the sphere S n , then we define
v(x)
ft (x) := cos(πt) + sin(πt) ,
∥v(x)∥
which is a homotopy from the identity map idS n : S n → S n to the antipodal map − idS n : S n → S n .
This simply follows from varying t from 0 to 1, where we have
v(x)
f0 (x) = cos(0)x + sin(0) = x ⇒ f0 = idS n ,
∥v(x)∥
while
v(x)
f1 (x) = cos(π)x + sin(π) = −x ⇒ f1 = − idS n .
∥v(x)∥
The last thing needs to be verified is that ft (x) is continuous, but this is trivial.
From the property of degree, we know that it’s a homotopy invariant, hence
which implies
(−1)n+1 = 1,
so n must be odd.
Conversely, if n is odd, say n = 2k − 1, we can define
Then v(x) is orthogonal to x, so v is a tangent vector field on S n , and |v(x)| = 1 for all x ∈ S n . ■
d : G → {±1}
g 7→ deg(τg )
CHAPTER 5. HOMOLOGY 97
Lecture 30: Degree
since τg is invertible (simply take τg−1 ) for each g ∈ G. Also, we see that d(g1 g2 ) = d(g1 ) · d(g2 ),
hence it is a homomorphism.
We want to show that the kernel is trivial, since then by the first isomorphism theorem G ∼ = Im,
and the image is either trivial or Z / 2Z. Suppose that g is a nontrivial element of G, then since G
acts freely we know that τg has no fixed points. With this in mind we have
Corollary 5.6.1. S 2n has only the trivial cover S 2n → S 2n or degree 2 cover (for example, S 2n →
RP 2n ).
Proof. This follows since any covering space action acts freely, then we simply apply Theorem 5.6.2.
■
Definition 5.6.2 (Local degree). Let f : S n → S n (n > 0). Suppose there exists y ∈ S n such that
f −1 (y) is finite, say, {x1 , . . . , xm }. Then let U1 , . . . , Um be disjoint neighborhoods of x1 , . . . , xm
that are mapped by f to some neighborhood V of y.
x1 y
U1 f V
x2
U2
The local degree of f at xi , denote as deg f |xi , is the degree of the map
f∗ : Z ∼
= Hn (Ui , Ui − {xi }) → Hn (V, V − {y}) ∼
= Z.
The two isomorphisms in the upper half come from excision, and the lower two isomorphisms come
from exact sequences of pairs.
Theorem 5.6.3. Let f : S n → S n with f −1 (y) = {x1 , . . . , xm } as in Definition 5.6.2, then we have
m
X
deg f = deg f |xi .
i=1
Remark. Thus, we can compute the degree of f by computing these local degrees.
Let’s work with some examples for our edification.
CHAPTER 5. HOMOLOGY 98
Lecture 31: Local Degree and Local Homology
Example. Consider S n and choose m disks in S n . Namely, we first collapse the complement of the
m disks to a point, and then we identify each of the wedged n-spheres with the n-sphere itself.
D1
D1 D3
D2
D3 D2
The result will be a map of degree m. We can see this by computing local degree.
x2 y
x1
x3
By choosing a good point in the codomain, we get one point for each disk in the preimage, and
the map is a local homeomorphism around these points which is orientation preserving. We could
likewise compose the maps to S n from the wedge with a reflection to construct a map of degree
−m.
Remark. We see that from the above construction, we can produce a map S n → S n in any degree.
Sn RP n RP n / RP n−1 ∼
= Sn
f
Sn RP n RP n / RP n−1 ∼
= Sn
x1 y y
x2
local homeomorphism
in neighborhood of
x1 , x2 ∈ f −1 (y)
CHAPTER 5. HOMOLOGY 99
Lecture 31: Local Degree and Local Homology
m
Z = Hn (S n , S n − {x1 , . . . , xm })
L
∼
= LES of a pair
i=1
excision ∼
=
m̀ m̀
Hn Ui , (Ui − {xi }) Hn (S n , S n − {y})
i=1 i=1
∼
= excision
homology of disjoint union ∼
=
L
Hn (Ui , Ui − {xi }) Hn (V, V − {y})
i
1 deg(f∗ )
P
(1, 1, . . . , 1) deg f = deg f |xi
where we trace around the outside of the diagram at the bottom, which just proves the result. ■
With degree, we have a very efficient way for computing the homology groups of CW complexes,
which is so-called cellular homology. But before we dive into this, we first grab some intuition about the
essential of which, namely,
what really is local homology?
Since S n \{xi } is homeomorphic to an open n-ball, we see that Hk (S n \{xi }) = Hk−1 (S n \{xi }) = 0.
With this in mind, j∗ is an isomorphism.
We want to think about what j∗ does when k = n, i.e., when this is an isomorphism Z ∼ =
Hn (S n ) → Hn (S n , S n \ {xi }) ∼
= Z.
We see that ∆1 − ∆2 generate Hn (S n ), where ∆1 , ∆2 are the top and bottom hemisphere
indicated below.
Hn (S n , S n \ {x})
x1 ∆1
∆1 − ∆2 = ∆1
∆2
∆1 − ∆2
Hn (M, M \ {x}) ∼
= Hn (U, U \ {x}) ∼
= Hn (Rn , Rn \ {x}).
And since Rn \{x} is homotopy equivalent to an n−1 sphere, this means that Hn (Rn , Rn \{x}) ∼
= Z.
By homework, this connecting homomorphism is given by taking the boundary of a relative cycle
as below.
u u
x δ x
Hn (Rn , Rn − x) Hn (Rn − x)
∆ ∂∆
an n − 1-sphere around x in u
We intuitively want to use this idea to compute degree using this idea. We use naturality of the
long exact sequence, namely the fact that where f : (Ui , Ui \ {xi }) → (V, y) is a map of pairs, then
the following diagram commutes.
By naturality of the long exact sequence and the isomorphism discussed above, we can compute the
local degree of a map S n → S n at a point x by computing the degree of the map
In fact the local degree will be the degree restricted to a small S n−1 n the neighborhood U .
U V
n−1
S
x f y
f (S n−1 )
e→C
f: C e
z 7→ z n .
We see that
deg f |0 = n.
zn wound n times
0
C C
Definition 5.7.1 (Cellular chain complex). The cellular chain complex on X is the chain complex
with cellular chain group and cellular boundary map defined as follows.
Definition 5.7.2 (Cellular chain group). The chain groups Cn (X) defined as
∂1 : C1 (X) → C0 (X)
⟨1-cell⟩ → ⟨0-cell⟩,
which is the usual simplicial boundary map.a For n > 1, the boundary map ∂n are defined as
X
∂n (enα ) = ∂αβ en−1
β
β
attaching quotient by
∂enα = Sαn−1 map X n−1 Sβn−1
X n−1 \ en−1
β
eβn−1
en−1
β
enα
Sαn−1 X n−1
∂αβ = deg
Hn−1 (∂Dαn ) ∼
= Hn−1 (S n−1 ) ∼
= Z.
Note. In Hatcher [HPM02], the approach of the definition of cellular chain complex is aP bit different,
especially for how we define the boundary maps. Here we simply define ∂n (enα ) := β ∂αβ eβn−1 ,
where this is so-called cellular boundary formula in Hatcher [HPM02]. Here, we just defined ∂n in
this way instead, but we should still check that this is well-defined of this definition. The proof is
given in Appendix A.3.
Definition 5.7.4 (Cellular homology group). We define the so-called cellular homology group by cel-
lular chain complex in our usual way of defining homology group.
Remark. We sometimes denote the cellular homology group as HnCW (X) if it causes confusion.
Proof. We need to check two things, namely the chain group Hn (X n , X n−1 ) defined in Defini-
tion 5.7.1 is indeed free Abelian with basis in each n-cell. But this is trivial since we have an
one-to-one correspondence with the n-cells of X as we have shown, and we can think of elements
of Hn (X n , X n−1 ) as linear combinations of n-cells of X.
The fact that the boundary map defined in Definition 5.7.1 has the property ∂ 2 = 0 will be
proved in Theorem 5.7.2. ■
Theorem 5.7.2 (Cellular homology agrees with singular homology). The cellular homology groups
coincide with the singular homology groups, i.e.,
HnCW (X) ∼
= Hn (X).
• If X has a CW complex with no cells in consecutive dimensions, then all ∂n = 0. Its homology
are free Abelian on its n-cells, namely the cellular chain groups.
Example. The last point in Corollary 5.7.1 is quite useful, as the following examples will show.
(a) S n , n ≥ 2. Since if we have S n with n ≥ 2, using the CW complex structure of en attached
to a single point x0 . The cellular chain complex is given as
So then all the boundary maps are zero, and we see that
(
n Z, if k = 0, n;
Hk (S ) =
0, otherwise.
(b) CP n , ∀n. In this case, we can let CP n equipped with a CW complex structure with one cell
of each even dimension 2k ≤ 2n, thus
(
n ∼ Z, if k = 0, 2, . . . , 2n;
Hk (CP ) =
0, otherwise.
(c) S n × S n , n > 1. We let S n × S n has the product CW structure consisting of a 0-cell, two
if k = 0, 2n;
Z,
n ∼
Hk (CP ) = Z2 , if k = n;
otherwise.
0,
Exercise. Redo this calculation for S n with other CW complex structures on S n , e.g. glue 2 k-cells
onto S k−1 and proceed inductively.
Proof. Let the torus equips with the following CW complex structure.
a D a
x b
a
a
b deg 0
∂D
We wind forwards then backwards around a,a so the degree is zero. The same thing happens
for b, so
0 · a + |{z}
∂2 D = |{z} 0 · b = 0,
∂αβa a ∂αβb b
where we assume that α is the index of D, and βa is the index of a and same for b.
This gives a nice principle, namely if a 2-cells D is glued down via some words w (this only
makes sense for 2-cells), then the coefficientb to a letter a in ∂2 D is the sum of the exponents of a
in w. In this case, for both a and b, the coefficients for are both 1 + (−1) = 0.
Now we just have that the homology groups are equal to the chain groups because the boundary
if k = 0, 2;
Z,
Hk (T ) = Z2 , if k = 1;
otherwise.
0,
⊛
a Intuitively, since we quotient out b, hence the gluing map is homotopy to constant maps.
b i.e. ∂αβ (a) where α is the index of a.
Example (Cellular homology group of Σg ). Calculate the cellular homology group of a genus g surface
Σg .
Example (Torus example: ∂2 in more detail). We’re going to work through this example a bit more
carefully.
y
x2 x1
want degree of this map
S1
Let’s zoom in on these two preimage points and use local homology to compute this: Fill this
up!
We’re now going to work towards proving that cellular homology agrees with singular homology. First
we need some nontrivial preliminaries.
(b) Hk (X n ) = 0 for all k > n. If X is finite dimensional, then Hk (X n ) = 0 for all k > dim X.
Xn
.
∼
X n−1 = wedge of one n-sphere for each n-cell.
The result then follows from Theorem 5.5.5 and its immediately corollary, namely
!
e n (Xα ) ∼
M _
H =Hen Xα
α α
Hk (X n ) Hk (X n , X n−1 ) ...
When k + 1 < n or k > n then Hk+1 (X n , X n−1 ) = 0 and Hk (X n , X n−1 ) = 0, so the above
map Hk (X n−1 ) → Hk (X n ) is an isomorphism. We also get sequences telling us the injective and
surjective maps when k = n or k = n − 1,
So the maps Hn (X n−1 ) → Hn (X n ) is injective, and the map Hn−1 (X n−1 ) → Hn−1 (X n ) is surjec-
tive.
Fix k, then we get a pile of maps induced by the inclusions X n ,→ X n+1
Hk (X 0 ) ∼
= Hk (X 1 ) ∼
= Hk (X 2 ) ∼
= ...
∼
=
Hk (X k+2 ) ∼
= Hk (X k+3 ) ∼
= ...
Note. This sequence is not exact. Descriptions of maps (in red) follow from our analysis of
the long exact sequence of a pair above.
To prove (b),
• k = 0, we do this by hand.
• k ≥ 1, then Hk (X 0 ) = 0, so we have that Hk (X 0 ), . . . , Hk (X k−1 ) are all zero from the
isomorphisms above. That is the k-th homology Hk (X n ) = Hk (X n ) is zero for every n-
skeleton where n < k, just as desired.
We also have the following collection of maps for fixed k
surj. ∼
= ∼
=
Hk (X k ) Hk (X k+1 ) Hk (X k+2 ) ...
This implies (c) when X is finite dimensional. For general X, we use the fact that every simplex
has image contained in some finite skeleton (since image is compact). ■
Exercise. Check (b) and (c) in Lemma 5.7.1 directly in the case that the CW complex structure is
a ∆-complex structure using simplicial chains.
We now prove Theorem 5.7.2.
Proof of Theorem 5.7.2. We get some exact sequences from our preliminaries,
These come from the long exact sequences of a pair combined with the things we’ve deduced in the
preliminaries. We can paste these together into a diagram, we have
0
0 Hn (X n+1 ) ∼
= Hn (X)
Hn (X n )
∂n+1 jn
dn+1
... Hn+1 (X n+1 , X n ) Hn (X n , X n−1 ) dn
Hn−1 (X n−1 , X n−2 ) ...
∂n jn−1
Hn−1 (X n−1 )
0
Hatcher [HPM02] tells us this diagram commutes, and what we’ve done here tells us that the two
red diagonal pieces crossing at Hn (X n ) are exact. We also have exactness of the bottom right
diagonal by just going down a degree.
Then the horizontal row has to at least be a chain complex since the diagram commutes, and
we have
dn ◦ dn+1 = (jn−1 ◦ ∂n ) ◦ (jn ◦∂n+1 ) = 0,
| {z }
0
ker dn = ker ∂n = Im jn .
Then we just do some group theory, the n-th cellular homology group is
Hn (X n )
. . .
ker dn ∼ Im jn ∼ ∼
Im dn+1 = Im(jn ◦ ∂n+1 ) = Im ∂n+1 = Hn (X).
There is one thing left to show, namely commutativity of this map. We claim that the differentials
dn = jn ◦ ∂n+1 satisfy the formula (in terms of degree) that we stated. This is done by direct
analysis of definitions of maps; details in Hatcher [HPM02]. ■
a This is the missing part of the proof of Theorem 5.7.1.
F, G : C → D,
F (Y ) ηY
G(Y )
equipped with natural transformations ∂ : Hn (X, A) → Hn−1 (A), where Hn−1 (A) := Hn−1 (A, ∅),
is called the boundary map.
Naturality here means that for any map f : (X, A) → (Y, B) we have a commutative diagram
∂
Hn (X, A) Hn−1 (A)
f∗ f∗
Hn (Y, B) ∂
Hn−1 (B)
ι : (X \ U, A \ U ) ,→ (X, A)
induces isomorphisms on Hn .
Definition 5.8.3 (Extraordinary homology theory). If H∗ satisfies all axioms but dimension, it is called
an extraordinary homology theory.
Theorem 5.8.1. If Hn : CW pairs → Ab is a homology theory and H0 (∗) = Z, then Hn are exactly
the singular homology functors up to a natural isomorphism of functors.
More generally, if H0 (∗) = G, then Hn are exactly the singular homology functors with coeffi-
cients in the Abelian group G.
Proof. Given H∗ , reconstruct the cellular chain groups Hn (X n , X n−1 ) using the axioms.
• Show the homology of this chain complex are the cellular homology groups of X.
• Show these agree with Hn (X n , X n−1 ). The exact same argument in Theorem 5.7.2 applies.
We then check that the cellular homology groups we just constructed satisfies the degree formula
as in our last step. This is a bit more difficult, but we won’t get into it. ■
After developing homology theory rigorously, we now see an incredibly useful and interesting application
and some of its implication as well.
6.1 Trace
Definition 6.1.1 (Trace). Let φ : Zn → Zn be a group homomorphism, we may represent this with
a matrix A = [aij ]i,j with trace being
Exercise. We have tr(AB) = tr(BA) and tr(A) = tr(BAB −1 ). Thus, trace is independent of change
of basis of Zn .
Example (Euler characteristic via homology). When f ≃ idX , we have f∗ = idHk (X) for all k. Then
111
Lecture 37: Simplicial Approximation
Definition 6.2.2 (Euler characteristic). Let X be a finite CW complex, then the Euler characteristic
χ(X) of x is defined by the alternating sum
X
χ(X) = (−1)n cn
n
Proposition 6.2.1. Definition 6.2.2 and the definition given in this example coincides
Proof. This is immediately by the fact that the homology group can be calculated by cellular
homology, hence cn is just rank(Hn (X)) for all n, so the result follows. ■
Theorem 6.2.1 (Lefschetz Fixed Point Theorem). Suppose X admits a finite triangulation,a or more
generally, X is a retract of a finite simplicial complex. If f : X → X is a map with τ (f ) ̸= 0, then
f has a fixed point.
a i.e. a finite simplicial complex structure.
Note. Note that the converse does not hold. And in particular, we have
X
τ (f ) = tr(f# ⟳ CkCW (X)).
k
Theorem 6.2.2. If X is a compact, locally contractible space that can be embedded in Rn for some
n, then X is a retract of a finite simplicial complex.
Remark. This includes compact manifolds and finite CW complexes. Hence, we see that Theo-
rem 6.2.1 applies on compact manifolds and finite CW complexes in particular.
Definition 6.2.3. Let F be a field, and let Hk (X; F) be the k-th homology of X with coefficients in
F. Then Hk (X; F) is always a vector space over F. Define τ F (X) be
X
(−1)k tr(f∗ : Hk (X; F) → Hk (X; F)).
k
Remark. The Lefschetz fixed point theorem still holds if we replace τ (X) ̸= 0 with τ F (X) ̸= 0.
Answer. Since the homology of f is concentrated in degree zero, the result follows. ⊛
Answer. From the last exercise, we have τ (f ) = 1. And since we have compactness, by Theo-
rem 6.2.1, the result follows. In particular, this recovers Brouwer’s Fixed Point Theorem. ⊛
Example (QR May 2016). Let X be a finite, connected CW complex. X e is its universal cover, and
e is compact. Show that X
X e cannot be contractible unless X is contractible.
e = d · χ(X).
1 = χ(X)
Exercise. A 1-sheeted cover is always injective and surjective. Furthermore, it’s a local homeomor-
phism. This suffices to show that a 1-sheeted cover is a homeomorphism.
Theorem 6.2.3. If X is a finite CW complex, with cellular chain groups Hn (X n , X n−1 ). If we have
a cellular map f : X → X, so f induces maps f∗ : Hn (X n , X n−1 ) → Hn (X n , X n−1 ). Then
X
τ (f ) = (−1)n tr(f∗ : Hn (X n , X n−1 ) → Hn (X n , X n−1 )).
n
0 A B C 0
α β γ
′
0 A B′ C′ 0
Definition 6.3.1 (Simplicial map). A simplicial map f : K → L is a continuous map that sends each
simplex of K to a (possibly smaller dimensional) simplex of L by a linear map in the form of
X X
ti vi 7→ ti f (vi ).
Remark. A simplicial map is completely determined by its restriction to the vertex set.
That is, we add a new vertex to the center of every subsimplex, filling things in like the above.
Note. This means that when we go one dimensional higher, the new vertex added in each
subsimplex carrying through to the next dimension as well.
For an n-simplex we end up with (n + 1)!-simplices with replace it.
(a) We first reduce to the case of a finite simplicial complex X. Suppose K is a finite simplicial
complex, with r : K → X a retraction. First notice that the following composite of maps
r f ι
K X X K
This implies that τ (f ) = τ (ι ◦ f ◦ r), therefore if we can prove the result for a finite simplicial
complex then we are done.
(b) Let X be a finite simplicial complex. We show that if f : X → X has no fixed points, then
τ (f ) = 0. The goal now is to find subdivisions K, L of X and g : K → L so that
• g is simplicial.
• g ≃ f , τ (f ) = τ (g).
• g(σ) ∩ σ = ∅ for all simplices σ.
So this becomes a few steps, none of which we’ll justify too formally. Firstly, since trace is
given by diagonal entries of g in a matrix which respects to basis of simplices of K and L, if
g fixes no basis entries, then it has trace 0. With this, we have the following steps.
• Choose a metric d on X.
• Since X is compact, and f has no fixed point, then d(x, f (x)) has some minimum value
ϵ > 0.
• Subdivide all simplices of X until simplices have diameter smaller than ϵ/100.b Call this
subdivision L.
• Use the simplicial approximation theorem to obtain a map g : K → L, where K is a
subdivision of L and g ≃ f .
• By the proofc of simplicial approximation theorem, we can construct g so that for all
simplices σ, g(σ) is not too far from f (σ). We can then conclude that g(σ) ∩ σ = ∅.
• Consider g# ⟳ C∗CW (K), so then g is a cellular map K → K that moves every cell. We
can then check that
X
τ (f ) = τ (g) = (−1)n tr(g# : cellular n-chains → cellular n-chains) = 0
| {z }
g# ⟳C∗CW (K)
■
b Just a random constant!
c We omit the proof, see Hatcher [HPM02].
Epilogue
Φ : ∗ π1 (Aα , x0 ) → π1 (X, x0 ).
α
We want to show that Φ is surjective. Take some γ : I → X, then by the compactness of the
interval I, we can show that there is a partition I with s1 < · · · < sn so that
γ|si ,si+1 =: γi
Exercise. Showing the above fact is a good exercise for point-set topology.
Specifically, since
• Aα is open for all α
• I is compact,
then for all i, we choose a path hi from x0 to γ(si ) in Aσi−1 ∩ Aαi , using path-connectedness
of the pairwise intersections. Now, take γ and write it as
Observe that each of these paths is fully contained in Aαi , so this implies that γ ∈ Im(Φ),
therefore Φ is surjective.
2. For the next step, we’ll show that the second part of Theorem 3.6.1. Assume that our triple
intersections are path-connected. We want to show that ker(Φ) is generated by
117
Lecture 39: Proof of Seifert-Van-Kampen Theorem
f ≃ f1 · f2 · · · · · fℓ .
We showed that every [f ] has a factorization in step 1 already. Now we want to show that
two factorizations
[f1 ] · · · · · [fℓ ] and [f1′ ] · · · · · [fℓ′′ ]
of [f ] must be related by two moves:
(a) [fi ] · [fi+1 ] = [fi · fi+1 ] if [fi ], [fi+1 ] ∈ π1 (Aα , x0 ). Namely, the reaction defining the free
product of groups.
(b) [fi ] can be viewed as an element of π1 (Aα , x0 ) or π1 (Aβ , x0 ) whenever
[fi ] ∈ π1 (Aα ∩ Aβ , x0 ).
Now, let Ft : I × I → X be a homotopy from f1 . . . fℓ to f1′ . . . fℓ′′ , since they both represent
[f ]. We subdivide I × I into rectangles Rij so that
for some αij using compactness. We also argue that we can perturb the corners of the squares
so that a corner lies only in three of the Aα ’s indexed by adjacent rectangles.
We also argue that we can set up our subdivision so that the partition of the top and bottom
intervals must correspond with the two factorizations of [f ]. We then perform our homotopy
one rectangle at a time.
Idea: Argue that homotoping over a single rectangle has the effect of using allowable moves
to modify the factorization.
At each triple intersection, choose a path from f (corner) to x0 which lies in the triple inter-
section, so we use the assumption that the triple intersections are path-connected.
Along the top and bottom, we make choices compatible with the two factorizations. It’s now
an exercise to check that these choices result in homotoping across a rectangle gives a new
factorization related by an allowable move.
≃ ≃
collapse
We then see that T #T ′ is a wedge of T and S 1 , hence the fundamental group is just Z2 ∗ Z,
where Z2 is from T and Z is from S 1 . ⊛
Exercise (QR Aug. 2019). Let X be a CW complex obtain from a k-sphere, k ≥ 1, by attaching
two (k + 1)-cells along attaching maps of deg m, n. Calculate H∗ (X).
(1, 0) m
(0, 1) n
Exercise (QR May 2017). Let S 1 be the complex numbers of absolute value 1 with induced topology.
Let K be the quotient of S 1 × [0, 1] by identifying (z, 0) with (z −2 , 1). Compute H∗ (K).
Answer. Again, we use cellular homology. We use the following gluing instruction.
b b
a D a
Exercise (QR Sep. 2016). Let Z = {(x, y) ∈ C2 | x = 0 or y = 0}. Find H∗ (C2 \ Z).
Exercise (QR May 2017). Let X be the connected CW complex, Hi (X) = 0 for all i > 0. Prove
that (
Z, if n = 0;
Hn (X × S k ) =
0, otherwise.
Answer. We can induct on k and use Mayer-Vietoris long sequence. Let U = X × upper hemisphere,
V = X × lower hemisphere, then we have
U ≃ X, V ≃ X,
Exercise (QR Jan. 2021). Let G be a connected topological space and w a topological group struc-
ture, i.e., continuous multiplication M : G × G → G and inverse i : G → G that define a group
structure. Assume G has finite CW complex structure. Show χ(G) = 0 unless G = {1}.
Answer. Let g ∈ G such that g ̸= id. We can then choose a path γ from g to id. Then we have
τg : G → G
h 7→ gh.
Exercise (QR May 2019). For what g ≥ 0 is it true that for all h ≥ g, a compact oriented genus-g
surface X (no boundary) has covering f : Y → X where Y is a compact, oriented surface of genus-h?
(2 − 2g)d = 2 − 2h
for a degree d covering. This implies h ≡ 1 (mod g − 1). We see that the only solutions for all h if
g = 2. Then for a particular d, we have the following cover.
2π
d
Exercise (QR Jan. 2019). Let S1 , S2 be two disjoint copies of n-sphere, n > 1. Choose distinct
points Ai , Bi ∈ Si , and let Z is obtained by gluing A1 ∼ A2 , B1 ∼ B2 . What’s the lowest number
of cells in CW complex structure on Z.
≃ ≃
Exercise (QR Jan. 2016). Let U, V ⊆ S n , n ≥ 2 be non-empty connected open sets such that
S n = U ∪ V . Show U ∩ V is connected.
(a) If χ(X) is not divisible by p, show that the action has a (global) fixed point.
(b) Give an example of such an action that is fixed point free when χ(X) = 0.
⊛
a Since we assume has no fixed point, with the fact that X admits a CW complex structure hence Hausdorff, so
Z / pZ is indeed a covering space action.
123
Appendix A
Additional Proofs
j0 Π(j0 )
(X0 ∩ X1 , x0 ) (X0 , x0 ) Π(X0 ∩ X1 ) Π(X0 )
Π
j1 i0 7−→ Π(j1 ) Π(i0 )
(X1 , x0 ) i1
(X, x0 ) Π(X1 ) Π(X)
Π(i1 )
F : Π(X0 ) → G , G : Π(X1 ) → G
such that
F ◦ Π(j0 ) = G ◦ Π(j1 ).
Π(j0 )
Π(X0 ∩ X1 ) Π1 (X0 )
Π(j1 ) Π(i0 )
F
Π1 (X1 ) Π1 (X)
Π(i1 )
∃!K
G G
We now only need to prove that there exists a unique functor K : Π(X) → G such that the above
diagram commutes.
We can define K as
124
Lecture 41: Prepare for Final
• on morphisms: For every p, q ∈ X, ⟨γ⟩ : p → q in HomΠ(X) (p, q), we need to define K(⟨γ⟩) ∈
HomG (K(p), K(q)). Our strategy is for every path γ from p to q, we define K(γ)
e ∈ HomG (K(p), K(q)).
Then if we also have K(γ) = K(γ ) for γ ≃ γ rel{0, 1}, then we can just let
e e ′ ′
K(⟨γ⟩) := K(γ).
e
X
γ
0 t0 1 p q
γ(t0 )
X0 X1
0 = t0 = t10 < t11 < · · · < t1K1 = t1 = t20 < t21 < · · · < tmKm = tm = 1.
As before, we denote γij : [0, 1] → X, t 7→ γ((1 − t)tij−1 + t · tij ). It’s clear that as long as
K(γ e iK ) ◦ K(γ
e i ) = K(γ e iK −1 ) ◦ · · · ◦ K(γ
e i0 ),
i i
then our claim is proved. But this is immediate since F and G are functor and for any i, we
only use either F or G all the time.
Now we prove γ ≃ γ ′ rel{0, 1}, then K(γ)
e e ′ ). This is best shown by some diagram.
= K(γ
H
H1 = γ ′ (1, 1)
X
H γ
γ′
(0, 0) H0 = γ
The left-hand side represents a partition P of [0, 1]×[0, 1] such that every small square’s image
in X under H is either entirely in X0 or in x1 . Consider all paths from (0, 0) to (1, 1) such
that it only goes right or up. We see that for any such path L, consider
γL : [0, 1] → L, t 7→ γL (t).
We let ΓL : H|L ◦ γL : [0, 1] → X, we see that ΓL is a path from p to q. Now, if for two paths
L1 and L2 such that they only differ from a square.
H1 = γ ′ (1, 1)
L1
L2
(0, 0) H0 = γ
We claim that γL1 , ΓL2 are two paths from p to q, and K(Γ e L ). Now, we denote
e L ) = K(Γ
1 2
Γ0 and Γ1 as follows.
H1 = γ ′ (1, 1) H1 = γ ′ (1, 1)
(0, 0) H0 = γ (0, 0) H0 = γ
Γ0 Γ1
It’s clearly that by only finitely many steps, we can transform Γ0 to Γ1 , hence
K(Γ
e 0 ) = K(Γ
e 1 ).
Π(j0 )
Π(X0 ∩ X1 ) Π1 (X0 )
Π(j1 ) Π(i0 )
F
Π1 (X1 ) Π1 (X)
Π(i1 )
K
G G
(j0 )∗
π1 (X0 ∩ X1 , x0 ) π1 (X0 , x0 )
(j1 )∗ (i0 )∗
π1 (X1 , x0 ) π1 (X, x0 )
(i1 )∗
is cocartesian.
Proof. The basic idea is that, for this diagram,
Π(X0 ∩ X1 ) Π(X0 )
Π(X1 ) Π(X)
(d) γp = cp
The proof is given in https://www.bilibili.com/video/BV1P7411N7fW?p=38&spm_id_from=
pageDriver. ■ If have
time.
is well-defined.
Proof. Here we are identifying the cells enα and en−1
β with generators of the corresponding summands
of the cellular chain groups, namely Cn (X). The summation in the formula contains only finitely
many terms since the attaching map of enα has compact image, so this image meets only finitely
many cells en−1
β . To derive the cellular boundary formula, consider the following commutative
diagram.
∂ ∆αβ ∗
Hn (Dαn , ∂Dαn ) e n−1 (∂Dαn )
H e n−1 (S n−1 )
H
∼
= β
Φα ∗ φα ∗ qβ ∗
∂n q∗
Hn (X n , X n−1 ) e n−1 (X n−1 )
H e n−1 (X n−1 / X n−2 )
H
jn−1 ∼
=
dn
∼
=
Hn−1 (X n−1 , X n−2 ) Hn−1 (X n−1 / X n−2 , X n−2 / X n−2 )
where
• Φα is the characteristic map of the cell enα and φα is its attaching map.
• ∆αβ : ∂Dαn → Sβn−1 is the composition qβ qφα , i.e., the attaching map of enα followed by the
quotient map X n−1 → Sβn−1 collapsing the complement of en−1 β in X n−1 to a point.
The map Φα∗ takes a chosen generator [Dαn ] ∈ Hn (Dαn , ∂Dαn ) to a generator of the Z summand
of Hn (X n , X n−1 ) corresponding to enα . Letting enα denote this generator, commutativity of the left
half of the diagram then gives
∂n (enα ) = jn−1 φα ∗ ∂[Dαn ].
In terms of the basis for Hn−1 (X n−1 , X n−2 ) corresponding to the cells en−1
β , the map qβ ∗ is the
projection of Hn−1 (X
e n−1
/X n−2
) onto its Z summand corresponding to eβ . Commutativity of
n−1
Algebra
This section aims to give some reference about the required algebra content, including Abelian groups,
and free Abelian group, which is used heavily when discuss homology, and also some homological algebra,
where we will focus on exact sequence specifically.
a · b = b · a.
Definition B.1.2 (Product of groups). Given two groups (G, ·), (H, ·), the product of G and H,
denoted by G × H is defined as
G × H = {(g, h) | g ∈ G, h ∈ H}
and
(g1 , h1 ) · (g2 , h2 ) := (g1 · g2 , h1 · h2 ).
Notation. For simplicity, given an index set I, we’ll denote the order pair (gα1 , gα2 , . . . ) as (gα )α∈I .
Note that the latter notation can handle the case that I is either countable or uncountable, while
the former can only handle the countable case.
Definition B.1.3 (Direct product). Given (Gα , +), α ∈ I as a collection of Abelian group, we define
their direct product as !
Y
Gα , + ,
α∈I
where Y
Gα = {(gα )α∈I | gα ∈ Gα }
α∈I
129
Lecture 41: Prepare for Final
n
and can be denoted as
Q
Gi , +
i=1
(G1 × · · · × Gn , +) .
M
Gα := (gα )α∈I | ∀ gα ∈ Gα , # non-zero elements in gα < ∞ .
α∈I
α∈I
(gα ) + (hα ) := gα + hα
Remark. We see that the operation + is indeed closed since the sum of g, g ′ ∈ α∈I Gα will have
L
only finitely non-zero elements if g, g ′ both have only finitely many non-zero elements.
We see that if I is a finite index set, given a collection of Abelian group {Gα }α∈I , then
G1 × · · · × Gn = G1 ⊕ · · · ⊕ Gn .
Definition B.1.5 (Internal direct sum). Given an Abelian group G, and a collection of the subgroups
{Gα }α∈I of G, we say G is an internal direct sum of {Gα }α∈I if for any g ∈ G, we can write
X
g= gα
α∈I
uniquely, where gα ∈ Gα has only finitely many non-zero elements. In this case, we denote
M
G= Gα .
α∈I
Intuitively, the external direct sum is to build a new group based on the given collection of groups
{Gα }α∈I , while the internal direct sum is to express an already known group G with an already
known collection of groups {Gα }α∈I .
Remark (Relation between InternalLand External direct sum). Given an Abelian group G and its
internal direct sum decomposition α∈I Gα , G is isomorphic to the external direct sum of {Gα }α∈I .
We see this from the following group homomorphism:
X
∀ g= gα 7→ (gα )α∈I .
g∈G
α∈I
Conversely, given a collection of Abelian group {Gα }α∈I , and let G = Gα as the external
L
α∈I
where (
gα0 , if α0 = α;
hα =
0, if α0 ̸= α
given α0 . Then M
G′α0 := iα0 (Gα0 ) < Gα
α∈I
and G is the internal direct sum of G′α0 , α0 ∈ I. This is because ∀g = (gα )α∈I ∈ G(=
L
α∈I Gα ),
we have X
g= iα (gα ).
α∈I
Note that the above sum is well-defined since there are only finitely many non-zero elements for
each gα . And additionally, we can see the uniqueness of this decomposition by defining πα0 such
that M
πα0 : Gα → Gα0 , (gα )α∈I 7→ gα0 ,
α∈I
for all β ∈ I, where the second equality is because this summation is finite. Hence, we have
X
g= iα (πα (g)).
α∈I
where
(φ + ψ)(g) := φ(g) + ψ(g).
Remark (Relation between direct sum and direct product). Given a collection of Abelian groups
{Gα }α∈I , and another Abelian group H, there exists a φ such that
!
M Y
φ : Hom Gα , H → Hom(Gα , H)
α∈I α∈I
f 7→ φ(f ) := (fα )α∈I
• φ is injective. This is obvious since ker(φ) = 0 from the fact that if φ(f ) = 0, then fα = 0 for
all α, hence f is 0.
Remark. If G is a free Abelian group, and {gα }α∈J is a basis, then for every α ∈ J, ⟨gα ⟩ is an
infinite cyclic group since
n · gα = 0 = 0 · gα ⇒ n = 0.
And from Definition B.2.1, we have M
G= ⟨gα ⟩.
α∈J
Conversely, assume there are a collection of infinite cyclic group ⟨gα ⟩ for α ∈ I in G such that
M
G= ⟨gα ⟩,
α∈I
Moreover, denote ϵα as the image of eα under the isomorphism u, namely ϵα = u−1 (eα ), then {ϵα }α∈I
is a basis of G.
Now, for every Abelian group H, we have
◦u
f ◦ u−1
L
Hom(G, H) ∼
=
Hom Z, H f
α∈I
∼
= φ
(f ◦ u−1 ◦ iα )α∈I
Q
∼
=
Hom (Z, H)
α∈I
∼
=
In other words, for all Abelian group H, a morphism from the set {ϵα }α∈I to H can be uniquely extended
to the group a homomorphism from G to H.
Remark. This means, to determine Hom(G, H), we only need to determine where each base element
in G will map to in H, and this is why it’s free.
We now want to generate free Abelian group by a set. Roughly speaking, given a set S, we can
generate a free Abelian group Z by defining
( )
X
Z := nx x | nx ∈ Z, # non-zero elements in nx < ∞
x∈S
Definition B.2.2 (Free Abelian group generated by a set). Given a set S, the free Abelian group
generated by S (Z, +) is defined as
with
+: Z × Z → Z
(f, g) 7→ f + g.
is the characteristic function at x. We see this by for all f ∈ S, f = f (x)ϕx , which is uniquely
P
x∈S
defined. Hence, (Z, +) is a free Abelian group.
X
nx · x.
x∈S
Theorem B.2.1 (The universal property of free Abelian group generated by a set). Denote a canonical
embedding i : S → Z, x 7→ ϕx . Then for all Abelian group H and f : S → H, there exists a unique
group homomorphism
fe: Z → H
such that fe ◦ i = f .
Proof. We define !
X X
fe nx · x := nx f (x),
x∈S x∈S
Proposition B.2.2. Given Z ′ as another Abelian group and i′ : S → Z ′ as another canonical embed-
ding such that for all Abelian group H and f : S → H, there exists a unique group homomorphism
fe: Z ′ → H such that fe ◦ i′ = f , then
Z′ ∼
= Z.
Namely, we can describe a free Abelian group by its universal property uniquely up to isomorphism.
Theorem B.2.2. Assume G is a free Abelian group. Assume there exists a finite basis {g1 , . . . , gn }
of G, and also assume that there exists another basis {hα }α∈I . Then we have
card(I) < ∞,
specifically, we have
card(I) = n.
Proof. Firstly, we observe that if we can show
card(I) ≤ n,
then by swapping {hα }α∈I and {gα }α∈I , we will have card(I) = n.
Suppose I is an infinite set, then we can find hα1 , . . . , hαm such that m > n and hαi ̸= hαj for
Specifically, we have 1
k12 . . . k1n
hα1 k1 g1
.. .. .. .. .. ,
. = . . . .
1 2 n
hαm km km . . . km gn
| {z }
K∈Mm×n (Z)⊂Mm×n (Q)
where kij ∈ Z. From linear algebra, we know that there exists (r1 , . . . , rm ) ∈ Qm \ {⃗0} such that
Multiplying both sides with the common multiple of the denominator of ri , we see that there exists
(ℓ1 , . . . ℓm ) ∈ Zm \ {⃗0} such that
(ℓ1 , . . . ℓm )K = (0, . . . , 0)
hα1 g1
.. ..
⇒(ℓ1 , . . . , ℓm ) . = (ℓ1 , . . . , ℓm )K . = (0, . . . , 0)
hαm gn
m
X
⇒ ℓi hαi = ⃗0 for (ℓ1 , . . . , ℓm ) ∈ Zm \ {⃗0}
i=1
⇒ card(I) < ∞.
Remark. Furthermore, one can prove that if G is a free Abelian group, then we can prove that any
two bases of G are equinumerous, which handle the case that the basis is an infinite set.
This induces the following definition.
Definition B.2.3 (Rank). Let G ba a free Abelian group, the rank of G is the cardinality of any
basis of G.
Definition B.3.1 (Torsion subgroup). Given an Abelian group G, we say that g ∈ G has finite order
if ∃n ∈ Z such that n · g = 0. Specifically, we say that
is a torsion subgroup.
If T = 0 given G, we say that G is torsion free.
Note. Note that T is indeed a subgroup, since for any g1 , g2 ∈ T , g1 + g2 ∈ T from the fact that it
still has finite order.
Remark. If G is a free Abelian group, then G is torsion free. Conversely, if G is torsion free, we can’t
deduce G is a free Abelian group. We see this from (Q, +). Firstly, we see that Q is torsion free,
Now, suppose Q is a free Abelian group, then there exists a basis {rα }α∈I of Q such that |I| > 1.
Now, consider α1 , α2 ∈ I such that α1 , α2 ∈ I, for rα1 , rα2 , there exists n, m ∈ Z and n, m ̸= 0 such
that
nrα1 + mrα2 = 0 ⇒ n = m = 0
F := Z ⊕ · · · ⊕ Z,
| {z }
n times
φ : F → G, ei 7→ gi
en n×n
bm1 . . . bmn
then we have
b1 e1
.. ..
. = B . .
bm en
Multiply a matrix on the right-hand side. Now, consider a p ∈ GL(n; Z), then
′
b1 e1 e1
.. −1 .. ..
. = B · P P . = (BP ) · . ,
bm en e′n
| {z }
new basis
where ′
e1 e1
−1 .. ..
P . =: . .
en e′n
We see that B · P is the coefficient matrix of generators b1 , . . . , bm under the new basis e′1 , . . . , e′n .
since Q is invertible, hence b′1 , . . . , b′m are also generators of K. We see that QB is the coefficient matrix
of new generators b′1 , . . . , b′m under basis e1 , . . . , en .
where di ∈ Z+ and d1 | d2 | · · · | dk .
Proof. In fact, P, Q can be taken as the multiplication of the following three types of square
matrices:
• Pij :
1
..
.
0 1(ij)
..
Pij = . ,
1(ji) 0
..
.
1
where the effect of multiplying Pij from the right is swapping column i, j.
where the effect of multiplying Pi (c) from the right is multiplying c to column i.
• Pij (a), a ∈ Z:
1
..
.
a(ij)
Pij = ,
..
.
1
where the effect of multiplying Pij (a) from the right is adding a times column i to column j.
We see that these are elementary column transformations in linear algebra. In particular, if we
multiply these matrices from the left, then it’s called elementary row transformations.
That is to say, we’re going to show
b11 . . . b1n
B = ... ..
.
..
.
can become
d1
..
.
dk ,
0
..
.
di ∈ Z+ , d1 | d2 | · · · | dk from elementary column/row transformations.
We now show the steps to make this happens.
• Step 1. Using elementary transformations, we make b11 > 0.
• Step 2. Using elementary transformations, we make b11 become a divisor of all elements in
the first column and row.
We see that if b11 ∤ b1i for i ̸= 1, we have b1i = r · b11 + s where 0 < s < b11 . Then we add
(−r) times the 1th column to the ith column and swapping the 1th and the ith column, which
makes B becomes !
s ...
.. . . ,
. .
for 0 < s < b11 . Since card({n ∈ Z | 0 < n < b11 }) < ∞, hence in finitely many steps we can
make B becomes !
d1 . . .
.. . . ,
. .
where d1 is a divisor of all other elements in the first column and row.
• Step 3. Using elementary transformations, we can multiply the first row by a proper integer
and add it to the other rows, do the same but for columns also, then we can make B becomes
d1 0 . . . 0
0
.. .
. B1
0
where di ∈ Z+ .
• Step 5. Using elementary transformations, by swapping columns and rows, we may assume
d1 ≤ d2 ≤ · · · ≤ dk .
• Step 6. Using elementary transformations, we can make B into
′
d1
..
.
′
dℓ
0
..
.
such that 0 < d′1 ≤ · · · ≤ d′ℓ , d′1 | d′2 | · · · | d′ℓ since if d1 ∤ di for some i ∈ {2, . . . , k}, then
d1 d1 di
.. ..
. .
dk → dk
,
0 0
.. ..
. .
where de1 ≤ · · · ≤ dej such that de1 < d1 . Since there are only finitely many integers which is
smaller than d1 , we see that by repeating these steps, we can always make
d1
..
.
dk
0
..
.
into e
d1
..
.
dep
0
..
.
such that d′1 | d′i for all i ̸= 1 and d′1 ≤ d′2 ≤ · · · ≤ d′p . By the same idea of Step 3., we have
the desired matrix.
Since all operations are elementary and there are only finitely many of them, hence the result
follows. ■
From the definition of Q · B · P and Proposition B.3.1, there exists a basis e′1 , . . . , e′n of F such that
K has finitely many generators d1 e′1 , . . . , dk e′k , hence
. . .
G∼= Z d Z ⊕ Z d Z ⊕ ··· ⊕ Z d Z ⊕ Z ⊕ ··· ⊕ Z.
1 2 | {z } k
n−k times
Theorem B.3.1 (Fundamental theorem of finitely generated Abelian group). Given a finitely generated
Abelian group, either G is a free Abelian group, or there exists a unique set of {mi ∈ Z | mi >
1, i = 1, . . . , t} such that m1 | m2 | · · · | mt and a unique non-negative integer s such that
. . .
G∼= Z m Z ⊕ Z m Z ⊕ ··· ⊕ Z m Z ⊕ Z ⊕ ··· ⊕ Z.
1 2 t | {z }
s times
Existence. From Proposition B.3.1, we obtain a basis e′1 , . . . , e′n of F and a basis d1 e′1 , . . . , dk e′k
in K such that d1 | · · · | dk . Let
(d1 , . . . , dk ) = (1, . . . , 1, m1 , . . . , mt ),
which implies
.
=F K
G∼
. . .
∼
= Z d Z ⊕ Z d Z ⊕ ··· ⊕ Z d Z ⊕ Z ⊕ ··· ⊕ Z
1 2 k
. . . .
Z Z Z Z
= 1Z ⊕ · · · ⊕ 1Z ⊕ m1 Z ⊕ · · · ⊕ mt Z ⊕ Z ⊕ · · · ⊕ Z
. .
= Z m Z ⊕ ··· ⊕ Z m Z ⊕ Z ⊕ ··· ⊕ Z.
1 t | {z }
∃!s times
Definition B.3.2 (Invariant factor). We call m1 , . . . , mt obtained from Theorem B.3.1 the invariant
factor.
m = pn1 1 · · · · · pns s
then
n n
n + ⟨pni i ⟩ = 1 + ⟨pni i ⟩, n + ⟨pj j ⟩ = 0 + ⟨pj j ⟩.
Above implies
ϕ(n) = (0, . . . , 0, 1 + ⟨pni i ⟩, 0, . . . , 0),
hence ϕ surjects, so . . .
Z ∼
= Z n ⊕ ··· ⊕ Z n .
mZ p1 Z
1
ps s Z
■
Combine Theorem B.3.1 and Lemma B.3.2, we see that we now only have
. . .
G∼= Z m Z ⊕ Z m Z ⊕ ··· ⊕ Z m Z ⊕ Z ⊕ · · · ⊕ Z,
1 2 t | {z }
s times
where p1 , . . . , pk are primes (which may includes repeated terms), si ∈ Z+ for all i.
{ps11 , . . . , pskk }
• coIm φ := A / ker φ
Consider a sequence of Abelian group homomorphism
ϕi−1 ϕi
... Ai−1 Ai Ai+1 ...
Im ϕi−1 = ker ϕi .
Definition B.4.2 (Exact sequence). We call S is an exact sequence if it’s exact at Ai for all i.
conversely,
ϕ
◦ B A is an exact sequence ⇔ ϕ is an injective homomorphism.
Definition B.4.3 (Short exact sequence). A short exact sequence is an exact sequence such that it
has the following form
ϕ ψ
◦ A B C ◦ .
ψ i
Remark. Let B −→ C as a surjective homomorphism and K = ker ψ, and we denote K −→ B as
an injection. Then
i ψ
◦ K B C ◦
is a short exact sequence. Conversely, if
ϕ ψ
◦ A B C ◦
i Proj2
◦ A A⊕B B ◦
i
a (a, 0)
Proj2
(a, b) b
i Proj2
◦ Z Z Z / nZ ◦
k k·n
Definition B.4.4 (Isomorphism between sequences). Given A· and B· defined as two sequences of
Abelian group homomorphisms
ϕi
A• : . . . Ai Ai+1 ...
and
ψi
B• : . . . Bi Bi+1 ...
And we say a morphism α from A• to B• is a series of group homomorphisms αi : Ai → Bi for
Definition B.4.5 (Split short exact sequence). Given a short exact sequence
ϕ ψ
◦ A B C ◦
◦ A A⊕C C ◦
A ⊕ C = i(A) ⊕ j(C).
∼
=
Consider θ−1 : A ⊕ C −→ B, then
hence
B = ϕ(A) ⊕ θ−1 (j(C)),
| {z }
D
ϕ ψ
◦ A B C ◦
∼
=
split implies B = ϕ(A) ⊕ D and ψ|D : D −→ C.
∼
=
Conversely, if B = ϕ(A) ⊕ D and ψ|D : D −→ C, then there exists a θ
θ: B → A ⊕ C
ϕ(a) + d 7→ (a, ψ(d))
◦ A A⊕C C ◦
ϕ(a) + d ψ(d)
commutes.
ϕ ψ
Hence, for a short exact sequence ◦ A B C ◦ is split if and only if
∼
=
B = ϕ(A) ⊕ D and ψ|D : D −→ C.
ϕ ψ
Remarkably, let ◦ A B C ◦ is a split short exact sequence, then D
constructed above is not unique. To see this, consider
i Proj2
◦ Z Z⊕Z Z ◦
n (n, 0)
(n, m) m
We have Z ⊕ Z = i(Z) ⊕ j(Z) where j : Z → Z ⊕ Z, n 7→ (0, n). We see that we can let D := j(Z).
Meanwhile, we can also let
D := {(n, n) | n ∈ Z} < Z ⊕ Z
such that Z ⊕ Z = i(Z) ⊕ D.
i Proj2
◦ Z Z Z / nZ ◦
k k·n
ϕ ψ
Lemma B.4.1 (Splitting lemma). If ◦ A B C ◦ is a short exact se-
quence, then the following are equivalent.
(a) This short exact sequence split.
(b) ∃p : B → A such that p ◦ ϕ = idA .
diagram commutes.
ϕ ψ
◦ A B C ◦
id θ id
i
◦ A A⊕C C ◦
Proj1
∼
=
• 1. ⇒ 3. Let θ : B −→ A ⊕ C such that it’s the isomorphism which makes the following diagram
commutes.
ϕ ψ
◦ A B C ◦
id θ id
Proj2
◦ A A⊕C C ◦
j
hence ψ ◦ q = idC .
• 2. ⇒ 1. We have
ϕ ψ
◦ A B C ◦
p
where p ◦ ϕ = idA . We claim that B = ϕ(A) ⊕ ker(p) since for every b ∈ B, ϕ(p(b)) ∈ ϕ(A),
and
b = ϕ(p(b)) + (b − ϕ(p(b)))
| {z } | {z }
∈ϕ(A) ∈ker(p)
ϕ(a1 − a2 ) = d2 − d1 ⇒ p(ϕ(a1 − a2 )) = 0 ⇒ a1 = a2 ⇒ d1 = d2 .
b = (b − q(ψ(b))) + q(ψ(b)),
| {z } | {z }
∈ker(ψ)=Im(ϕ) ∈q(C)
B = ϕ(A) ⊕ q(C)
similarly.
[AM94] M.F. Atiyah and I.G. MacDonald. Introduction To Commutative Algebra. Addison-Wesley
series in mathematics. Avalon Publishing, 1994. isbn: 9780813345444. url: https://books.
google.com/books?id=HOASFid4x18C.
[Arm13] M.A. Armstrong. Basic Topology. Undergraduate Texts in Mathematics. Springer New York,
2013. isbn: 9781475717938. url: https://books.google.com/books?id=NJbuBwAAQBAJ.
[HPM02] A. Hatcher, Cambridge University Press, and Cornell University. Department of Mathe-
matics. Algebraic Topology. Algebraic Topology. Cambridge University Press, 2002. isbn:
9780521795401. url: https://books.google.com/books?id=BjKs86kosqgC.
148