Master
Master
Master
Gilles Castel
Lecture 1: Introduction
di 01 okt 10:33
• First part: fundamental groups. We follow Munkres:
After the exam, hand in your solutions of the other problems. After a
quick look, the points of 3. and 4. can be ±1 (in extreme cases ±2)
• No exercise classes for this course!
1
Chapter 0
Introduction
• fundamental groups
• homology groups
g∗ ◦ f∗ = 1GX f∗ ◦ g∗ = 1GY ,
2
Lecture 1: Introduction
f∗ [α] = [f ◦ α].
r(x)
B 2 f (x)
x
CHAPTER 0. INTRODUCTION 3
Lecture 1: Introduction
The map from π(S 1 ) → π(S 1 ) is the identity map, but the first map maps
everything on 1.
CHAPTER 0. INTRODUCTION 4
Chapter 9
Fundamental group
Proof.
5
Lecture 1: Introduction
f
X
• Any two paths f, g : I → C with f (0) = g(0) and g(1) = f (1) are
path homotopic.
Choose H : X × I → C : (x, t) → H(x, t) = (1 − t)f (x) + tg(x).
Product of paths
Let f : I → X, g : I → X be paths, f (1) = g(0). Define
(
f (2s) 0 ≤ s ≤ 21
f ∗ g : I → X : s 7→
g(2s − 1) 12 ≤ s ≤ 1.
Theorem 2.
1. [f ] ∗ ([g] ∗ [h]) is defined iff ([f ] ∗ [g]) ∗ [h] is defined and in that case,
they are equal.
Then k ◦ f ≃p k ◦ g using k ◦ H.
• If f ∗ g (not necessarily path homotopic). Then k ◦ (f ∗ g) = (k ◦ f ) ∗
(k ◦ g).
Now, the proof
2. Take e0 : I → I : s 7→ 0. Take i : I → I : s 7→ s Then e0 ∗ i is a
path from 0 to 1 ∈ I. The path i is also such a path. Because I is
a convex subset, e0 ∗ i and i are path homotopic, e0 ∗ i ≃p i. Using
one of our observations, we find that
f ◦ (e0 ∗ i) ≃p f ◦ i
(f ◦ e0 ) ∗ (f ◦ i) ≃p f
ex0 ∗ f ≃p f
[ex0 ] ∗ [f ] = [f ].
f ◦ (i ∗ i) ≃p f ◦ e0
f ∗ f ≃p e x 0
[f ] ∗ [f ] = [ex0 ].
ka,b : [0, 1] −→ X
lin. f
[0, a] −−→ [0, 1] −
→X
lin. g
[a, b] −−→ [0, 1] −
→X
lin. h
[b, 0] −−→ [0, 1] −
→X
Then f ∗ (g ∗ h) = k 21 , 34 and (f ∗ g) ∗ h = k 41 , 12
Let γ be that path γ : I → I with the following graphs:
0 a b 1
kc,d ◦ γ ≃p kc,d ◦ i
ka,b ≃p kc,d ,
from x0 to x1 . Then
α̂ : π(X, x0 ) −→ π(x, x1 )
[f ] 7−→ [α] ∗ [f ] ∗ [α].
We can also construct the inverse, proving that these groups are isomor-
phic.
X
x0 f
x1
Remark. If trivial for one base point, it’s trivial for all base points.
Example (Wrong proof of π(S 2 ) being trivial). Let f be a path from [0, 1] →
S 2 . Then pick y0 ̸∈ Im(f ). Then S 2 \ {y0 } ≈ R2 . Then use R2 .
This is wrong because we cannot always find y0 ̸∈ Im(f ). Space filling
loops! We’ll see the correct proof later on.
And therefore α ≃p β. (Note: make sure end and start point matchs when
using ∗)
• Vα open in E
• Vα ∩ Vβ = ∅ if α ̸= β
• p|Vα : Vα → U is a homeomorphism.
Vα
p−1 (U )
p
U
b
S1
V1 V2
R
S1 S1
p
Here we have three copies for each point. We say that the covering has
3 sheets. Note that this is independent of which point we take. This is
always the case! We can show that these are the only coverings of S 1 : R
and z 7→ z n .
Proof. Exercise.
Example. Let T 2 = S 1 × S 1 .
These are the only types of coverings of the torus. We’ll prove this later
on.
(1, 1)
B0 = S 1 × {1} ∪ {1} × S 1
Figure 9.7: Covering of the torus. B0 is the figure 8 space: two connected
circles.
E
f˜
f
X B
E
e0
f˜(si )
f˜
p
B
f Ub
b0
I
s0 sn
si+1
si
• f˜(0) = e0
• Assume f˜ has been defined on [0, si ]. Let U be an open such that
f [si , si+1 ] ⊂ Ub .
There is exactly one slice Vα in p−1 (Ub ) containing f˜(si ). We define
∀s ∈ [si , si+1 ] : f˜(s) = (p|Vα )−1 ◦ f (s). By the pasting lemma, f˜ is
continuous.
• In this way, we can construct f˜ on the whole of I.
Uniqueness works in exactly the same way, by induction.
E
F̃
e0
p
F
B
b0
B
F
b0
• F (0, t) = b0 , F (1, t) = b1
We’ve shown that homotopy from below lifts to above. The converse is easy.
Now we’re ready to discuss the relation between groups and covering spaces.
e0 e1
E
f
b0 B
f
Example. Take the circle and the real line as covering space. Then p−1 (1) =
Z. So we know that as a set π(S 1 ) is countable. Therefore, p ◦ f˜ ≃p p ◦ g̃.
This implies that f ≃p g, and therefore [f ] = [g].
Theorem 6. π1 (S 1 , 1) ∼
= (Z, +).
Lecture 3: Retractions
di 15 okt 10:30
Example. Let X be the figure 8 space, and denote the right circle by A.
Then it’s easy to see that there exists a retract from the whole space to A
by mapping the left circle onto the right.
Proof. Look at the proof of the first lecture. Now that we’ve actually
proven that π(S1 ) = Z and π(B2 ) = 1, the proof is complete.
Proof. We have to show that for all f : I → X with f (0) = f (1) = x0 that
h ◦ f ≃p k ◦ f , i.e. h∗ [f ] = k∗ [f ].
H
G:I ×I X ×I Y
.
(s, t) (f (s), t) H(f (s), t)
• Then G is continuous.
– H(x, 0) = x
– H(x, 1) = r(x)
– H(a, t) = a for all a ∈ A
This means that 1X is homotopic to i◦r via a homotopy leaving A invariant.
x′
S1 R2 \ {(0, 0)}
Example. Consider the torus and a circle on the torus. Then it is a retract,
but not a deformation retract.
Remark. This means that the fundamental group of R20 is the same as the
one of S 1 , which is Z.
Example. The fundamental group of the figure 8 space and the θ-space are
isomorphic. These spaces are not deformations of each other, but we can
show that they are deformation retracts of R2 \ {p, q}. We say that these
spaces are of the same homotopy type.
R2 \ {p, q} R2 \ {p, q}
Figure 9.12: The figure eight and theta space seen as a deformation retract
of R2 \ {p, q}
Definition 10. Let X, Y be two spaces, then X and Y are said to be of the
same homotopy type iff there exists f : X → Y and g : Y → X such that
g ◦ f ≃ 1X and f ◦ g ≃ 1Y . We say that f, g are homotopy equivalences
and are homotopy inverses of each other.
π(Y, y0 ) ∋ [g]
h∗
π(X, x0 ) α̂
k∗
I ×I X ×I Y
F (s, t) = (f (s), t) H
β1
(x0 , 1) f1 y1
γ1 γ2 α
c
f0 y0
β0
(x0 , 0)
f∗,x0 g∗,x0
π(X, x0 ) π(Y, y0 ) π(X, x1 )
α̂
1π(X,x0 ) =(1X )∗
π(X, x0 )
.
g∗,x0 f∗,x1
π(Y, y0 ) π(X, x1 ) π(Y, y1 )
β̂
1π(Y,y0 ) =(1Y )∗
π(Y, x0 )
Then [f ] = [f1 ] ∗ [f2 ] ∗ · · · ∗ [fn ]. Note that all fi have images inside U or
V . Now,
[f ] = [a0 ] ∗ [f1 ] ∗ [α1 ] ∗ [α1 ] ∗ [f2 ] ∗ [α2 ] ∗ [α2 ] ∗ [f3 ] ∗ · · · ∗ [αn−1 ] ∗ [fn ] ∗ [αn ]
= [α0 ∗ (f1 ∗ α1 )] ∗ [α1 ∗ (f2 ∗ α2 )] ∗ · · · .
Every path of the form αi−1 ∗ (fi ∗ αi ) is a loop based at x0 lying entirely
inside U or V . This means that
f2 V
U f (a2 ) f3
α2
α1
f1
α0
α3
f (a3 )
f4
2
Example. RP2 = S∼ Then p : S 2 → RP2 , which is by definition continuous
by definition of the topology on the projective plane.
p a a
e1 e0
a ≃p a
This means that (S, p) is a covering of the projective plane. The lifting
correspondence says that
p
˜
b̃
ã ˜
ã
e0 b̃
b a
b0
i j ′
X FX X FX
j i
∃ϕ ∃ψ .
′
FX FX
Then
i
X FX
i .
ψ◦ϕ
FX
Then by uniqueness, ψ ◦ ϕ is 1FX , and likewise for ϕ ◦ ψ.
Note. The free group on a set always exists. You can construct it using
“irreducible words"
27
Lecture 4: Seifert-Van Kampen theorem
This is not a irreducible word. The reduced form is aaba−1 bba = a2 ba−1 b2 a.
Then FX is the set of irreducible words.
Example. If X = ∅, then FX = 1.
commute
π(U, x0 )
i1 Φ1
j1
i Φ
π(U ∩ V, x0 ) π(x, x0 ) H .
i2 Φ2
j2
π(V, x0 )
Proof. Not covered in this class. You can have a look in the book, but not
insightful.
G1
f1
f
G1 ∗ G2 H .
f2
G2
The groups U and V generate the whole group. The rest of this theorem follows from
the previous theorem.
π(U, x0 )
i1 Φ1
j1
i Φ
π(U ∩ V, x0 ) π(X, x0 ) H .
i2 Φ2 k
j2
π(V, x0 )
u v
a
deformation
U Z
b
deformation Z
V
deformation p 1
U ∩V
Conclusion: π1 (X, p) ∼
= Z ∗ Z = F{[a],[b]}
p b p
torus without inside
U a b
V = T 2 \ {q} = π(V ) = F2
a q a U = disc π(U ) = 1
x0 U ∩V = π1 (U ∩ V, x0 ) = Z
γ1
p b p
0 ) ∼ π(V,p)
Then π(X, x0 ) = π(V,x
N = γ̂(N ) . With N the normal subgroup generated
by the image of i2 : π(U ∩ V, x0 ) → π(T 2 \ {q}, x0 ).
i2 : π(U ∩ V, x0 ) ∼
= Z −→ π(T 2 \ {q}, 0)
⟨[f1 ] ∗ [f2 ] ∗ [f3 ] ∗ [f4 ]⟩ 7−→ ⟨[f1 ] ∗ [f2 ] ∗ [f3 ] ∗ [f4 ]⟩ .
p b p
γ2
γ3
f2
a f1 q f3 a
f4 γ4
γ1
p b p
γ̂([f1 ] ∗ [f2 ] ∗ [f3 ] ∗ [f4 ]) = [γ1 ] ∗ [f1 ] ∗ [γ2 ] ∗ [γ2 ] ∗ [f2 ] ∗ [γ3 ]
| {z } | {z }
[a] [b]
Therefore,
π(V, p)
π(X, p) =
⟨⟨[a][b][a−1 ][b−1 ]⟩⟩
F{[a],[b]}
=
⟨⟨[a] ∗ [b] ∗ [a]−1 ∗ [b]−1 ⟩⟩
= ⟨α, β | αβ = βα⟩ .
b b c
a a a d
U
b q
d b d
a c
b c
c c c
a
d b a a d
Same idea. (Note, we’ve renamed the edges in the figure on the right.)
V = X \ {a}. π1 (V, x0 ) = F{[a],[b],[c],[d]} , π(U ∩ V, x0 ) = Z. Conclusion
π(X, x0 ) = ⟨α, β, γ, δ | [α, β][γ, δ] = 1⟩
In this way, we can calculate the fundamental group of any surface,
e.g. projective space (Z2 ), klein bottle ( α, β | αβα−1 β = 1 , ‘just read
the boundary’), . . .
End of Chapter 11.
Classification of covering
spaces
(E, e0 )
f˜
p .
f
(Y, y0 ) (B, b0 )
Example. Take Y = [0, 1]. Then f is a path, then we showed that every
map can be lifted. And indeed, the condition holds: f∗ (π(Y, y0 )) = 1, the
trivial group, which is a subgroup of all groups.
34
Lecture 4: Seifert-Van Kampen theorem
f˜(y)
e0 E
f˜
p
α
I
f
y b0 f (y)
y0
B
e0 γ e1
E
˜
δ =f ◦β
f˜
p
f
y b0 f ◦ α f (y)
y0 α
f ◦β B
β
β
(E, e0 )
f˜
p .
f
(Y, y0 ) (B, b0 )
E
N
V
e0
f˜(y1 )
f˜
U
W
y y1 f b0 f (y1 ) f (y)
y0 α β
(E, e0 )
f˜
p
f
(Y, y0 ) (B, b0 )
B
′
is commutative. p ◦ h = p.
h
(E, e0 ) (E ′ , e′0 )
p
p′
(B, b0 )
(E, e0 ) (E ′ , e′0 )
l◦k k◦l
p p′
′
p p
(E, e0 ) (B, b0 ) (E ′ , e′0 ) (B, b0 )
Note that this doesn’t answer the question ‘is there a equivalence between
two coverings’, it only answers the question ‘is there an equivalence between two
coverings mapping e0 → e′0 ’. So now, we seek to understand the dependence of
H0 on the base point.
This completely answers the question: ‘When are two covering spaces equiv-
alent’:
Remark. Any two universal coverings are equivalent. Even more, we can
choose any base point we want.
h(e0 )=e′0
(E, e0 ) (E ′ , e′0 )
p
p′
(B, b0 )
Remark. We don’t cover Lemma 75.1, and we only cover point (a) from
75.2
X
q
p Y
r
Z
If p and r are covering maps, then also q is a covering map. (Also: if q and
p are covering maps, then so is r. Not the case for q, r ⇒ p!)
S
Consider a Uα . Then q(Uα ) is connected and contained in β∈J Vβ ,
but all these Vβ are disjoint, so there is exactly one Vβ such that
q(Uα ) ⊂ Vβ .
Now, let I ′ = {α | q(Uα ) ⊂ V }. For any α ∈ I ′ , we have the diagram
Uα
q
p V
r
U
Why is this useful? Because now we can say why the universal covering
space is a universal covering space.
E
q
p X
r
B
X
q
r
p
E B
of B. We define
p
E B
and, h1 and h2 are both lifts of p and there is a unique lift when fixing
the base point, so h1 and h2 agree.
Goal: what is the structure of the group C(E, p, B) in terms of fundamental
groups? Let (E, p) be a covering of B. p(e0 ) = b0 , H0 = p∗ π(E, e0 ). Remember:
Nπ(B,b0 ) (H0 )
Lemma 14. Im ψ = Φ , where
H0
e1
γ
e0
δ e2 e3
Then
h
(E, e1 ) (E, e2 )
p p
B
This h exists iff H1 = H2 . H1 = p∗ π(E, e1 ) and H2 = p∗ π(E, e2 ). If H0 is
normal. Then H1 and H2 are the same as H0 , because they are conjugate to
H0 . So for H1 , H2 the h exists. Conversely: exercise.
In this case, (Φ−1 ◦ ψ) : C(E, p, B) → π(B,bH0
0)
is an isomorphism. Special
case: Let (E, p) be the universal covering space, so π(E, e0 ) = 1, or H0 = 1. In
this case, Φ−1 ◦ ψ : C(E, p, B) → π(B, b0 ) is an isomorphism.
Example. C(R, p, S 1 ) ∼
=Z∼
= π(S 1 , b0 ).
p
Example. Consider S 2 − → RP 2 . C(S 2 , p, RP 2 ) = {1S 2 , −1S 2 }, because
2
take a point e0 ∈ S , then under a covering transformations, it can only be
mapped to e0 or to −e0 . So these are the only covering transformations.
And indeed, π(RP 2 ) = Z2
2
Note that RP 2 ∼= S∼ , where ∼ is in a sense defined by the groupactions
of 1S 2 and −1S 2 . More on this next week.
Lemma 15 (75.4). If B admits a universal covering space (E, p), then any
element b of B has a neighbourhood U such that
i8 : π1 (U, b) −→ π1 (B, b)
[α] 7−→ [1].
E
α̃
αU B
Lecture 6
di 12 nov 10:27
Recap: p : E → B and C(E, p, B) = {h : E → E | h homeo, p ◦ h = h}.
Nπ(B,b0 ) (H0 )
C(E, p, B) ∼
= H0 , where H0 = p∗ π(E, e0 ).
When p : E → B is regular (H0 is normal in π(B, b0 )), then
π(B, b0 )
C(E, p, B) ∼
= .
H0
In particular when E is simply connected,
C(E, p, B) ∼
= π1 (B, b0 ).
note
αk β ℓ (x, y) = ((−1)ℓ x + k, y + ℓ).
′ ′
Therefore αk β l = αk β l ⇔ k = k ′ ∧ l = l′ . As a set G = Z2 , but not
isomorphic.
Claim: G acts properly discontinuously on R2 . Indeed consider B(x, 41 ).
So
π1 (R2 /G) = G = α, β | βα = α−1 β .
Remark. These two spaces are the only possible spaces you can create using
affine actions. Projective plane? Not properly discontinuously?
Singular Homology
49
Lecture 6
x0 = (1, 0, 0, . . . , 0)
x1 = (0, 1, 0, . . . , 0)
xp = (0, 0, . . . , 0, 1).
X
σp = {(t0 , . . . , tp ) | ti > 0 and ti = 0}.
h : σp −→ p-simplex
X
(t0 , . . . , tp ) 7−→ t i xi .
ϕ : σp −→ X.
Definition 22. Let X be a topological space. Then Sn (X) is the free abelian
group on all singular n-simplices. This is a giant group. ByPthe definition
of a free abelian group, elements of Sn (X) are of the form nϕ ϕ, where
x2
ϕ ∂σ2
∂ 1 σ2 ∂0 σ2
− +
x0 σ2
+ ∂2 σ2
x1
Now, define
X X
∂ : Sn (X) → Sn−1 (X) : nϕ ϕ 7→ nϕ (∂ϕ),
where X
∂ϕ = (−1)i ∂i ϕ ∈ Sn−1 (X).
Then certainly, ∂ is a morphism of abelian groups. In the case of the figure, we
get
∂ϕ = ∂0 ϕ − ∂1 ϕ + ∂2 ϕ,
which really does goes around the simplex in the correct direction. The (−1)i
fixes the direction of the simplices.
Proposition 5 (1.3). ∂ ◦ ∂ = 0.
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
−
→ Sn+1 (X) −
→ Sn (X) −
→ Sn−1 (X) −
→ ··· −
→ S1 (X) −
→ S0 (X) −
→0−
→0−
→ ··· .
We claimed that Bn ⊂ Zn .
Zn (X)
Definition 23. The n-th singular homology group of X is Hn (X) = Bn (X) .
Two cycles are the same if they differ with a boundary.
A cycle is more or less a loop. Then two cycles are the same when they differ
with a boundary, in this case the boundary of the triangles indicated with the
dashed lines.
Lecture 7
di 19 nov 10:27
Consider the more general situation.
∂n+1 n ∂
C∗ : · · · →Cn+1 −−−→ Cn −→ Cn−1 → . . .
↓ ϕn+1 ↓ ϕn
′
∂n+1 n ∂′
D∗ : · · · →Dn+1 −−−→ Dn −→ Dn−1 → . . .
.
where ϕn ◦ ∂n = ∂n′ ◦ ϕn
Define Zn = Ker ∂n ≤ Cn and Bn = Im ∂n+1 ≤ Cn and Bn ≤ Zn .
Then we define
Zn (C∗ )
Hn (C∗ ) = .
Bn (C∗ )
Φ∗ : Hn (C∗ ) −→ Hn (D∗ )
z + Bn (C∗ ) 7−→ ϕn (z) + Bn (D∗ ).
We claim that this is well-defined. Indeed, first we check that ϕn (z) is a cycle.
n ∂ ϕn−1
Indeed, using the commutativity of the diagram, z −→ 0 −−−→ 0 implies that
′
∂n
ϕn (z) −→ 0. Also, a boundary stays a boundary by the same reasoning.
We apply this in connection with a continuous map f : X → Y . Consider
the following:
X Y
ϕ f
f# (ϕ) = f ◦ ϕ
∂ ∂
Sn+1 (X) ϕ ∈ Sn (X) ∂ϕ ∈ Sn−1
f# f# f#
∂ ∂
Sn+1 (Y ) f# (ϕ) ∈ Sn (Y ) Sn−1
Now, the question is, is ∂f# (ϕ) = f# (∂ϕ). It’s enough to check this for ∂i :
f∗ : Hn (X) −→ Hn (Y )
z + Bn (x) 7−→ f# (z) + Bn (Y ),
f∗ : Hn (X) −→ Hn (Y )
⟨z⟩ 7−→ ⟨f# (z)⟩ .
so this is a functor.
Example. What are the homology groups of the space consisting of one
point, X = {x0 }. Note that ∀n ∈ N, there is exactly one singular n-
simplex, ϕn : σn → X : (t0 , . . . , tn ) 7→ x0 . Sn is the free abelian group on
ϕn ∼
= Z for all n.
∂ : S0 (X) → S−1 (X) : ϕ0 7→ 0.
So all odd maps are 0 and if odd, then it maps the generator of Sn to
the generator of Sn−1 :
0 1Z 0 1Z 0 0 0
S2n+3 S2n+2 S2n+1 S2n S2n−1 S1 S0 0
Z Z Z Z Z Z Z 0
Now,
Z0 (X) Z ∼
H0 (X) = = =Z
B0 (X) ⟨0⟩
Z
H2n+1 (X) = ∼=0
Z
0
H2n (X) = ∼ = 0.
0
so this element,
P which we’ve assumed lied in the kernel is also a
boundary, i.e. ni xi ∈ B0 (X).
What happens if our space is not path-connected?
Suppose C∗ is a chain complex and there exists an I such that for all n,
M
Cn = Cni ,
i∈I
H0 (Xα ) ∼
L L
Proposition 6. H0 (X) = α = αZ
Proof. We will define a map going from Sn (X) → Sn+1 (X). Fix x ∈ X.
Take ϕ : σn → X, and we’ll build up a θ : σn+1 → X.
Now consider the purple point, we can write is as a convex combination
of the blue and the red point:
t1 t2 tp+1
t0 (1, 0, . . . , 0) +(1 − t0 ) 0, , ,..., .
| {z } 1 − t0 1 − t0 1 − t0
| {z }
Then we define:
t1 t2 tp+1
θ(t0 , t1 , . . . , tn+1 ) = t0 x + (1 − t0 )ϕ , ,...,
1 − t0 1 − t0 1 − t0
0 t →0
≤ |(1 − t0 )|(M + C) −− −→ 0.
θ x
v0
originally: v0
v2
ϕ
v1
originally: v0
This does not work if we’re working with a 0-simplex. (Because then
∂i−1 becomes zero, and the only homomorphisms that we have are the
identity, so . . . )
Now, consider the total boundary,
n
X
∂T ϕ = ∂0 T ϕ + (−1)i ∂i T ϕ
i=1
n
X
=ϕ+ (−1)i T ∂i−1 ϕ
i=1
n−1
X
=ϕ+ (−1)j+1 T ∂j ϕ
j=0
n−1
X
=ϕ+T (−1)j+1 ∂j ϕ
j=0
= ϕ − T (∂ϕ).
∂ ∂
Cn+1 Cn Cn−1
gn+1 fn+1 gn fn gn+1 fn−1
T T
∂ ∂
Dn+1 Dn Dn−1
If f − g = ∂T + T ∂ for some T for some T : Cn → Dn+1 for all n,
where f and g are chain maps. Then f∗ = g∗ : Hn (C∗ ) → Hn (D∗ ). When
f − g = ∂T + T ∂ is satisfied, we say that f and g are chain homotopic.
This indeed has something to do with homotopic maps.
If you have two maps that are homotopic, then f# and g# will be chain
homotopic maps.
There is a proof of the theorem in the book: long, not difficult. There
is a better proof, see problem 8.
Idea of the proof:
g∼f g
ϕ
Tϕ
P
≈ σ3
∂T ϕ T ∂ϕ
g# (ϕ) f# (ϕ)
This goes from two dimensional stuff to three dimensional stuff. But the
prism on the left is not a simplex. But we can divide it into simplices. Now, we
look at ∂T ϕ + T ∂ϕ, and we see that (being careful with signs) is g# (ϕ) − f# (ϕ).
Conclusion: If f ∼ g : X → Y , then for all n ≥ 0, we have that f∗ = g∗ :
Hn (X) → Hn (Y ).
Theorem 20. If X and Y have the same homotopy type, then they also
have isomorphic homology groups.
Proof. • The last bullet point is trivial. This follows from induced
maps from homotopic maps.
• Let r : X → A be a retraction. Then we have
i r
A X A
r◦i=1A
and
r∗ (γ − i∗ r∗ γ) = r∗ γ − r∗ i∗ r∗ γ
= r∗ γ − r∗ γ = 0.
Lecture 8: Ill
di 03 dec 23:02
Lecture 9
di 10 dec 10:31
Recap:
inj
0 → H1 (S 1 ) = Z −−→ H0 (U ∩ V ).
and ⟨c + d⟩ 7→ x − y.
x y V
U
Dn
(1S n )∗ =1Hn−1
Z = Hn−1 (S n−1 ) Hn−1 (S n−1 ) =Z
i∗ r∗
Hn−1 (Dn ) = 0
S n \ {p}
Then
f∗
Hn (S n ) Hn (S n )
f∗′
i∗
Hn (S n \ {p}) = 0
Proof. By induction on n.
Base case, n = 1. S 1 = {(x1 , x2 ) | x21 + x22 = 1}.
z
c
x y
d
z′
This is
−∆−1 (i∗ r∗ ∆(⟨α⟩)).
As i∗ and r∗ are inverses, we have that d(f ) = −1.
Suppose f : S n → S n . Then there is a map Σf : S n+1 → S n+1 (the
suspension of f ), given by
S n+1 Σf S n+1
f
Sn Sn
Σf
Σf
Using formulas:
Then
Σf : S n+1 −→ S n+1
(
(x, t) if x = 0, north and south pole .
(x, t) 7−→ x
∥x∥f ( ∥x∥ ), t if x ̸= 0
[−1, 1] × X
Proof. Let h : S n → S n be the map that exchanges the first coordinate and
the ith coordinate. Then h is a homeomorphism: h−1 = h, so h ◦ h = 1S n ,
so d(h) = ±1 Then fi = h ◦ f ◦ h. So
Proof.
d(A) = d(f1 )d(f2 ) · · · d(fn+1 ) = (−1)n+1 .
So for even dimensional spheres, d(A) = −1 and for odd dimensional sphere,
d(A) = 1. This is a big difference.
f (x)
Sn
Ag(x)
Proof. The line does not go through the center, as f (x) ̸= g(x). Define a
homotopy,
(1 − t)f (x) + tAg(x)
(x, t) 7→ .
∥(1 − t)f (x) + tAg(x)∥ =
̸ 0
This is a homotopy between f and Ag
Not in the text:
Corollary. If |d(f )| =
̸ |d(g)|, then there is a point x such that f (x) = g(x).
Proof. The previous theorem says that d(f ) = d(A)d(g), so |d(f )| = |d(g)|.
You can even make this better by making a distinction between odd and
even dimensional spheres.
Corollary. If |d(f )| =
̸ 1, then f has a fixed point, i.e. an x such that f (x) =
x = Id(x).
Proof. If |d(f )| =
̸ d(Id) = 1, then . . .
which is a contradiction.
Remark. This does not hold for odd dimensional spheres. For example,
consider S1 , and make a small rotation, then no point is mapped onto
itself or onto its antipodal point. Doing the same for S2 , we see that when
we do a small rotation, the north and south pole are switched.
Remark. It’s harder to show that for odd dimensional spheres, there always
exist such a map.
Now, we prove the hairy ball theorem, which only holds in even dimensional
spheres.
Proof. By the previous theorem, there is a point for which f (x) = ±x, and
x ̸⊥ ±x.
Proof. Suppose ϕ does exists. Then, map the tangent vector to a point on
the sphere:
ϕ(x)
ψ : S 2n → S 2n : x 7→ .
∥ϕ(x)∥
x
f (x)
Remark. For odd dimensional spheres, there always exists a such a vector
field. We even know how many linearly independent vector fields exist!
Torus
Goal: Hn (T 2 ).
a
V
c1
p
b c2 b
H2 (U ) ⊕ H2 (U ) → H2 (T 2 ) → H1 (U ∩ V ) → H1 (U ) ⊕ H1 (U ) → H1 (T 2 ) → H0 (U ∩ V ) → H0 (U ) ⊕ H0 (U ) → H0 (T 2 )
So we have
→ 0 ⊕ 0 → H2 (T 2 ) → Z → Z2 ⊕ 0 → H1 (T 2 ) → Z → Z ⊕ Z → Z → 0.
i r
U ∩V U figure 8
ir r∗
H1 (U ∩ V ) H1 (U ) H1 (figure 8)
⟨c1 + c2 + c3 + c4 ⟩ ⟨c1 + c2 + c3 + c4 ⟩ ⟨r ◦ c1 + r ◦ c2 + r ◦ c3 + r ◦ c4 ⟩
a+b+a+b
Remark. H2 (S 1 × S 1 ) ̸= H2 (S 1 ) × H2 (S 1 ).
n
Remark. Hn (T k ) = Z(k ) , where n
k = 0 when k > n.
Klein bottle
a
V
c1
p
b c2 b
• V = B 2 So H0 = Z and Hn = 0 for n ≥ 1.
• U ∩ V = B 2 \ {p}, which retracts to a circle. So H0 = Z and H1 = Z =
grp(⟨c1 + c2 + c3 + c4 ⟩)
g∗
→ H2 (U ) ⊕ H2 (V ) → H2 (K) → H1 (U ∩ V ) −→ H1 (U ) ⊕ H1 (V ) → H1 (K) → H0 (U ∩ V ) → H0 (U ) ⊕ H0 (V ) → H0 (K) → 0
Klein bottle is path connected, so
g∗
→ 0 ⊕ 0 → H2 (K) → Z −→ Z2 ⊕ 0 → H1 (K) → Z → Z ⊕ Z → Z → 0
We can cut of the sequence because all spaces are path connected.
g∗
→ 0 ⊕ 0 → H2 (K) → Z −→ Z2 ⊕ 0 → H1 (K) → 0
Now, we have to understand what g∗ is
Then
i r
U ∩V U figure 8
ir r∗
H1 (U ∩ V ) H1 (U ) H1 (figure 8)
⟨c1 + c2 + c3 + c4 ⟩ ⟨c1 + c2 + c3 + c4 ⟩ ⟨r ◦ c1 + r ◦ c2 + r ◦ c3 + r ◦ c4 ⟩
a+b+a+b
So we get
− ⟨a⟩ − ⟨b⟩ + ⟨a⟩ − ⟨b⟩ = −2 ⟨b⟩ .
So g∗ : Z → Z2 : 1 7→ (0, −2), or z → (0, −2z). So, the image of g∗ =
0 × 2Z ≤ Z × Z. The kernel of g∗ is {0}. So g∗ is injective. And 0 × 2Z is the
kernel of h∗
g∗ h
→ 0 ⊕ 0 → H2 (K) → Z −→ Z2 −→
∗
H1 (K) → 0
Now, the zero on the right implies that h∗ is surjective. so (H1 (U ) ⊕ 0)/ Ker h∗ ,
Z2
which is 0×2Z , which is Z ⊕ Z2 .
Now, on the left. As g∗ is injective, so H2 (K) → H1 (U ∩ V ) = Z is the zero
map. On the other hand, we have zeros on the left, so H2 (K) → H1 (U ∩ V ) = Z
is also injective. The only way the zero map can be injective is when the spaces
are zero. Therefore, H2 (K) = 0.
Higher dimensional: trivial.
Let n ≥ 3
0 ⊕ 0 = Hn (U ) ⊕ Hn (V ) → Hn (K) → Hn−1 (U ∩ V ) = 0,
π1 (X)
Remark. Fact: The H1 (X) = [π1 (X),π 1 (X)]
. Abelianized version of funda-
mental group. Remember, we did compute the fundamental group of the
klein bottle! π1 (K) = α, β | βα = α−1 β . Making this group commuta-
tive, we have
H1 (K) = α, β | βα = α−1 β, αβ = βα .
H1 (K) = α, β | αβ = βα, α2 = 1 = Z2 ⊕ Z.
• H1 = Z2 (abelianized version)
• H2 = 0 (orientable)
• Hn≥3 = 0
Proof. Picture:
c f A◦γ
f (a)
f ◦c
A(a) = b a
d γ
A∗ (⟨α⟩) = A∗ ⟨c + d⟩ = ⟨A ◦ c + A ◦ d⟩ = ⟨d + c⟩ = ⟨α⟩ .
So
m ⟨α⟩ = A∗ ⟨f ◦ c + γ⟩ = ⟨A ◦ f ◦ c + A ◦ γ⟩ .
Now, A ◦ γ is the path from f (a) to f (b), going counter clockwise. Now,
we assume A ◦ f = f ◦ A, so
m ⟨α⟩ = ⟨f ◦ d + A ◦ γ⟩ .
Now, also
2m ⟨a⟩ = ⟨f ◦ c + γ⟩ + ⟨f ◦ d + A ◦ γ⟩
= ⟨f ◦ c + f ◦ d + γ + A ◦ γ⟩
= ⟨f ◦ c + f ◦ d⟩ + ⟨γ + A ◦ γ⟩
= f∗ ⟨c + d⟩ + ⟨α⟩
= f∗ ⟨α⟩ + ⟨α⟩ .
So
f∗ ⟨α⟩ = (2m − 1) ⟨α⟩ ,
for some integer m.
TODO: In
problem:
Remark. Exercise: even maps have even degrees: f (Ax) = f (x). say that one
cycles
S2 S1
f
S1
f
S2 S1
i
f |S 1
S1
Then we have
f∗
H1 (S 2 ) = 0 H1 (S 1 ) = Z
i∗
·(2k+1)
1
H1 (S ) = Z
f (x) − f (−x)
g : S 2 → R2 : x 7→ ∈ S1.
∥f (x) − f (−x)∥
b1 S2
v
h
b2
Now, consider
Then f1 (v) is the volume of bread b1 above Pv . Then f2 (v) is the volume
of bread b2 above Pv .
Now, you should believe that f1 and f2 are continuous. (Proving this
precisely needs measure theory etc.) So, now, we can use the Borsak
Ulam theorem. So there exists a v ∈ S 2 such that f (v) = f (−v). So
f1 (v) = f1 (−v), so volume of bread b1 above Pv is the volume of bread b1
below Pv , and similar for f2 . This proves the Ham sandwich theorem.
Last result.
Remark. This also works for 2 (Just take A3 = ∅), but this doesn’t work
for 4.
A4
A2
A3
A1
Remark. Time on schedule is when you have to enter the room. Hand over
problems after exam.