Topbookf
Topbookf
UNIVERSITY OF WISCONSIN-MADISON
A L G E B R A I C TO P O L O G Y:
A COMPREHENSIVE
INTRODUCTION
i
Contents
Preface ix
2 Fundamental group 3
2.1 Definition 3
2.2 Basepoint (in)dependence 7
2.3 Functoriality 8
2.4 Homotopy invariance of fundamental group 9
2.5 Contractible spaces. Deformation Retracts 10
2.6 Fundamental group of a circle 12
2.7 Some Immediate Applications 15
Brower’s fixed point theorem 16
Fundamental Theorem of Algebra 17
Exercises 18
2.8 Seifert–Van Kampen’s Theorem 19
Free Groups 19
Free Products 21
Seifert-Van Kampen Theorem 25
Exercises 29
4 Covering spaces 47
4.1 Definition. Properties 47
4.2 Covering transformations 52
4.3 Universal Covering Spaces 54
4.4 Group actions and covering maps 56
Exercises 60
5 Homology 63
5.1 Singular Homology 63
5.2 Homotopy Invariance 66
5.3 Homology of a pair 70
Exercises 79
5.4 π1 vs. H1 81
5.5 Cellular Homology 83
Degrees 84
Exercises 87
How to Compute Degrees? 88
CW Complexes 90
Exercises 92
Cellular Homology 92
Exercises 99
5.6 Euler Characteristic 101
Exercises 102
5.7 Lefschetz Fixed Point Theorem 103
Exercises 105
iii
Bibliography 311
vii
List of Figures
Preface
1
Foreword about invariants of spaces
Theorem 1.0.2. Any continuous map f : [0, 1] → [0, 1] has a fixed point, i.e.,
there exists x ∈ [0, 1] so that f ( x ) = x.
2
Fundamental group
2.1 Definition
(t, 0) 7→ γ(t)
(t, 1) 7→ δ(t)
(0, s) 7→ x
(1, s) 7→ y
s
δ
γs
x y
F
t
γ
F F̄
• symmetric: if γ ∼ δ, then δ ∼ γ via F̄ (t, s) = F (t, 1 − s).
F G H
• transitive: let γ ∼ δ and δ ∼ φ, then γ ∼ φ via
1
F (t, 2s) 0≤s≤ 2
H (t, s) =
G (t, 2s − 1) 1
2 ≤s≤1
Note that
0 γ 1 δ 1
2
0 γ s δ 1
1
γ′ 2 δ′
F G
γ 1 δ
2
6 algebraic topology
[γ] · [δ] := [γ ∗ δ]
0 γ s′ δ 1
γ δ
0 s 1
(γ ∗ δ) ∗ η ∼ (γ ∗ δ) ∗ 2 η ∼ γ ∗ 1 (δ ∗ η ) ∼ γ ∗ (δ ∗ η )
3 3
Indeed, we have,
2t s +1
γ(
s +1 ) 0≤t≤ 2
G (t, s) =
s +1
y = γ (1)
2 ≤t≤1
x y
γ ey
γ ey
F
And similarly for ex ∗ γ ∼ γ. Therefore, ex is the identity element in
(π1 ( X, x ), ·).
fundamental group 7
Finally, let
γ̄(t) = γ(1 − t)
and set,
[γ]−1 := [γ̄]
We claim that, γ ∗ γ̄ ∼ ex ∼ γ̄ ∗ γ, i.e., [γ̄] is the inverse of [γ] in
(π1 ( X, x ), ·). Indeed, γ ∗ γ̄ ∼ ex via:
ex
γ (1 − s )
s
γs γ̄s
γ x γ̄
γ 7→ δ̄ ∗ γ ∗ δ
2.3 Functoriality
f ∗ : π1 ( X, x ) → π1 (Y, y)
[γ] 7→ [ f ◦ γ]
f g
1. If ( X, x ) → (Y, y) → ( Z, z), then ( g ◦ f )∗ = g∗ ◦ f ∗
f : ( X, x ) → (Y, y)
is a homeomorphism, then
f ∗ : π1 ( X, x ) → π1 (Y, y)
is an isomorphism.
f ∗ = g∗ : π1 ( X, x0 ) → (Y, y0 ).
h∗
π1 ( X, x0 ) π1 (Y, y0 )
α#
k∗
π1 (Y, y1 )
f g f
( X, x0 ) −
→ (Y, y0 ) −
→ ( X, x1 ) −
→ (Y, y1 )
f ∼ f ∗ ey ∼ f ∗ ḡ ∗ g ∼ ex ∗ g ∼ g,
p
fundamental group 13
ϕ(m + n) = [ p ◦ ω
e n+m ] = [ p ◦ (ω
e m ∗ τm (ω
e n ))]
= [( p ◦ ω
e m ) ∗ ( p ◦ τm (ω
e n )]
= [ ωm ∗ ωn ] = [ ωm ] · [ ωn ]
= ϕ ( m ) · ϕ ( n ),
hence ϕ is a group homomorphism.
To prove that ϕ is bijective, we need two lemmas.
(R, x̃0 )
∃! f˜
p
f
( I, 0) ( S1 , x0 )
Assuming the two lemmas for now, let x0 = (1, 0) and choose x̃0 = 0.
Let f : I → S1 be a loop at (1, 0) representing [ f ] ∈ π1 (S1 , x0 ). By the
path lifting lemma, there is a path f˜ : I → R such that p ◦ f˜ = f and
f˜(0) = 0 ∈ Z. Say f˜(1) = n ∈ Z, so f˜ is a path in R from 0 to n. Then
ϕ(n) = [ p ◦ f˜] = [ f ]. Since f was arbitrary, ϕ must be surjective.
Now suppose ϕ(m) = ϕ(n) for some m, n ∈ Z. So [ωm ] = [ωn ],
or ωm ∼ ωn . Let f s be a homotopy with f 0 = ωm and f 1 = ωn .
By homotopy lifting, there exists a homotopy f˜s : I → R such that
p ◦ f˜s = f s and f˜s (0) = 0. But f˜s (1) is independent of s, so f˜0 (1) =
f˜1 (1). Now f˜0 and ω e m are both lifts to R of f 0 = ωm which start at
0. By the uniqueness of path lifting, this gives f˜0 = ω̃m . In particular,
f˜0 (1) = ω
e m (1) = m. Similarly, f˜1 (1) = ωe n (1) = n. So m = n.
The path and homotopy lifting Lemmas 2.6.2 and 2.6.3 are conse-
quences of the following general lifting lemma which we prove here.
R
Fe p
∃! Fe
F
Y × {0} Y×I S1
(Step 1) There is an open cover {Uα }α of S1 so that for each α, one has
p−1 (Uα ) = β U e β is an open interval in R that satisfies
F
e β , where each U
p (U e β ) = Uα and such that p restricts to a homeomorphism between U eβ
and Uα . For all pairs (y0 , s) ∈ Y × I, let α be such that F (y0 , s) ∈ Uα .
Since F is continuous, there is a neighborhood Ns × ( as , bs ) of (y0 , s)
so that F ( Ns × ( as , bs )) ⊆ Uα . Since {y0 } × I is compact, it can be
covered by finitely many such Ns × ( as , bs ). We can choose a single
neighborhood N of y0 and a partition of I given by 0 = s0 < s1 <
· · · < sm = 1 so that for each i there is an αi with F ( N × [si , si+1 ]) ⊂
Uαi . Assume (for induction) that Fe has been defined on N × [0, si ],
starting with the given lift on N × {0} for i = 0. We can extend it
to N × [si , si+1 ] as follows. Recall that since F ( N × [si , si+1 ]) ⊂ Uαi ,
we have Fe( N × {si }) ⊂ U e β , for a unique β i as above. Define Fe on
i
− 1
N × [si , si+1 ] by F = ( p |Uα : Uαi → U
e e β ) ◦ F.
i i
(Step 2) Now we show uniqueness for the case when Y is a single point.
Choose a partition of I by 0 = s0 < s1 < · · · < sm = 1 so that for all
i there is an open Uαi that completely contains F ([si , si+1 ]). Assume
we have Fe and Fe′ , two lifts of F : I → S1 . We have that Fe(0) = Fe′ (0),
since we are choosing a specific starting point x̃0 ∈ R. For induction,
suppose Fe and Fe′ coincide on [0, si ]. Since Fe is continuous and [si , si+1 ]
is connected, we have that Fe([si , si+1 ]) is connected. Thus there is a
unique U e β that completely contains Fe([si , si+1 ]). Similarly, there is a
i
unique Uβi′ ⊃ Fe′ ([si , si+1 ]).
e
⊂R
F
i Uβ i
e
Fe p
I ⊃ [ s i , s i +1 ] Uαi ⊂ S1
F
Since Fe(si ) = Fe′ (si ) by the induction hypothesis, and given that the sets
{Ue β } are either disjoint or equal, we must have that U
i
eβ = U
i
e β . Also,
i′
p| e is a homeomorphism, so p is injective on U e β and p ◦ Fe = p ◦ Fe′ .
Uβ i
i
Hence Fe = Fe′ on [si , si+1 ].
fundamental group 15
γ−1 (y) is compact, only finitely many intervals ( ai , bi ) cover γ−1 (y).
Let ( a j , b j ), j ∈ A be so that ( a j , b j ) ∩ γ−1 (y) ̸= ∅. Let γ j := γ|[ a j ,bj ] ⊂ B̄y .
So γ( a j ), γ(b j ) ∈ ∂ B̄y = Syn−1 . As Syn−1 is path connected, there is a path
δj in Syn−1 from γ( a j ) to γ(b j ). Since B̄y is contractible, we then obtain
16 algebraic topology
that δj ∼ γ j in B̄y . Note that y ̸∈ Im δj . Homotop γ by deforming
γ j to δj , and keeping the rest of γ unchanged. Repeat the process
for all j’s such that ( a j , b j ) ∩ γ−1 (y) ̸= ∅. We get a loop η ∼ γ with
Im(η ) ∩ {y} = ∅.
R2 \{0} ∼
= Rn \{ f (0)}
hence
π1 (R2 \{0}) ∼
= π1 (Rn \{ f (0)})
But we know that
(
Z, n=2
π1 (R \{ a}) ∼
n
= π 1 ( S n −1 ) =
0, n > 2.
r(x)
D2
f (x)
B = S2 ∩ {( x1 , x2 , x3 ) ∈ R3 | x1 , x2 , x3 ≥ 0} ∼
= D2 .
( pn )∗
Z∼
= π1 (Cr , r ) Z∼
= π 1 (C∗ , r n )
δ#
(e an )∗
Z∼
= π 1 (C∗ , a n )
Exercises
1. Show that if h, h′ : X → Y are homotopic and k, k′ : Y → Z are
homotopic, then k ◦ h and k′ ◦ h′ are homotopic.
6. Using the fact that the fundamental group of the circle S1 is Z, show
that there are no retractions r : X → A in the following cases:
fundamental group 19
9. Let A be a real 3 × 3 matrix, with all entries positive. Show that A has
a positive real eigenvalue. (Hint: Use Brower’s fixed point theorem.)
Free Groups
Definition 2.8.1. Let G be a group, and let { x j } j∈ J be a set of elements of
G. We say that the set { x j } j∈ J generates the group G if every element of G
can be written as a product of powers of the elements of { x j } j∈ J . If the family
{ x j } j∈ J is finite, we say that G is finitely generated.
w1 ∼ w1′
w2 ∼ w2′
then,
w1 w2 ∼ w1′ w2′ .
Thus, the binary law on W ( X ) given by concatenation descends to
F ( X ). Moreover, we have:
Theorem 2.8.7. The set F ( X ) of equivalence classes of words, with the induced
binary operation, is a group called the free group on the set X.
i : X −→ F ( X ), x 7→ [ x ]
be the map that sends every element of X to the equivalence class of the word
it defines, and let
j : X −→ G
be a set map from X to a group G. Then, there is a unique group homomorphism
f : F ( X ) −→ G
such that f ◦ i = j.
fundamental group 21
F ( X ) = { x n | n ∈ Z} ∼
= Z.
G = ⟨ a | a n = 1⟩.
Then,
j : X → G, x 7→ a
gives a homomorphism
f : F ( X ) −→ G
G∼
= F ( X )/ ker( f ).
f : F ( X ) −→ G.
Therefore,
G∼
= F ( X )/ ker( f ) = ⟨ x ∈ X | r ∈ ker( f )⟩
Free Products
Let H and K be groups. We form a new group H ∗ K from them, which
is called the free product of H and K, defined as follows. First we
consider the following set of words:
W ( H, K ) = { g1 g2 . . . gn | gi ∈ H or gi ∈ K }.
w = w1 ∼ e . . . ∼ e w k = w ′ .
h1 k 1 h2 k 2 . . . hr k r
with:
hi ∈ H for all i = 1, . . . , r
k ∈ K
for all j = 1, . . . , r
j
i ̸= 1
h for all i = 2, . . . , r
k ̸= 1 for all j = 1, . . . , r − 1.
j
Theorem 2.8.14. The set H ∗ K, endowed with the operation induced from
concatenation, is a group, called the free product of H and K.
i : H −→ H ∗ K, h 7→ [h]
and
j : K −→ H ∗ K, k 7→ [k ]
be the maps that send every element of H (respectively, every element of K) to
the equivalence class of the word it defines, and let
p : H −→ G
fundamental group 23
and
q : K −→ G
be any pair of group homomorphisms. Then, there is a unique group homomor-
phism
f : H ∗ K −→ G
such that f ◦ i = p and f ◦ j = q, or equivalently, such that the following
diagram is commutative.
p
i
∃! f
H∗K G
j
q
f ( h1 k 1 h2 k 2 . . . hr k r ) = p ( h1 ) · q ( k 1 ) · p ( h2 ) · q ( k 2 ) · . . . · p ( hr ) · q ( k r ) ∈ G
Fi := F ( xi ) ∼
=Z
F(X ) ∼
= F1 ∗ . . . ∗ Fn ∼
= Z ∗ . . . ∗ Z = : Z∗ n
Example 2.8.19. If
H = ⟨h | rh ⟩
K = ⟨k | rk ⟩
are presentations of H and K by generators and relations, then
H ∗ K := ⟨h, k | rh , rk ⟩.
ω : Z2 ∗ Z2 −→ Z2 , x 7→ length of x mod 2.
ker(ω ) = ⟨ ab⟩ ∼
=Z
We define the action ϕ of Z2 on Z = ⟨ ab⟩ by
( Hα \{1}) ∩ ( Hβ \{1}) = ∅.
For a free product of an arbitrary number of groups we also have
the Universal Mapping Property, namely:
Proposition 2.8.22 (UMP). Let { φα : Hα −→ G }α∈ A be a collection of
group homomorphisms, and let iα : Hα −→ α∈∗ A Hα be the inclusion for all
α ∈ A. Then, there exists a unique group homomorphism
∗
φ: α∈ A Hα −→ G
such that, for all α ∈ A,
φ ◦ iα = φα .
Sketch of proof. Let h1 h2 . . . hn be a word in α∈∗A Hα , with hi ∈ Hαi for all
i = 1, . . . , n. Define the map φ as:
φ ( h 1 h 2 . . . h n ) = φ α1 ( h 1 ) · φ α2 ( h 2 ) · . . . · φ α n ( h n ) ∈ G
φ β : Hβ −→α∈×A Hα
be the inclusion for all β ∈ A. Then it follows from the UMP that there
exists a unique homomorphism
∗
φ: α∈ A Hα −→α∈×A Hα
that preserves every subgroup Hα .
fundamental group 25
and
[
X= Aα .
α∈ J
The inclusion
jα : Aα ,→ X
induces a homomorphism
( jα )∗ : π1 ( Aα , x0 ) −→ π1 ( X, x0 )
for all α ∈ J, so, by the UMP, there exists a unique homomorphism
∗
φ: α ∈ J π1 ( A α , x0 ) −→ π1 ( X, x0 ).
For every α, β ∈ J, with α ̸= β, we denote by iαβ the inclusion
iαβ : Aα ∩ A β ,→ Aα .
iαβ Aα jα
Aα ∩ A β X
i βα jβ
Aβ
(iαβ )∗ π1 ( A α , x0 )
( jα )∗
π1 ( A α ∩ A β , x0 ) π1 ( X, x0 )
(i βα )∗ ( jβ )∗
π1 ( A β , x0 )
S
Theorem 2.8.24 (Seifert-Van Kampen). If X = Aα is a topological
α∈ J
space, where Aα is a path-connected open set such that x0 ∈ Aα for all α ∈ J,
then
is surjective.
Proof.
1) Let f : I −→ X be a loop at x0 ∈ X. By the continuity of f and the
compactness of I, there exists a partition 0 = s0 < s1 < . . . < sm = 1
such that
f ([si−1 , si ]) ⊂ Aαi
for some αi ∈ J. Denote by Ai the set Aαi , and let f i := f |[si−1 ,si ] . We
have that
f = f1 ∗ f2 ∗ . . . ∗ fm
with f i a path in Ai .
The set Ai ∩ Ai+1 is path-connected, and { x0 , f (si )} ⊂ Ai ∩ Ai+1 .
Thus, there exists a path gi in Ai ∩ Ai+1 from x0 to f (si ) for all i =
1, . . . , m − 1. Therefore,
f ∼ ( f 1 ∗ g 1 ) ∗ ( g1 ∗ f 2 ∗ g 2 ) ∗ . . . ∗ ( g m − 1 ∗ f m ) ,
[ f ] = [ f 1 ∗ g 1 ] · [ g1 ∗ f 2 ∗ g 2 ] · . . . · [ g m − 1 ∗ f m ] ,
where
• f 1 ∗ g1 is contained in A1 .
• gm−1 ∗ f m is contained in Am .
∗
Thus, if we see the classes of these loops as letters in α ∈ J π1 ( A α , x0 ), we
get that
g1 f ( s1 )
f1
Aα
x0 f2
Aβ
2) Clearly, (iαβ )∗ (ξ )((i βα )∗ (ξ ))−1 ∈ ker( φ) for all α, β ∈ J and for all
ξ ∈ π1 ( Aα ∩ A β , x0 ), so
φ( gN ) = φ( g) = ( jα )∗ ( g),
so by applying k we get
as desired. (More details can be found Hatcher’s book, see also the
book of Munkres for the case | J | = 2.)
i U u
U∩V U∪V
j V v
Then
π1 (U, x0 ) ∗ π1 (V, x0 )
π1 ( X, x0 ) ∼
= .
N ⟨i∗ (ξ ) j∗ (ξ )−1 | ξ ∈ π1 (U ∩ V, x0 )⟩
Corollary 2.8.26. If U ∩ V is simply connected, then π1 (U ∪ V, x0 ) =
π1 (U, x0 ) ∗ π1 (V, x0 ).
Corollary 2.8.27. The union of two simply connected spaces is simply con-
nected, provided their intersection is nonempty and path connected.
Exercises
1. Let X be the space obtained from D2 by identifying two distinct
points on its boundary. Is there a retract from X to its boundary?
Explain.
(v) Möbius band. Are the cylinder and the Möbius band homeomor-
phic?
(vi) The complement in R3 of a line and a circle. Note: There are two
cases to consider, one where the line goes through the interior of the
circle and the other where it doesn’t. Are these two spaces homotopy
equivalent?
3. Show that RP3 and RP2 ∨ S3 have the same fundamental group.
Are they homeomorphic?
f1 f2 f3
X1 → X2 → X3 → · · ·
30 algebraic topology
obtained from the disjoint union of cylinders Xi × [0, 1] via the identifi-
cation of ( xi , 1) ∈ Xi × {1} with ( f i ( xi ), 0) ∈ Xi+1 × {0}. Compute the
fundamental group of M in the case when each Xi is a circle S1 and
f i : S1 → S1 is the map z 7→ zi (for each i ≥ 1).
5. For relatively prime positive integers m and n, the torus knot Km,n ⊂
R3 is the image of the embedding f : S1 → S1 × S1 ⊂ R3 , f (z) =
(zm , zn ), where the torus S1 × S1 is embedded in R3 in the standard
way. Compute π1 (R3 \ Km,n ).
classification of compact surfaces 31
3
Classification of compact surfaces
a3 a2
p4 p1
a2 a1
p5 a3 p0
a1 a2 a1−1 a3−1 a2 a3 .
P a P
b T2 b
P a P
with labeling scheme aba−1 b−1 . First, we glue the a labels together to
get a cylinder:
a
P P
b b
a
P
b
a a
with labeling scheme aa−1 , we get the sphere S2 by gluing the a labels
together.
Example 3.1.5 (The Projective Plane RP2 ). From the polygonal region
a a
with labeling scheme aa, we get the (real) projective plane RP2 by
gluing the a labels together.
Example 3.1.6 (The Klein Bottle K). We start with the following polygo-
nal region:
P a P
b K b
P a P
not glue together so nicely, that is, the surface we get by gluing them
together cannot be embedded in R3 . The resulting surface is called the
Klein bottle (figure 3.2).
π −1 ( X \π (C )) = P\π −1 (π (C ))
Lemma 3.1.9. If L1 and L2 are labeling schemes for M and N, then their
concatenation L1 L2 is a labeling scheme for M#N.
b1 b1
D1
a1
and let T22 be the following torus, and D2 a disk inside of it:
a2
b2 b2
D2
a2
a1 b1 a2 b2
b1 a1 b2 a2
∂D1 ∂D2
To get the connected sum of the two tori, we need to glue ∂D1 with
∂D2 , and we get the polygonal region which has the labeling scheme
that is, the concatenation of the labeling schemes of T12 and T22 .
36 algebraic topology
a1 b1 Figure 3.3: T 2 #T 2
b1 a1
a2 b2
b2 a2
n times
z }| {
Tn := T 2 # . . . #T 2
and
n times
z }| {
Pn := RP2 # . . . #RP2 .
• π1 (U, A) = Z ∗ Z = ⟨ a, b⟩
• π1 (V, x0 ) is trivial
• π1 (U ∩ V, x0 ) ∼
=Z∼
= ⟨c⟩
classification of compact surfaces 37
A a A
0
b b
x0
V
δ
A a A
π1 ( T 2 , x0 ) ∼
= π1 (U, x0 )/N ⟨i∗ ξ | ξ ∈ π1 (U ∩ V, x0 )⟩.
π1 ( T 2 , x0 ) ∼
= ⟨ ã, b̃⟩/N ⟨i∗ c⟩.
Hence π1 ( T 2 , x0 ) ∼
= ⟨ ã, b̃ | ãb̃ ã−1 b̃−1 ⟩ ∼
= Z × Z.
Similar calculations yield the following theorem:
Theorem 3.2.1. If X is the identification space of a labeling scheme
ϵ
a11 a2ϵ2 . . . aϵnn
2
Pn = RP
| . . . RP}2 : a1 a1 . . . an an
#{z
n−times
Notice that all vertices are identified in each labelling scheme. The
above theorem gives us:
π1 ( Tn ) = ⟨ a1 , b1 , . . . , an , bn | a1 b1 a1−1 b1−1 . . . an bn a− 1 −1
n bn = 1 ⟩
π1 ( Pn ) = ⟨ a1 , . . . , an | a21 . . . a2n = 1⟩
π1ab ( Pn ) = ⟨ a1 , . . . , a n | 2( a1 + . . . + a n ) = 0⟩
| {z }
An
= ⟨ a1 , . . . , an−1 , An | 2An ) = 0⟩ ∼
= Zn−1 × Z/2
Example 3.2.9. In the notation of Figure 3.4, we cut along the diagonal
labeled c and glue along a to show that
K∼
= P2 .
b c
b II
a
c a
b I
b c
c
K#RP2 ∼
= T 2 #RP2 ∼
= P3 .
Figure (3.7). Similarly, K#RP2 looks like a Möbius band with a twisted
handle attached to it. Cutting the band between the legs of the handle
leads to the same space as in Figure (3.7). Hence the assertion follows.
We begin with two results whose proofs you can find in Munkres’ book.
a a
The proof of the above theorem is based on the fact that each 2-
dimensional compact surface has a triangulation, and when we glue
half discs together along a common edge, we get a disc.
The arguments involved in the following classification of labelling
schemes provides the algorithm needed to identify any surface in the
list S2 , Tn , Pn , n ∈ N).
• S2 : aa−1
• Tn : a1 b1 a1−1 b1−1 . . . an bn a− 1 −1
n bn
• Pn : a1 a1 a2 a2 . . . an an
• second kind: a . . . a
Here are the steps involved in the cut and paste algorithm.
Step 1: Adjacent edges of the first kind can be removed. (See Figure
(3.9), where the edge labeled a is removed.)
a
a
c P
a
Q
c b
a Q P
Step 3: Make any pair of edges of second kind adjacent. (See Figure
(3.11).)
Here we cut along the edge c and, after flipping one of the two
pieces obtained, we glue along a. After removing the interior label a,
we created the subword cc, which corresponds to a pair of adjacent
edges of second kind.
Step 4: If a is an edge of the first kind, then there are two edges of the
−1
first kind which alternate: . . . a . . . a′ . . . a−1 . . . a′ . . ..
If this is not the case, the the edges of the region connecting the
vertices P in Figure (3.12) only get identified to edges from the same
region. The same applies for the region between the vertices labeled Q.
But then the endpoints of the edge a cannot be identified, contradicting
Step 2.
a
a a
P Q
P P
Step 5: Any two pairs of the first kind can be made consecutive. See
Figure (3.13).
a a
c
a a
b
b c
c
c d
a a
a
d
c c
44 algebraic topology
Here we first cut along c and glue along b, then cut along d and glue
along a.
a a a a
b d
e
b b e a
f
a a d d
d a
e e
e g e d
g
f g d f
f f
we first cut along e and glue along b, then cut along f and glue along a,
and finally cut along g and glue along d.
Exercises
1. There are six ways to obtain a compact surface by identifying pairs
of sides in a square. In each case determine what surface one obtains.
• abc−1 c−1 ba
• a 1 a 2 · · · a n a 1 −1 a 2 −1 · · · a n −1
• Show that π does not map all the vertices of P to the same point of
X.
A := {( x, y) ∈ R2 | 1 ≤ x2 + y2 ≤ 4}.
4
Covering spaces
In this chapter we introduce covering spaces and show how they can be
used for computing fundamental groups. In addition, we will use the
fundamental group as a tool for studying covering spaces.
Example 4.1.2. It is easy to check that the following maps are coverings.
(i) p : R → S1 , t 7→ e2πit .
(ii) id X : X → X.
(iii) p : X × {1, . . . , n} → X, ( x, k) 7→ x.
(iv) p : S1 → S1 , z 7→ zn .
(vi) p : C → C∗ , z 7→ ez .
Theorems 4.1.9 and 4.1.10 are generalizations from the case of the
covering p : R → S1 , with similar proofs.
is a bijection. Since the fiber p−1 (b0 ) has only two elements, we must
have that π1 (RPn , b0 ) ∼
= Z/2Z as a group isomorphism.
Example 4.1.15. Let p : R → S1 , t 7→ e2πit . Since R is both simply
connected and path-connected, Theorem 4.1.13 yields that
ϕe0 : π1 (S1 , b0 ) → Z
We have
ϕe0 ([γ] · [δ]) = ϕe0 ([γ ∗ δ]) = (^ e0 ∗ δe∗ )(1) = δe∗ (1)
γ ∗ δ )0 (1) = ( γ
= n + m,
where we set δe∗ (t) = n + δe0 (t) so that δe∗ (0) = n, δe∗ (1) = n + m. Thus,
ϕe0 is a homomorphism and therefore an isomorphism.
Proposition 4.1.16. If p : E → B is a covering and B is path-connected, then
for b0 , b1 ∈ B there is a bijection p−1 (b0 ) → p−1 (b1 ).
=γ
eδ̃e (1) ( 1 ) .
0
( f ◦^ f ◦ α ) e0 ∗ ( ]
(α ∗ β))e0 = ( ] f ◦ β)(]
f ◦α)
f ◦ α)e0 ∗ (^
= (] f ◦ β)(]
f ◦α)
e0 ( 1 ) e0 ( 1 )
f ◦ α)e0 ∗ (^
= (^ f ◦ β)(]
f ◦α) e0 ( 1 )
In this section, all spaces are assumed path-connected and locally path-
connected.
k◦h ( E, e0 )
id E p
p
( E, e0 ) ( B, b0 )
covering spaces 53
Using Lemma 4.2.10, there is e1 ∈ p−1 (b0 ) such that p′∗ π1 ( E′ , e0′ ) =
H ′ = p∗ π1 ( E, e1 ). By the lifting property (Theorem 4.1.19), there is an
equivalence h : E → E′ , thus finishing the proof of Theorem 4.2.9.
Theorem 4.3.5. Let B be path connected, locally path connected and semi-
locally simply connected. Let b0 ∈ B and H ⊂ π1 ( B, b0 ) a subgroup.
Then there is a covering p : E → B and a point e0 ∈ p−1 (b0 ) such that
p∗ π1 ( E, e0 ) = H.
E = { α# | α ∈ P }.
Example 4.3.6. The infinite earring has no universal cover, since it is not
semi-locally simply connected. The infinite earring is the space
[
X= Cn ,
n ≥1
1
where Cn is the circle of center (1/n, 0) and radius n. We claim that
j j∗
Cn X π1 (Cn , 0) π1 ( X, 0)
k k∗
i i∗
U π1 (U, 0)
In this section we study in more depth the relation between the fun-
damental group of the base of a covering on the one hand, and deck
transformations of the covering on the other hand. All spaces are again
path connected and locally path connected.
H = p∗ π1 ( E, e0 ) ⊂ π1 ( B, b0 ),
then
D( E, p) ∼
= N ( H )/H,
where N ( H ) = { g ∈ π1 ( B, b0 ) | gHg−1 = H } is the normalizer of H.
(Recall that N ( H ) is the largest subgroup of G which contains H as a normal
subgroup.)
ψ : D( E, p) → F, ψ(h) = h(e0 ).
ψ([α] H ) = e1 , ψ([ β] H ) = e2 .
covering spaces 57
Now let e3 = h(k (e0 )), so that ψ(h ◦ k) = e3 . We need to show that
ψ([α ∗ β] H ) = e3 .
D( E, p) ∼
= π1 ( B, p(e))/p∗ π1 ( E, e).
{(t, s) 7→ (t + n, s + m) | n, m ∈ Z} ∼
= Z2 .
π −1 (π (U )) =
[
gU
g∈ G
Using the above theorem and Corollary 4.4.2, we get the following.
π : S2n+1 → L( p; r2 , . . . , rn+1 )
Exercises
1. Show that the map p : S1 → S1 , p(z) = zn is a covering. (Here we
represent S1 as the set of complex numbers z of absolute value 1.)
3.
4.
5.
(i) Show that the torus T 2 is a two-fold cover of the Klein bottle.
5
Homology
We use the word “singular” because the image of such a map can
have “singularities”.
Let Cn ( X ) be the free abelian group with basis the singular n-
simplices in X, i.e.,
Cn ( X ) = ∑ i i i
n σ | n ∈ Z, σi : ∆ n
→ X continuous ,
i
where each formal sum ∑i ni σi is finite, i.e., all but finitely many ni are
zero. We call an element of Cn ( X ) an n-chain in X.
We define boundary maps
∂n : Cn ( X ) → Cn−1 ( X )
Note that both Im(∂n+1 ) and ker(∂n ) are subgroups of the abelian
group Cn ( X ). The above lemma yields that Im(∂n+1 ) is a subgroup of
ker(∂n ). Hence we can make the following.
Hn ( X ) := ker(∂n )/ Im(∂n+1 ).
Proposition 5.1.9. Suppose X is a space and { Xα }α∈ A are the path connected
components of X. Then, Hn ( X ) ∼
M
= Hn ( Xα ).
α∈ A
Cn ( X ) ∼
M
= Cn ( Xα ).
α
connected components.
Proof. From
1∂ ∂0
C1 ( X ) −
→ C0 ( X ) −
→0
and ∂0 = 0, we get that H0 ( X ) ∼
= C0 ( X )/ Im(∂1 ). Define the augmenta-
tion map
ϵ : C0 ( X ) −→ Z
∑ ni σi 7→ ∑ ni
i i
66 algebraic topology
H0 ( X ) ∼
=He0 ( X ) ⊕ Z
In this section we show that the homology groups are homotopy invari-
ants.
Let f : X → Y be a continuous map. Then, we have an induced
homomorphism
f # : Cn ( X ) → Cn (Y )
defined by f # (∑ ni σi ) = ∑ ni ( f ◦ σi ).
(b) (id X )∗ = id H∗ (X )
e n ( X ) = 0 for every n.
Corollary 5.2.8. If X is contractible, then H
Once defined, this will then show that g# and f # have the same effect on
homology. For, if α ∈ Cn ( X ) is a cycle, then by (⋆), we get that g# (α) −
f # (α) = ∂P(α) + P∂(α) = ∂P(α). Since f # and g# differ by a boundary,
they are homologous, so when quotient out by the boundaries, we get
that f ∗ ([α]) = g∗ ([α]) in homology.
It suffices to define P(σ ), for σ : ∆n → X a singular n-simplex, and
then we can extend P by linearity. We have the following maps:
(σ,id) F
∆n × I −−−→ X × I −
→Y
w0 w1
v0 v1
w2
w0 w1
v2
v0 v1
and
− F ◦ (σ, id)|[v0 ,...,vn ,wcn ] = − f ◦ σ = − f # (σ).
The terms with i ̸= j in the sum for ∂P(σ ) are exactly − P∂(σ ).
.. .. ..
. . .
i j
0 A n +1 Bn+1 Cn+1 0
∂ ∂ ∂
i j
0 An Bn Cn 0
∂ ∂ ∂
i j
0 A n −1 Bn−1 Cn−1 0
.. .. ..
. . .
homology 71
c + ∂(c′ ) = c + ∂( j(b′ ))
= c + j(∂(b′ ))
= j(b + ∂(b′ ))
So then, b will be replaced by b + ∂(b′ ), which leaves ∂(b) unchanged,
hence a unchanged.
One can use the connecting homomorphism ∂ just defined to prove
the following statement.
f # : Cn ( X, A) → Cn (Y, B).
0 → C• ( A) → C• ( X ) → C• ( X, A) → 0
· · · → Hn ( A) → Hn ( X ) → Hn ( X, A) → Hn−1 ( A) → · · ·
0 / C• ( A) / C• ( X ) / C• ( X, A) /0
ε ε 0
0 /Z /Z /0 /0
0 0 0
homology 73
· · · −→ H
e n ( A) −→ H
e n ( X ) −→ Hn ( X, A) −→ H
e n−1 ( A) −→ · · ·
en (X) ∼
H = Hn ( X, x0 )
for all n.
Corollary 5.3.5. There is a long exact sequence for the homology of a triple
( X, A, B), where B ⊆ A ⊆ X:
· · · −→ Hn ( A, B) −→ Hn ( X, B) −→ Hn ( X, A) −→ Hn−1 ( A, B) −→ · · ·
Proof. Start with the short short exact sequence of chain complexes
0 / C• ( A, B) / C• ( X, B) / C• ( X, A) /0
where maps are induced by inclusions of pairs, then take the associated
long exact sequence for homology as in Theorem 5.3.1.
for all n. Equivalently, if A, B ⊆ X are such that X = int( A) ∪ int( B), the
inclusion ( B, A ∩ B) ,→ ( X, A) induces isomorphisms
∼
=
Hn ( B, A ∩ B) −
→ Hn ( X, A)
for all n.
74 algebraic topology
Remark 5.3.8. To see that the two statements of the Excision Theorem
are equivalent, just take B = X \ Z (or Z = X \ B). Then A ∩ B = A \ Z,
and the condition Z ⊂ int( A) is equivalent to X = int( A) ∪ int( B).
HnU ( X ) → Hn ( X )
induced by inclusion.
ΣS1 ∼
= S2
S1
2. Hi (ΣX, S) ∼
= Hi (ΣX, Σ− X ). This can be seen in two ways:
3. Hi (ΣX, Σ− X ) ∼
= Hi (Σ+ X, X ). This follows by excising int(Σ− X ) and
using homotopy invariance for the homology of a pair.
(
Z, i=n
Corollary 5.3.10. H
ei (Sn ) =
0, i ̸= n.
Hk (U, U \ { x }) ∼
= Hk (Rm , Rm \ { x }) ∼ =He k−1 (Rm \ { x })
(
∼ e k −1 ( S m −1 ) ∼ Z, k = m
=H =
0, k ̸= m.
∼ i∗ r∗ ∼
Z
= /H
e n −1 ( S n −1 ) /H
e n −1 ( D n ) /H
e n −1 ( S n −1 ) =
4/ Z
idZ
ϕ ψ ∂
· · · → Hn ( A ∩ B) → Hn ( A) ⊕ Hn ( B) → Hn ( X ) → Hn−1 ( A ∩ B) →
· · · → H0 ( X ) → 0.
/ Cn ( A ∩ B) / Cn ( A) ⊕ Cn ( B) / Cn ( A + B)
/0
ϕ ψ
0
(5.3.1)
where, ϕ( x ) = ( x, − x ) for all x ∈ Cn ( A ∩ B) and ψ( x, y) = x + y for all
( x, y) ∈ Cn ( A) ⊕ Cn ( B). We claim that this sequence is exact:
b K
M2
b
a M1 a M1
Exercises
1. Show that if X is a path-connected topological space and f : X → X
is a continuous function, then the induced map f ∗ : H0 ( X ) → H0 ( X ) is
the identity map.
W W
5. For a wedge sum α Xα , the inclusions iα : Xα ,→ α Xα induce an
isomorphism
M M _
iα ∗ : e n ( Xα ) → H
H en ( Xα ),
α α α
provided that the wedge sum is formed at basepoints xα ∈ Xα such
that the pairs ( Xα , xα ) are good.
6. Show that:
∑ rankAn < ∞.
n ≥0
(i) Suppose
∂ ∂ ∂ ∂
· · · → Cn → Cn−1 → · · · → C1 → C0 → 0
(ii) Suppose we are given three finite type graded abelian groups A• ,
B• , C• , which are part of a long exact sequence
i
k jk ∂
· · · → Ak → Bk → Ck →k Ak−1 → · · · → A0 → B0 → C0 → 0.
Show that
χ( B• ) = χ( A• ) + χ(C• ).
homology 81
5.4 π1 vs. H1
h([ f ]) := f ∗ (α),
σ2
x0 y0
σ1
f
82 algebraic topology
∂(σ) = g − f · g + f .
v1
f g
v0 v2
f g
j : H1 ( X ) → π1 ( X, x0 )ab .
σ : = ϕx1 ∗ σ ∗ ϕx2 .
b
x3
σ2 σ1
x1 x2
σ0
j : H1 ( X ) → π1 ( X, x0 )ab .
(iv)
ϕ pi + σi + ϕqi = ϕ pi + σi − ϕqi ,
Degrees
Definition 5.5.1. The degree of continuous map f : Sn → Sn is defined as:
e n (Sn ) = Z → H
where f ∗ : H e n (Sn ) = Z is the homomorphism induced by f
in homology, and 1 ∈ Z denotes the generator.
1. deg idSn = 1.
g h
f : Sn −→ Sn \ {y} −→ Sn .
r ( x 0 , . . . x n ) = ( x 0 , . . . , x n −1 , − x n ).
so deg r = −1.
ei (X ) ∼
H =He i+1 (ΣX ), ∀i ≥ 0.
We already proved this fact by using the excision theorem 5.3.9. Here
we give another proof by using the Mayer-Vietoris sequence 5.3.14
for the decomposition
ΣX = C+ X ∪ X C− X,
where C+ X and C− X are the upper and lower cones of the suspen-
sion joined along their bases:
··· → H
e i+1 (C+ X ) ⊕ H
e i+1 (C− X ) → He i+1 (ΣX ) →
→H ei (X ) → H
e i (C+ X ) ⊕ H
e i (C− X ) → · · ·
(S f )∗ f∗
? ?
e i +1 ( S n +1 ) ∂- e n
H Hi (S )
∼
Note that C+ Sn /Sn ∼= Sn+1 so the boundary map ∂ at the top and
bottom of the diagram are the same map. So by the commutativity
of the diagram, since f ∗ is defined by multiplication by some integer
m, then (S f )∗ must be given by multiplication by the same integer
m.
where ri : Si → Si by ( x0 , x1 , . . . , xi ) 7→ (− x0 , x1 , . . . , xi ). So r0 : S0 →
S0 is given by x0 7→ − x0 . Note that S0 is two points but in reduced
homology we are only looking at one integer. Consider
e0 (S0 ) → H0 (S0 ) −
0→H →Z→0
ϵ
Sn
−x
f (x)
x
homology 87
((1 − t) · f ( x ) + t · (− x )
gt ( x ) : = : Sn → Sn .
||(1 − t) · f ( x ) + t · (− x )||
Exercises
1. Let f : Sn → Sn be a map of degree zero. Show that there exist points
x, y ∈ Sn with f ( x ) = x and f (y) = −y.
U1
x1
V
U2
x2
y
Um
xm
Note that by using excision and homology long exact sequences, one
has:
Hn (Ui , Ui \ xi ) ∼
= Hn (Sn , Sn \ xi ) ∼
=He n (Sn ) ∼
=Z
and
Hn (V, V \ y) ∼
= Hn (Sn , Sn \ y) ∼
=He n (Sn ) ∼
= Z.
Theorem 5.5.3. The degree of f equals the sum of local degrees at points in a
generic fiber, that is,
m
deg f = ∑ deg f |xi .
i =1
∼ exc ∼
=, ki = exc
?
m ?
Pi M exc f∗
Z∼
= Hn (Sn , S
n
\ xi ) Hn (Ui , Ui \ xi ) ∼
= Hn (Sn , Sn \ f −1 (y)) - Hn (Sn , Sn \ y)
i =1
∼
6
6 = l.e.s.
∼
= , l.
e.s
l.e.s. j
f∗
Z∼
=He n (Sn ) - e n (Sn ) ∼
H =Z
· deg f
k i (1) = (0, . . . , 0, 1, 0, . . . , 0)
where the entry 1 is in the ith place. Also, Pi ◦ j(1) = 1, for all i, so
m
j(1) = (1, 1, . . . , 1) = ∑ k i (1).
i =1
where the last equality follows from the commutativity of the upper
square.
CW Complexes
We next introduce cellular complexes (also referred to as cell-complexes
or CW complexes), and discuss a few important examples.
Start with a discrete set X0 , whose points are called 0-cells. In-
ductively, we form the n-skeleton Xn from Xn−1 by attaching n-cells
φn
eλn = Int( Dλn ) via maps ∂Dλn = Sλn−1 −→
λ
Xn−1 , i.e.,
Xn = Xn−1 ⨿λ Dλn ∼
Note that Φnλ | Int( Dn ) is a homeomorphism onto eλn , while the restriction
λ
of Φnλ to ∂Dλn is the attaching map φnλ .
A CW complex is endowed with the weak topology, i.e., A ⊂ X
is open ⇐⇒ A ∩ Xn is open for all n. An n-cell will be denoted by
eλn = Int( Dλn ). One can think of X as a disjoint union of cells of various
dimensions, or as ⨿n,λ Dλn ∼, where ∼ means that we are attaching
Φ : D2n → CPn
by v
n −1
u
: t1 − ∑ | z i |2 .
u
( z 0 , . . . , z n −1 ) 7 → z 0 : . . . : z n −1
i =0
Then Φ takes ∂D2n into the set of points with zn = 0, i.e., into CPn−1 .
Let φ := Φ|∂D2n . It is easy to check that Φ factors through CPn−1 ∪ φ D2n
and, moreover, the resulting map
Exercises
1. Let X and Y be finite CW complexes. Show that X × Y has the
structure of a finite CW complex with an (open) n + m dimensional cell
e × e′ for each n dimensional cell e in X and each m dimensional cell e′
in Y.
Cellular Homology
In this section, we show how to compute the homology of a CW
complex (assumed, for simplicity, to be of finite type). We first introduce
cellular homology and then we show that it can be identified with the
singular homology.
We start with the following preliminary result:
over, the homology long exact sequence of the pair ( Dλn , Dλn − { xλ })
yields that
Z, k=n
Hk ( Dλn , Dλn − { xλ }) ∼
=He k −1 ( S n −1 ) ∼
λ =
0, k ̸= n.
(b) Consider the following portion of the long exact sequence of the
pair for ( Xn , Xn−1 ):
Hk ( Xn ) ∼
= Hk ( Xn−1 ) ∼
= ··· ∼
= Hk ( X0 ).
(c) For simplicity, we only prove here the statement for finite dimen-
sional CW complexes. Let k < n and consider the following portion of
the long exact sequence for the pair ( Xn+1 , Xn ):
Hk ( Xn ) ∼
= Hk ( Xn+1 ).
Hk ( Xn ) ∼ = Hk ( Xn+2 ) ∼
= Hk ( Xn+1 ) ∼ = ··· ∼
= Hk ( Xn+l ) = Hk ( X ),
dn : Cn ( X ) −→ Cn−1 ( X )
defined by the following diagram, with diagonal arrows induced from long
exact sequences of pairs:
94 algebraic topology
Hn ( Xn+1 , Xn ) = 0
-
Hn ( Xn−1 ) = 0 Hn ( Xn+1 ) ∼
= Hn ( X )
-
in
-
Hn ( Xn )
-
1 jn
∂ n+
-
d n +1 dn
Hn+1 ( Xn+1 , Xn ) - Hn ( Xn , Xn−1 ) - Hn−1 ( Xn−1 , Xn−2 )
-
∂n
- jn −1
Hn−1 ( Xn−1 )
-
Hn−1 ( Xn−2 ) = 0
HnCW ( X ) ∼
= Hn ( X )
Hn ( X ) ∼
= Hn ( Xn )/ ker in ∼
= Hn ( Xn )/Im ∂n+1 .
Now, Hn ( Xn ) ∼
= Im jn ∼= ker ∂n ∼= ker dn . The first isomorphism
comes from jn being injective, the second follows by exactness, and
homology 95
Hn ( X ) ∼
= Hn ( Xn )/Im ∂n+1 = ker dn /Im dn+1 = HnCW ( X ),
Example 5.5.13. Recall that CPn has one cell in each even dimension
0, 2, 4, . . . , 2n. So CPn has no two cells in adjacent dimensions, meaning
we can apply Consequence (c) above to obtain:
Z, i = 0, 2, 4, . . . , 2n
Hi (CPn ) =
0, otherwise.
Example 5.5.14. When n > 1, Sn × Sn has one 0-cell, two n-cells, and
one 2n-cell. Since n > 1, these cells are not in adjacent dimensions so
again Consequence (c) above applies to give:
Z i = 0, 2n
n n
Hi (S × S ) = Z2 i=n
0 otherwise.
of the cellular chain complex. Let us consider the n-cells {eαn }α as the
basis for Cn ( X ) and the (n − 1)-cells {enβ−1 } β as the basis for Cn−1 ( X ).
In particular, we can write:
φnα
Sαn−1 = ∂Dαn −→ Xn−1 = Xn−2 ⊔γ eγn−1
collapse
−−−−→ Xn−1 /( Xn−2 ⊔γ̸= β eγn−1 ) = Snβ−1 ,
where φnα is the attaching map of eαn , and the collapsing map sends Xn−2 ⊔γ̸= β
eγn−1 to a point.
enβ−1
X n −1
X n −2
Proof. We will proceed with the proof by chasing the following diagram:
(∆αβ )∗
Hn ( Dαn , Sαn−1 )
∂ /H
e n−1 (Sαn−1 ) /H
e n −1 ( S n −1 )
≃
O β
(Φnα )∗ ( φnα )∗ q β∗
q∗
Hn ( Xn , Xn−1 )
∂n
/H
e n −1 ( X n −1 ) /H
e n−1 ( Xn−1 /Xn−2 )
dn
jn−1 ≃
'
Hn−1 ( Xn−1 , Xn−2 )
≃ / Hn−1 ( Xn−1 , Xn−2 )
X n −2 X n −2
where:
• Φnα is the characteristic map of the cell eαn and φnα is its attaching map.
Note that (∆αβ )∗ is defined so that the top right square commutes.
Recall that our goal is to compute dn (eαn ). The upper left square is
natural and therefore commutes (it is induced by the characteristic map
Φ : ( D ∗ , S∗−1 ) → ( X∗ , X∗−1 ) of a cell), while the lower left triangle
is part of the exact diagram defining the chain complex C∗ ( X ) and
is defined to commute as well. The map (Φnα )∗ takes the generator
[ Dαn ] ∈ Hn ( Dαn , Sαn−1 ) to a generator of the Z-summand of Hn ( Xn , Xn−1 )
corresponding to eαn , i.e.,
Since the top left square and the bottom left triangle both commute,
this gives that
we see from the definition of the above maps and the fact that ∂([ Dαn ])
e n−1 (Sαn−1 ), that (∆αβ )∗ is multiplication by dαβ .
is a generator of H
98 algebraic topology
d3 d2 d1 d0
0 /Z / Z2g /Z /0
d2 (e) = a1 + b1 − a1 − b1 + · · · a g + bg − a g − bg = 0.
d3 d2 d1 d0
0 /Z / Zg /Z /0
d2 (1) = (2, 2, · · · , 2)
H1 ( Ng ) ∼
= Zg /Im d2 ∼
= Zg /2Z ∼
= Zg−1 ⊕ Z/2.
Altogether,
homology 99
Z i=0
Hn ( Ng ) = Zg−1 ⊕ Z2 i=1
0 otherwise.
Example 5.5.18. Recall that RPn has a CW structure with one cell ek in
each dimension 0 ≤ k ≤ n. Moreover, the attaching map of ek in RPn
is the two-fold cover projection φ : Sk−1 → RPk−1 . The cellular chain
complex for RPn looks like:
d n +1 dn d2 d1 d0
0 /Z / ··· /Z /Z /0
In particular,
0 if k is odd
dk =
2 if k is even,
Exercises
1. Describe a cell structure on Sn ∨ Sn ∨ · · · ∨ Sn and calculate H∗ (Sn ∨
S n ∨ · · · ∨ S n ).
S k −2
q
∆
S k −1
S k −1
id
S k −1
4. Show that RP5 and RP4 ∨ S5 have the same homology and funda-
mental group. Are these spaces homotopy equivalent?
X×I
Tf = .
( x, 0) ∼ ( f ( x ), 1)
(a) Show that one can construct a CW-structure on L with one cell ek in
each dimension k ≤ 2n − 1.
G ≃ Zr × Z n 1 × · · · Z n k .
102 algebraic topology
d n +1 dn d2 d1 d0
0 / Cn / ... / C1 / C0 /0
0 / Zi ι / Ci di
/ / Bi−1 /0
d i +1 q
0 / Bi / Zi / Hi /0
and
rk( Zi ) = rk( Bi ) + rk( Hi ).
Substitute the second equality into the first, multiply the resulting
equality by (−1)i , and sum over i to get that χ( X ) = ∑in=0 (−1)i · rk( Hi ).
Finally, we apply this result to the cellular chain complex Ci =
Hi ( Xi , Xi−1 ), and use the identification between cellular and singular
homology.
Exercises
1. A graded abelian group is a sequence of abelian groups A• :=
( An )n≥0 . We say that A• is of finite type if
∑ rankAn < ∞.
n ≥0
homology 103
0 → An → Bn → Cn → 0, n ≥ 0.
χ( B• ) = χ( A• ) + χ(C• ).
Show that
χ( B• ) = χ( A• ) + χ(C• ).
χ ( X × Y ) = χ ( X ) · χ (Y ) .
where the latter trace is the linear algebraic trace of the map φ̄ : Zr →
Zr , with r = rk( G ). It is a fact that the trace is independent of the
choice of a basis for Zr .
104 algebraic topology
dim( X )
τ( f ) = ∑ (−1)i · Tr( f ∗ : Hi ( X ) → Hi ( X )). (5.7.2)
i =0
Remark 5.7.2. Notice that homotopic maps have the same Lefschetz
number since they induce the same maps on homology.
This follows from the fact that all the other homology groups are zero
and that the map induced on H0 is the identity.
The proof of this result is omitted for now. We proceed with sketch-
ing the proof of the Lefschetz theorem.
Proof. (sketch)
The general case reduces to the case when X is a finite CW complex.
Indeed, if r : K → X is a retraction of a finite CW complex K onto X,
the composition f ◦ r : K → X ⊂ K has exactly the same fixed points as
f and since r∗ : Hi (K ) → Hi ( X ) is projection onto a direct summand,
we have that Tr( f ∗ ◦ r∗ ) = Tr( f ∗ ), so τ ( f ◦ r ) = τ ( f ). We can therefore
assume that X is a finite CW complex.
Let us suppose that f has no fixed points.
By cellular approximation, the map f : X → X is homotopic to
a cellular map g : X → X. In particular, τ ( f ) = τ ( g). Moreover,
since f ( x ) ̸= x for all x ∈ X, it is possible to choose the cellular map
g : X → X so that g(eiλ ) ∩ eiλ = ∅, for all i and λ. Since the {eiλ }λ
generate Ci ( X ) := Hi ( Xi , Xi−1 ), we get that
∑(−1)i · Tr( g∗ : Ci (X ) → Ci (X )) = 0.
i
Furthermore, using the fact that the trace is additive for short exact
sequences, if follows as in the case of the Euler characteristic (Theorem
5.6.2) that
τ ( g) = ∑(−1)i · Tr( g∗ : Ci (X ) → Ci (X )).
i
Exercises
1. Is there a continuous map f : RP2k−1 → RP2k−1 with no fixed
points? Explain.
Tensor Products
Let A, B be abelian groups. Define the abelian group
A ⊗ B = ⟨ a ⊗ b | a ∈ A, b ∈ B⟩/ ∼ (5.8.1)
A×B
i / A⊗B
φ
∃! φ
$
C
Proof. Indeed, φ : A ⊗ B → C can be defined by a ⊗ b 7→ φ( a, b).
(1) A ⊗ B ∼
= B ⊗ A via the isomorphism a ⊗ b 7→ b ⊗ a.
Ai ) ⊗ B ∼
= i ( Ai ⊗ B) via the isomorphism ( ai )i ⊗ b 7→ ( ai ⊗ b)i .
L L
(2) ( i
(3) A ⊗ ( B ⊗ C ) ∼
= ( A ⊗ B) ⊗ C via the isomorphism a ⊗ (b ⊗ c) 7→ ( a ⊗
b) ⊗ c.
(4) Z ⊗ A ∼
= A via the isomorphism n ⊗ a 7→ na.
(5) Z/nZ ⊗ A ∼
= A/nA via the isomorphism l ⊗ a 7→ la.
Proof. These are easy to prove by using the above universal property.
We sketch a few.
(1) The map φ : A × B → B ⊗ A defined by ( a, b) 7→ b ⊗ a is clearly
bilinear and therefore induces a homomorphism φ : A ⊗ B → B ⊗ A
with a ⊗ b 7→ b ⊗ a. Similarly, there is the reverse map ψ : B × A →
A ⊗ B defined by (b, a) 7→ a ⊗ b which induces a homomorphism
ψ : B ⊗ A → A ⊗ B with b ⊗ a 7→ a ⊗ b. Clearly, φ ◦ ψ = id B⊗ A and
ψ ◦ φ = id A⊗ B and A ⊗ B ∼
= B ⊗ A.
(4) The map φ : Z × A → A defined by (n, a) 7→ na is a bilinear
map and therefore induces a homomorphism φ : Z ⊗ A → A with
homology 107
φ : C × G → B ⊗ G/Im(i ⊗ 1G )
φ ◦ ψ = id and ψ ◦ φ = id.
φ : C ⊗ G → B ⊗ G/Im(i ⊗ 1G )
Ci ( X; G ) = Ci ( X ) ⊗ G (5.8.2)
∂iG := ∂i ⊗ idG .
(C∗ ( X; G ), ∂∗G )
Ci ( X, A; G ) := Ci ( X; G )/Ci ( A; G ),
homology 109
∂iG+1 ∂Gi ∂G
2 1 ∂G ϵ
· · · −→ Ci ( X; G ) −→ · · · −→ C1 ( X; G ) −→ C0 ( X; G ) −→ G → 0,
CiG ( X ) = Hi ( Xi , Xi−1 ; G ) ∼
= G# i−cells .
where dαβ is as before the degree of a map ∆αβ : Si−1 → Si−1 . This
follows from the easy fact that if f : Sk → Sk has degree m, then
f ∗ : Hk (Sk ; G ) ≃ G → Hk (Sk ; G ) ≃ G is the multiplication by m. As it
is the case for integers, we get an isomorphism
HiCW ( X; G ) ≃ Hi ( X; G )
for all i.
X = S n ∪ g e n +1 ,
where the (n + 1)-cell en+1 is attached to Sn via the map g. Let f be the
quotient map f : X → X/Sn . Define Y = X/Sn = Sn+1 . The homology
of X can be easily computed by using the cellular chain complex:
d n +2 d n +1 dn d1 d1 d0
0 /Z /Z / ... /0 /Z /0
m
Therefore,
Z i=0
Hi ( X; Z) = Zm i=n
0 otherwise.
Exercises
1. Calculate the homology of the 2-torus T 2 with coefficients in Z, Z2
and Z3 , respectively. Do the same calculations for the Klein bottle.
∂n ∂ n −1 1 ∂
C• : · · · → Cn −→ Cn−1 −−→ · · · −
→ C0 → 0
Let us next explain the Tor functor appearing in the statement of the
universal coefficient theorem.
Definition 5.9.2. A free resolution of an abelian group H is an exact sequence:
f2 f1 f0
· · · → F2 −
→ F1 −
→ F0 −
→ H → 0,
F• ⊗ G : · · · → F2 ⊗ G → F1 ⊗ G → F0 ⊗ G → 0.
We define
Torn ( H, G ) := Hn ( F• ⊗ G ). (5.9.3)
Moreover, the following holds:
Lemma 5.9.3. For any two free resolutions F• and F•′ of H there are canonical
isomorphisms Hn ( F• ⊗ G ) ∼ = Hn ( F•′ ⊗ G ) for all n. Thus, Torn ( H, G ) is
independent of the free resolution F• of H used for its definition.
Proposition 5.9.4. For any abelian group H, we have that
and
Tor0 ( H, G ) ∼
= H ⊗ G. (5.9.5)
0 → F1 ,→ F0 ↠ H → 0
Tor( H, G ) := Tor1 ( H, G ).
(1) Tor( A, B) ∼
= Tor( B, A).
homology 113
Ai , B ) ∼
L L
(2) Tor( i = i Tor( Ai , B).
(4) Tor( A, B) ∼
= Tor(Torsion( A), B), where Torsion( A) is the torsion sub-
group of A.
n
(5) Tor(Z/nZ, A) ∼
= ker( A −
→ A ).
n
(5) The exact sequence 0 → Z −
→ Z → Z/nZ → 0 is a free resolution
of Z/nZ. Tensoring with A and dropping the right-most term yields
n ⊗1
the complex Z ⊗ A −−−→ A
Z ⊗ A → 0, which by property (4) of the
n n
→ A → 0. Thus, Tor(Z/nZ, A) = ker( A −
tensor product is A − → A ).
F1 = 0 → F0 = A → A → 0
0 0 0
0 F1 ⊗ B F1 ⊗ C F1 ⊗ D 0
0 F0 ⊗ B F0 ⊗ C F0 ⊗ D 0
0 0 0
Rows are exact since tensoring with a free group preserves exactness.
Thus we get a short exact sequence of chain complexes. Recall now
114 algebraic topology
0 → H1 ( F• ⊗ B) → H1 ( F• ⊗ C ) → H1 ( F• ⊗ D )
→ H0 ( F• ⊗ B) → H0 ( F• ⊗ C ) → H0 ( F• ⊗ D ) → 0
0 → Tor( A, B) → A ⊗ F1 → A ⊗ F0 → A ⊗ B → 0.
0 → Tor( B, A) → F1 ⊗ A → F0 ⊗ A → B ⊗ A → 0.
So we get a diagram:
0 Tor( A, B) A ⊗ F1 A ⊗ F0 A⊗B 0
ϕ ≃ ≃ ≃
0 Tor( B, A) F1 ⊗ A F0 ⊗ A B⊗A 0
with the arrow labeled ϕ defined as follows. The two squares on the
right commute since ⊗ is naturally commutative. Hence, there exists
ϕ : Tor( A, B) → Tor( B, A) which makes the left square commutative.
Moreover, by the 5-lemma, we get that ϕ is an isomorphism.
We can now prove the torsion free case of (3). Assume that B is torsion
f
free. Let 0 → F1 → F0 → A → 0 be a free resolution of A. The claim
about the vanishing of Tor( A, B) is equivalent to the injectivity of the
map f ⊗ id B : F1 ⊗ B → F0 ⊗ B. Assume ∑i xi ⊗ bi ∈ ker( f ⊗ id B ). So
homology 115
Hn ( X; Q) ≃ Hn ( X ) ⊗ Q.
= Z2 ⊗ Z/4 = (Z/4)2
Exercises
1. Prove Lemma 9.2.
6
Basics of Cohomology
with
C n := Hom(Cn , G ), (6.1.3)
and where the coboundary map
δ n : C n → C n +1 (6.1.4)
is defined by
It follows that
Ext groups
Let H and G be given abelian groups. Consider a free resolution of H,
f2 f1 f0
F• : · · · −→ F1 −→ F0 −→ H −→ 0.
f∗ f∗ f∗
F1∗ ←− F0∗ ←− H ∗ ←− 0,
2 1 0
· · · ←−
Extn ( H, G ) := H n ( F• ; G ). (6.2.1)
Lemma 6.2.1. The Ext groups are well-defined, i.e., they are independent of
the choice of the free resolution F• of H.
As in the case of the Tor functor, one can thus work with the free
resolution of H given by
0 −→ F1 −→ F0 −→ H −→ 0,
Ext0 ( H, G ) = Hom( H, G ).
Proposition 6.2.2. The Ext group Ext( H, G ) satisfies the following proper-
ties:
Proof. Indeed, H decomposes into a free part and a torsion part, and
the claim follows by Proposition 6.2.2.
H n (C• ; G ) ∼
= Ext( Hn−1 (C• ), G ) ⊕ Hom( Hn (C• ), G ). (6.2.5)
120 algebraic topology
Proof. (Sketch)
The homomorphism h : H n (C• ; G ) → Hom( Hn (C• ), G ) is defined as
follows. Let Zn = ker ∂n , Bn = Im ∂n+1 , in : Bn ,→ Zn the inclusion map,
and Hn (C• ) = Zn /Bn . Let [ϕ] ∈ H n (C• ; G ). Then ϕ is represented by a
homomorphism ϕ : Cn → G, so that δn ϕ := ϕ∂n+1 = 0, which implies
that ϕ| Bn = 0. Let ϕ0 := ϕ| Zn , then ϕ0 vanishes on Bn , so it induces a
quotient homomorphism ϕ¯0 : Zn /Bn → G, i.e., ϕ¯0 ∈ Hom( Hn (C• ), G ).
We define h by
h([ϕ]) = ϕ¯0 .
Notice that if ϕ ∈ Im δn−1 , i.e., ϕ = δn−1 ψ = ψ∂n , then ϕ| Zn = 0, so
ϕ¯0 = 0, which shows that h is well-defined. It is not hard to show that
h is an epimorphism, and
where the Ext group is defined with respect to the free resolution of
Hn−1 (C• ) given by
i n −1
0 −→ Bn−1 −→ Zn−1 −→ Hn−1 (C• ) −→ 0.
Corollary 6.2.6. Let (C• , ∂• ) be a chain complex so that its (integral) homol-
ogy groups H∗ are finitely generated, and let Tn = Torsion( Hn ). Then we
have natural short exact sequences:
H n (C• ; Z) ∼
= Tn−1 ⊕ Hn /Tn . (6.2.8)
The claim follows by the five-lemma, since α∗ and its dual are isomor-
phisms.
C n ( X; G ) := Hom(Cn ( X ), G ). (6.3.1)
δn : C n ( X; G ) → C n+1 ( X; G )
It follows that
δn+1 ◦ δn = 0, (6.3.3)
and for a singular (n + 1)-simplex σ : ∆ n +1 → X we have:
n +1
δn ψ(σ ) = ∑ (−1)i · ψ(σ|[v0 ,··· ,v̂i ,··· ,vn+1 ] ). (6.3.4)
i =0
H 0 ( X; G ) = Hom( H0 ( X ), G ), (6.3.7)
Remark 6.3.3. Theorem 6.2.4 also works for modules over a PID. In
particular, if G = F is a field, then
H n ( X; F ) ≃ Hom( Hn ( X ), F ) ≃ Hom F ( Hn ( X; F ), F ) = Hn ( X, F )∨
And since
Z, i=0
Hi ( X ) =
0, otherwise,
we get
G, i=0
Hom( Hi ( X ), G ) =
0, otherwise.
Altogether,
G, i=0
H i ( X; G ) =
0, otherwise.
· · · −→ C1 ( X ) −→ C0 ( X ) −→ Z −→ 0,
∂ ∂ ϵ
ϵ∗
· · · ←− C1 ( X; G ) ←− C0 ( X; G ) ←− G ←− 0.
δ δ
e i ( X; G ) = H i ( X; G ), if i > 0,
H
e 0 ( X; G ) = Hom( H
H e 0 ( X ), G ).
δ : C n ( X, A; G ) → C n+1 ( X, A; G ) (6.3.11)
i j
0 −→ Cn ( A) −→ Cn ( X ) −→ Cn ( X, A) −→ 0
i∗ j∗
0 ←− C n ( A; G ) ←− C n ( X; G ) ←− C n ( X, A; G ) ←− 0, (6.3.12)
124 algebraic topology
i∗ j∗
0 ←− C ∗ ( A; G ) ←− C ∗ ( X; G ) ←− C ∗ ( X, A; G ) ←− 0. (6.3.13)
j∗ i∗
· · · → H n ( X, A; G ) → H n ( X; G ) → H n ( A; G ) → H n+1 ( X, A; G ) → · · ·
δ
(6.3.14)
We can also consider above the augmented chain complexes on X and
A, and get a long exact sequence for the reduced cohomology groups,
with H e n ( X, A; G ) = H n ( X, A; G ):
· · · → H n ( X, A; G ) → H e n ( A; G ) → H n+1 ( X, A; G ) → · · ·
e n ( X; G ) → H
(6.3.15)
In particular, if A = x0 is a point in X, we get by (6.3.15) that
e n ( X; G ) ∼
H = H n ( X, x0 ; G ). (6.3.16)
Induced homomorphisms
Recall that if f : X → Y is a continuous map, we have induced chain
maps
f# : Cn ( X ) / Cn (Y )
(σ : ∆n → X )
f
/ ( f ◦ σ : ∆n → σ
X → Y)
f # : C n (Y; G ) → C n ( X; G ),
In fact, we can repeat the above construction for maps of pairs, say
f : ( X, A) → (Y, B). And note that the universal coefficient theorem
also works for pairs because Cn ( X, A) = Cn ( X )/Cn ( A) is free abelian.
So, by naturality, we get a commutative diagram for a map of pairs
f : ( X, A) → (Y, B):
Homotopy invariance
In this subsection we show that cohomology groups are homotopy
invariants of spaces.
f ∗ = g∗ : H n (Y, B; G ) → H n ( X, A; G ).
Proof. Recall from the proof of the similar statement for homology that
there is a prism operator
satisfying
f # − g# = P∂ + ∂P, (6.3.18)
Cn ( X, A)
∂ / Cn−1 ( X, A)
P
f# g#
x x P
Cn+1 (Y, B)
∂ / Cn (Y, B)
f # − g# = δP∗ + P∗ δ. (6.3.20)
Excision
Theorem 6.3.9. Given a topological space X, suppose that Z ⊂ A ⊂ X, with
cl ( Z ) ⊆ int( A). Then the inclusion of pairs i : ( X \ Z, A \ Z ) ,→ ( X, A)
induces isomorphisms
i∗ : H n ( X, A; G ) → H n ( X \ Z, A \ Z; G ) (6.3.21)
(i ∗ ) ∗ i∗ (i ∗ ) ∗
0 / Ext( Hn−1 ( X \ Z, A \ Z ), G ) / H n ( X \ Z, A \ Z; G ) / Hom( Hn ( X \ Z, A \ Z ), G ) / 0
Mayer-Vietoris sequence
Theorem 6.3.10. Let X be a topological space, and A and B be subsets of X
so that
X = int( A) ∪ int( B).
Then there is a long exact sequence of cohomology groups:
ψ ϕ
· · · −→ H n ( X; G ) −→ H n ( A; G ) ⊕ H n ( B; G ) −→ H n ( A ∩ B; G )
−→ H n+1 ( X; G ) −→ · · · (6.3.22)
basics of cohomology 127
ψ ϕ
0 C n ( A + B; G ) C n ( A; G ) ⊕ C n ( B; G ) C n ( A ∩ B; G ) 0
Hom(Cn ( A + B), G )
and
ϕ(α, β) = α|Cn ( A∩ B) − β|Cn ( A∩ B) .
e i (Sn ; G ) ∼
H =He i −1 ( S n −1 ; G ) ∼
= ··· ∼
=He i − n ( S0 ; G )
(
∼ G, i = n
=
0, otherwise.
Cellular cohomology
C n ( X; G ) := H n ( Xn , Xn−1 ; G ),
dn = δn ◦ jn
128 algebraic topology
fitting in the following diagram (where the coefficient group for cohomology is
by default G):
H n −1 ( X n −1 )
:
j n −1 δ n −1
#
· · · / H n −1 ( X n −1 , X n −2 ) / H n ( X n , X n −1 ) / H n +1 ( X n +1 , X n ) / · · ·
d n −1 dn
<
jn δn
!
H n ( Xn )
Here, the diagonal arrows are part of cohomology long exact sequences for the
relevant pairs. For this reason, it follows that jn δn−1 = 0, and therefore
for all n and any coefficient group G. Moreover, the cellular cochain com-
plex (C • ( X; G ), d• ) is isomorphic to the dual of the cellular chain complex
(C• ( X ), d• ), obtained by applying Hom(−; G ).
Proof. Recall from Section 5.5 that for the cellular chain complex of X
we have that
Cn ( X ) := Hn ( Xn , Xn−1 ) ∼
= Z# of n-cells ,
C n ( X; G ) := H n ( Xn , Xn−1 ; G ) ∼
= Hom(Cn ( X ), G ) (6.3.24)
since the Ext term vanishes. The universal coefficient theorem also
yields that
H i ( Xn , Xn−1 ; G ) = 0 if i ̸= n, (6.3.25)
since the groups Hi ( Xn , Xn−1 ) are either free or trivial. From the long
exact sequence of the pair ( Xn , Xn−1 ), that is,
· · · −→ H k ( Xn , Xn−1 ; G ) −→ H k ( Xn ; G ) −→ H k ( Xn−1 ; G )
−→ H k+1 ( Xn , Xn−1 ; G ) −→ · · · ,
basics of cohomology 129
H k ( Xn ; G ) ∼
= H k ( X n −1 ; G ). (6.3.26)
H k ( Xn ; G ) ∼
= H k ( X n −1 ; G ) ∼
= H k ( X n −2 ; G ) ∼
= ··· ∼
= H k ( X0 ; G ) = 0
(6.3.27)
since X0 is just a set of points.
We next claim that there is an isomorphism
H n ( X n +1 ; G ) ∼
= H n ( X; G ). (6.3.28)
First recall from Lemma 5.5.10(c) that the inclusion Xn+1 ,→ X induces
isomorphisms on homology groups Hk , for k < n + 1. So by the
naturality of the universal coefficient theorem, we get the following
diagram with commutative squares:
h
0 Ext( Hn−1 ( X ), G ) H n ( X; G ) Hom( Hn ( X ), G ) 0
(i∗ )∗ ∼
= i∗ (i∗ )∗ ∼
=
h
0 Ext( Hn−1 ( Xn+1 ), G ) Hn (X n +1 ; G ) Hom( Hn ( Xn+1 ), G ) 0
i∗ : H n ( X; G ) → H n ( Xn+1 ; G )
is also an isomorphism.
Altogether, by using (6.3.27) and (6.3.28), we get the following dia-
gram (where the diagonal arrows are part of long exact sequences of
pairs):
H n −1 ( X n −2 ) ∼
=0
9
H n −1 ( X n −1 )
:
j n −1 δ n −1
%
··· / H n −1 ( X n −1 , X n −2 )
d n −1
/ H n ( X n , X n −1 )
dn
/ H n +1 ( X n +1 , X n ) / ···
<
jn δn
$
n
H ( Xn )
:
: "
α
H (X) ∼
n
= H ( X n +1 ) n
H n ( X n −1 ) ∼
=0
H n ( X; G ) ∼
= H n ( X n +1 ; G )
∼
= Im(α)
∼
= ker(δn )
∼ ker(dn )/ ker( jn )
= (6.3.29)
∼
= ker(dn )/Im(δn−1 )
∼
= ker(dn )/Im(δn−1 jn−1 )
∼
= ker(dn )/Im(dn−1 ).
d n = ( d n +1 ) ∗ . (6.3.30)
and, similarly, the boundary map dn+1 of the cellular chain complex is
given by:
∂ n +1 jn
dn+1 : Hn+1 ( Xn+1 , Xn ) −→ Hn ( Xn ) −→ Hn ( Xn , Xn−1 ).
Let us now consider the following diagram:
jn
H n ( X n , X n −1 ; G ) / H n ( Xn ; G ) δn
/ H n +1 ( X n +1 , X n ; G )
∼
= h h ∼
= h
( jn )∗ )∗
Hom( Hn ( Xn , Xn−1 ), G ) / Hom( Hn ( Xn ), G) ( ∂ n +1
/ Hom( Hn+1 ( Xn+1 , Xn ), G)
The composition across the top is the cellular coboundary map dn , and
we want to conclude that it is the same as the composition (dn+1 )∗
across the bottom row. The extreme vertical arrows labelled h are iso-
morphisms by the universal coefficient theorem, since the relevant Ext
terms vanish (by using (6.3.25)). So it suffices to show that the diagram
commutes. The left square commutes by the naturality of universal
coefficient theorem for the inclusion map ( Xn , ∅) ,→ ( Xn , Xn−1 ), and
the right square commutes by a simple diagram chase.
Example 6.3.15. Let X = RP2 . Then X has one cell in each dimension
0, 1, and 2, and the cellular chain complex of X is:
0 /Z 2 /Z 0 /Z /0.
0o Zo Zo Zo
2 0
0.
basics of cohomology 131
Thus, we have
Z,
i=0
i 2
H (RP ; Z) = Z/2, i = 2
0, otherwise.
We then have:
(
Z/2, i = 0, 1, or 2
H i (RP2 ; Z/2) ∼
=
0, otherwise.
Therefore, we have
(
Z/3, i = 0 or 1
H i (K; Z/3) =
0, otherwise.
Exercises
1. Prove Lemma 6.2.1.
10. Let X be the space obtained by attaching two 2-cells to S1 , one via
the map z 7→ z3 and the other via z 7→ z5 , where z denotes the complex
coordinate on S1 ⊂ C. Compute the cohomology groups H ∗ ( X; G ) of
X with coefficients:
(a) G = Z.
(b) G = Z/2.
(c) G = Z/3.
cup product in cohomology 133
7
Cup Product in Cohomology
Let us motivate this chapter with the following simple, but hopefully
convincing example. Consider the spaces X = CP2 and Y = S2 ∨ S4 .
As CW complexes, both X and Y have one 0-cell, one 2-cell and one
4-cell. Hence the cellular chain complex for both X and Y is:
0 0 0 0
0 −→ Z −→ 0 −→ Z −→ 0 −→ Z −→ 0
π1 ( X ) = π1 (Y ) = 0.
k +1
(δϕ ∪ ψ)(σ) = ∑ (−1)i ϕ(σ|[v0 ,··· ,bvi ,··· ,vk+1 ] ) · ψ(σ|[vk+1 ,··· ,vk+l+1 ] )
i =0
and
k + l +1
(−1)k (ϕ ∪ δψ)(σ) = ∑ (−1)i ϕ(σ|[v0 ,··· ,vk ] ) · ψ(σ|[vk ,··· ,bvi ,··· ,vk+l +1 ] ).
i =k
When we add these two expressions, the last term of the first sum
cancels with the first term of the second sum, and the remaining terms
are exactly δ(ϕ ∪ ψ)(σ ) = (ϕ ∪ ψ)(∂σ ) since
k + l +1
∂σ = ∑ (−1)i σ |[v0 ,··· ,bvi ,··· ,vk+l +1 ] .
i =0
Corollary 7.1.3. The cup product of two cocycles is again a cocycle. That is,
if ϕ, ψ are cocycles, then δ(ϕ ∪ ψ) = 0.
Example 7.1.5. Let us consider the real projective plane RP2 . Its Z/2Z-
cohomology is computed by:
Z/2Z for i = 0, 1, 2
H i (RP2 ; Z/2Z) =
0 otherwise.
α2 := α ∪ α ∈ H 2 (RP2 ; Z/2Z).
Consider the cell structure on RP2 with two 0-cells v and w, three
1-cells e, e1 and e2 , and two 2-cells T1 and T2 . The 2-cell T1 is attached
by the word e1 ee2−1 , and the 2-cell T2 is attached by the word e2 ee1−1
(see the figure below). We can of course regard these cells as singular
simplices as well.
w
e1
e T1 v T2 e
e2
ϕ : C1 (RP2 ) → Z/2Z
with ϕ(e) = 1, where we use the fact that e represents the generator of
H1 (RP2 ). The cocycle condition for ϕ translates into the identities:
C k ( X; R) × C l ( X; R) −→ C k+l ( X; R)
C k ( X, A; R) × C l ( X; R) −→ C k+l ( X, A; R),
C k ( X, A; R) × C l ( X, A; R) −→ C k+l ( X, A; R),
and
C k ( X; R) × C l ( X, A; R) −→ C k+l ( X, A; R)
since Ci ( X, A; R) can be regarded as the set of cochains vanishing on
chains in A, and if ϕ or ψ vanishes on chains in A, then so does ϕ ∪ ψ.
So there exist relative cup products:
∪
H k ( X, A; R) × H l ( X; R) −→ H k+l ( X, A; R),
∪
H k ( X, A; R) × H l ( X, A; R) −→ H k+l ( X, A; R),
and
∪
H k ( X; R) × H l ( X, A; R) −→ H k+l ( X, A; R).
In particular, if A is a point, we get a cup product on the reduced
cohomology H e ∗ ( X; R).
More generally, there is a cup product
∪
H k ( X, A; R) × H l ( X, B; R) −→ H k+l ( X, A ∪ B; R)
C k ( X, A; R) × C l ( X, B; R) −→ C k+l ( X, A + B; R),
f # ( ϕ ∪ ψ ) = f # ( ϕ ) ∪ f # ( ψ ),
f # (ϕ) ∪ f # (ψ)(σ : ∆k+l → X ) = ( f # ϕ)(σ |[v0 ,··· ,vk ] ) · ( f # ψ)(σ |[vk ,··· ,vk+l ] )
= ϕ(( f # σ)|[v0 ,··· ,vk ] ) · ψ(( f # σ)|[vk ,··· ,vk+l ] )
= (ϕ ∪ ψ)( f # σ)
= ( f # (ϕ ∪ ψ))(σ).
k ≥0
with respect to the cup product operation. If R has an identity, then so does
H ∗ ( X; R). Similarly, we define the cohomology ring of a pair H ∗ ( X, A; R) by
using the relative cup product.
Remark 7.1.9. By scalar multiplication with elements of R, we can
regard these cohomology rings as R-algebras.
The following is an immediate consequence of Lemma 7.1.6:
Corollary 7.1.10. If f : X → Y is a continuous map then we get an induced
ring homomorphism
f ∗ : H ∗ (Y; R) → H ∗ ( X; R).
e∗( Xα ; R ) ∼ ∏ He ∗ (Xα ; R)
_
H = (7.1.6)
α α
Example 7.1.13.
H ∗ ( S n , Z) = Z[ α ] / ( α2 )
where α is a generator of H n (Sn ; Z). Indeed, we have
Z for i = 0, n
H i ( S n ; Z) =
0 otherwise.
RP∞ = RPn
[
n ≥0
(a)
H ∗ (RPn ; Z/2Z) ∼
= Z/2[α]/(αn+1 )
where α is the generator of H 1 (RPn ; Z/2Z).
(b)
H ∗ (RP∞ ; Z/2Z) ∼
= Z/2[α]
where α is the generator of H 1 (RPn ; Z/2Z).
(c)
H ∗ (CPn ; Z) = Z[ β]/( βn+1 )
where β is the generator of H 2 (CPn ; Z).
cup product in cohomology 139
(d)
H ∗ (CP∞ ; Z) = Z[ β]
Before discussing the proof of the above theorem, let us get back to
the following motivating example:
Example 7.1.15. We saw at the beginning of this chapter that the spaces
X = CP2 and Y = S2 ∨ S4 have the same homology and cohomology
groups, and even the same CW structure. The cup products can be
used to decide whether these spaces are homotopy equivalent. Indeed,
let us consider the cohomology rings H ∗ ( X; Z) and H ∗ (Y; Z). From
the above theorem, we have that:
H ∗ (CP2 ; Z) = Z[ β]/( β3 ),
Let us now get back to the proof of Theorem 7.1.14. We will discuss
below the proof in the case of RPn . The result in the case of RP∞ fol-
lows from the finite-dimensional case since the inclusion RPn ,→ RP∞
induces isomorphisms on H i (−; Z/2) for i ≤ n by cellular cohomol-
ogy. The complex projective spaces are handled in precisely the same
manner, using Z-coefficients and replacing H k by H 2k and R by C.
We next prove the following result:
Theorem 7.1.16.
Pn := RPn
Let
i
Si = {( x0 , · · · , xi , 0, · · · , 0) | ∑ xl2 = 1}
l =0
and
n
S j = {(0, · · · , 0, xn− j , · · · xn ) | ∑ xl2 = 1}
l =n− j
Sn
Sj
Si
Pi ∩ P j = { p } = ( 0 : · · · : 0 : 1 : 0 : · · · : 0 )
is a homeomorphism U ∼
= Rn which takes p to 0 ∈ Rn .
cup product in cohomology 141
P j −1
Pj
Pi − 1 p Pi − 1
Pi
P j −1
H i ( P n , P n − P j ) × H j ( P n , P n − Pi ) / H n (Pn , Pn − { p})
H ( R , R − R ) × H j ( R n , R n − Ri )
i n n j / H n (Rn , Rn − {0})
∼
= H j ( D j , S j −1 )
and
H n (Rn , Rn − {0}) ∼
= H n ( D n , S n −1 )
∼
= H n ( D i × D j , S i −1 × D j ∪ S j −1 × D i ).
[ D i ] × [ D j ] 7 → [ D n ].
The same will be true for the top row, provided we show that the four
vertical maps in the above diagram are isomorphisms.
For the bottom right vertical arrow, we have by excision that
H n (Pn , Pn − { p}) ∼
= H n (U, U − { p}) ∼
= H n (Rn , Rn − {0}), (7.1.8)
H n (Pn , Pn − { p}) ∼
= H n (Pn , Pn −1 ) ∼
= Z/2, (7.1.9)
H n (Pn , Pn − { p}) ∼
= H n (Pn ) (7.1.10)
To show that the two left vertical arrows are isomorphisms, consider
the following commutative diagram.
(5)
H i (Pn ) o H i ( P n , Pi − 1 ) o H i (Pn , Pn − P j ) / H i (Rn , Rn − R j )
(2) (4)
(1) (3) (6) (7)
(10)
H i ( Pi ) o H i ( Pi , Pi − 1 ) o H i (Pi , Pi − { p}) / H i (Ri , Ri − {0})
(8) (9)
Example 7.1.17. Let us consider the spaces RP2n+1 and RP2n ∨ S2n+1 .
First note that these spaces have the same CW structure and the same
cellular chain complex, so they have the same homology and cohomol-
ogy groups. However, we claim that RP2n+1 and RP2n ∨ S2n+1 are not
homotopy equivalent. In order to justify the claim, we first compute
their Z/2Z-cohomology rings.
From the above theorem, the cohomology ring of RP2n+1 is:
with H ∗ (RP2n ; Z/2Z) ∼ = Z/2Z[ β]/( β2n+1 ) for β the degree 1 genera-
tor of H 1 (RP2n ; Z/2Z), and H ∗ (S2n+1 ; Z/2Z) ∼ = Z/2Z[γ]/(γ2 ) for γ
the generator of H 2n + 1 (S 2n + 1 ; Z/2Z) of degree 2n + 1.
If there was a homotopy equivalence f : RP2n+1 → RP2n ∨ S2n+1 , then
the generators of degree one would correspond isomorphically to each
other, i.e., we would get f ∗ ( β) = α. But as f ∗ is a ring isomorphism,
this would then imply that: f ∗ ( β2n+1 ) = ( f ∗ ( β))2n+1 = α2n+1 . How-
ever, this yields a contradiction, since β2n+1 = 0, thus f ∗ ( β2n+1 ) = 0,
while α2n+1 ̸= 0 since α2n+1 generates H 2n+1 (RP2n+1 ; Z/2Z).
144 algebraic topology
a map
f : RPn → RPm
[ x ] 7→ [ g( x )]
which makes the following diagram commutative:
g
Sn / Sm
p′ ↷ p
f
RPn / RPm
f ∗ : π1 (RPn ) ∼
= Z/2 → π1 (RP1 ) ∼
=Z
We have:
0 = f ∗ (αm +1 ∗
m ) = f (αm )
m +1
,
so f ∗ (αm ) ∈ H 1 (RPn ; Z/2) has order m + 1 < n + 1. Therefore,
f ∗ (αm ) ̸= αn .
f ∗ (αm ) = 0.
j∗ (αm ) ̸= 0.
( f ◦ i )∗ (αm ) = i∗ ( f ∗ (αm )) = 0.
Exercises
1. Show that if X is the union of contractible open subsets A and B,
then all cup products of positive-dimensional classes in H ∗ ( X ) are
zero. In particular, this is the case if X is a suspension. Conclude that
spaces such as RP2 and T 2 cannot be written as unions of two open
contractible subsets.
nullhomotopic? Explain.
4. Show that RP3 and RP2 ∨ S3 have the same cohomology rings with
integer coefficients.
5.
(a) Show that H ∗ (CPn ; Z) ∼
= Z[ x ]/( x n+1 ), with x the generator of
H 2 (CPn ; Z).
(a) Show that the Lefschetz number τ f of a map f : CPn → CPn is given
by
τ f = 1 + d + d2 + · · · + d n ,
where f ∗ ( x ) = dx for some d ∈ Z, and with x as in part ( a).
(c) Show that for n even, any map f : CPn → CPn has a fixed point.
Cross product
Let us motivate this section by consider the spaces S2 × S3 and S2 ∨ S3 ∨
S5 . Both spaces are CW complexes with cells {e0 , e2 , e3 , e5 } in degrees,
0, 2, 3 and 5, respectively. So the cellular chain complex for both spaces
is:
0
0 → Z → 0 → Z −→ Z → 0 → Z → 0
Hence both spaces have the same homology and cohomology groups.
It is then natural to ask the following:
The aim of this section is to convince the reader that the answer is
No. More precisely, we will show that the two spaces have different
cohomology rings.
The cohomology ring H ∗ (S2 ∨ S3 ∨ S5 ; Z) can be computed from the
ring isomorphism
e ∗ ( S2 ∨ S3 ∨ S5 ; Z) ∼
H =He ∗ ( S2 ; Z) ⊕ H
e ∗ ( S3 ; Z) ⊕ H
e ∗ ( S5 ; Z),
with H ∗ (S2 ; Z) ∼
= Z[ α ] / ( α2 ), H ∗ ( S3 ; Z) ∼
= Z[ β]/( β2 ) and H ∗ (S5 ; Z) ∼
=
Z[γ]/(γ ), where α is the generator of H (S ; Z), β is the generator of
2 2 2
p : S2 ∨ S3 ∨ S5 → S2 ∨ S3
α ∪ β = p∗ ᾱ ∪ p∗ β̄ = p∗ (ᾱ ∪ β̄) = 0
since ᾱ ∪ β̄ = 0.
By the end of this section, we will show that the product of the gen-
erators of degree 2 and degree 3 in the cohomology ring of S2 × S3 is
the generator in degree 5, so it is non-zero. This will then completely
answer the above question.
α ∪ β = (−1)kl · β ∪ α. (7.3.1)
×( a ⊗ b) := a × b.
Indeed, we have:
H n ( X × Y; R) ∼ H i ( X; R) ⊗ R H j (Y; R)
M
= (7.3.4)
i + j=n
H ∗ ( S2 × S3 ; Z) ∼
= H ∗ ( S2 ; Z) ⊗Z H ∗ ( S3 ; Z)
150 algebraic topology
αi α j = −α j αi , if i ̸= j
α2i = 0.
H ∗ ( S3 × S5 × S7 ; Z) ∼
= ΛZ [ a3 , a5 , a7 ], (7.3.5)
H ∗ ( S3 × S5 × S7 ; Z) ∼
= H ∗ ( S3 ; Z) ⊗Z H ∗ ( S5 ; Z) ⊗Z H ∗ ( S7 ; Z).
• a3 = ×(α3 ⊗ 1 ⊗ 1)
• a5 = ×(1 ⊗ α5 ⊗ 1)
• a7 = ×(1 ⊗ 1 ⊗ α7 )
= × ( α3 ⊗ 1 ⊗ 1) · ( α3 ⊗ 1 ⊗ 1)
= ×(α23 ⊗ 1 ⊗ 1)
= 0
and a similar result for a25 and a27 .
a3 a5 = ×(α3 ⊗ 1 ⊗ 1) ∪ ×(1 ⊗ α5 ⊗ 1)
= × ( α3 ⊗ 1 ⊗ 1) · (1 ⊗ α5 ⊗ 1)
= (−1)0·0 × (α3 ⊗ α5 ⊗ 1)
= ×(α3 ⊗ α5 ⊗ 1)
a5 a3 = ×(1 ⊗ α5 ⊗ 1) ∪ ×(α3 ⊗ 1 ⊗ 1)
= × (1 ⊗ α5 ⊗ 1) · ( α3 ⊗ 1 ⊗ 1)
= (−1)3·5 × (α3 ⊗ α5 ⊗ 1)
= − a3 a5
We have similar results for the other products too. The above cal-
culations show that we have an isomorphism H ∗ (S3 × S5 × S7 ; Z) ∼
=
ΛZ [ a3 , a5 , a7 ].
Remark 7.3.10. It is easy to see that a similar result holds for the
cohomology ring of any (finite) product of odd dimensional spheres.
• If n = 2, then both the spaces CP3 and S2 × S4 have one cell in each
of the dimensions {0, 2, 4, 6}. Thus they also have the same cellular
chain/cochain complex and, in particular, their homology/cohomol-
ogy groups are isomorphic. We will, however, distinguish these
spaces by their cohomology rings.
We will now show that for n > 1 the two spaces have non-isomorphic
cohomology rings. First, the Künneth formula yields that:
H ∗ (S2 × S4 × · · · × S2n ; Z)
∼
= H ∗ (S2 ; Z) ⊗Z H ∗ (S4 ; Z) ⊗Z · · · ⊗Z H ∗ (S2n ; Z)
Example 7.3.14. Let us use cup products and the Künneth formula in
order to show that Sn ∨ Sm is not a retract of Sn × Sm , for n, m ≥ 1.
First, consider the product CW structure on Sn × Sm : it consists of cells
{e0 , em , en , em+n } with attaching maps ϕ : ∂em → e0 and ϕ′ : ∂en → e0
coming from the factors. Hence Sn ∨ Sm is a subset of Sn × Sm . (Note
that we also allow the case n = m.) Next, suppose by contradiction that
there is a retract
r : Sn × Sm → Sn ∨ Sm .
r ∗ : H ∗ (Sn ∨ Sm ) −→ H ∗ (Sn × Sm )
cup product in cohomology 153
( a × 1) ∪ (1 × b) = ×( a ⊗ 1) ∪ ×(1 ⊗ b)
= ×[( a ⊗ 1) · (1 ⊗ b)]
= ×( a ⊗ b) (7.3.6)
= a×b
̸= 0,
since a ⊗ b ̸= 0 in H ∗ (Sn ) ⊗ H ∗ (Sm ). We also have a ring isomorphism
e ∗ (Sn ∨ Sm ) ∼
H =He ∗ (Sn ) ⊕ H
e ∗ ( S m ).
Then we have:
(d ◦ d)( a ⊗ b) = d (∂a) ⊗ b + (−1) p ( a ⊗ ∂′ b)
following question:
Question 7.3.15. How is the homology H∗ ((C ⊗ C ′ )• ) related to H∗ (C• )
and H∗ (C•′ )?
The answer is provided by the following result from homological
algebra:
Theorem 7.3.16 (Künneth exact sequence). Let R be a PID, and assume
that for each i, Ci is a free R-module. Then for all n, there is a split short exact
sequence:
Hn ( X; R) ∼
= Hn ( X ) ⊗ R ⊕ Tor( Hn−1 ( X ), R). (7.3.13)
Remark 7.3.19. Note that (7.3.13) can also be obtained from (7.3.10) by
taking Y to be a point.
Proof. (Sketch.) Let us indicate how this result is obtained from Theo-
rem 7.3.16. We would like to apply the Künneth exact sequence to the
chain complexes defined by:
and
′ n ′ n
C− n : = C (Y; R ), ∂−n : = δY .
However, note that Ci and Ci′ are not necessarily R-free. Indeed,
Exercises
1. Are the spaces S2 × RP4 and S4 × RP2 homotopy equivalent? Justify
your answer!
2. Using cup products, show that every map Sk+l → Sk × Sl induces the
trivial homomorphism Hk+l (Sk+l ) → Hk+l (Sk × Sl ), assuming k > 0
and l > 0.
8
Poincaré Duality
8.1 Introduction
In particular, we get:
(8.1.1) (UCT )
Hi ( M; Q) ∼
= H n−i ( M; Q) ∼
= Hom( Hn−i ( M; Q), Q) (8.1.2)
Hi ( M; Q) × Hn−i ( M; Q) → Q.
β i ( M ) = β n −i ( M ).
(1)
Hi ( M, M \ { x }; Z) ∼
= Hi (Ux , Ux \ { x }; Z)
(2)
∼
= Hi (Rn , Rn \ {0}; Z)
(3)
∼ e i −1 (Rn \ { 0 } ; Z )
= H (8.2.1)
(4)
∼ e i −1 ( S n −1 ; Z )
= H
Z, if i = n
=
0 , otherwise,
a deformation retract.
∼
=
µ x ∈ Z = Hn ( M, M \ { x }) ←
− Hn ( M, M \ B)
∼
=
→ Hn ( M, M \ {y}) = Z ∈ µy ,
−
ρK : Hi ( M, M \ L) → Hi ( M, M \ K ).
Theorem 8.2.7. For any oriented manifold M of dimension n and any compact
K ⊂ M, there is a unique µK ∈ Hn ( M, M \ K; Z) such that ρ x (µK ) = µ x
for all x ∈ K.
(i) Hi ( M, M \ K ) = 0 if i > n.
Before proving the above lemma, let us finish the proof of Theorem
8.2.7.
Step II: If the theorem holds for compact subsets K1 and K2 and for their
intersection K1 ∩ K2 , we show that it holds for their union K = K1 ∪ K2 .
Indeed, the Mayer-Vietoris sequence for the open cover
M \ ( K1 ∩ K2 ) = ( M \ K1 ) ∪ ( M \ K2 ) ,
with intersection
M \ K = ( M \ K1 ) ∩ ( M \ K2 )
φ
0 → Hn ( M, M \ K ) −
→ Hn ( M, M \ K1 ) ⊕ Hn ( M, M \ K2 )
ψ
−
→ Hn ( M, M \ (K1 ∩ K2 )) → . . .
where φ( a) = ρK1 ( a) ⊕ ρK2 ( a) and ψ(b ⊕ c) = ρK1 ∩K2 (b) − ρK1 ∩K2 (c).
By our assumption, there exist unique µK1 ∈ Hn ( M, M \ K1 ) and µK2 ∈
Hn ( M, M \ K2 ) restricting to local orientations at points x ∈ K1 and,
resp., x ∈ K2 , hence
ρ x ◦ ρ K1 ∩ K2 ( µ K i ) = ρ x ( µ K i ) = µ x (8.2.3)
µK ∈ Hn ( M, M \ K )
such that φ(µK ) = µK1 ⊕ µK2 . By the uniqueness part, we also have that
µK restricts to local orientations at points x ∈ K.
Hi ( M, M \ K ) ∼
= Hi ( M, M \ { x })
∼
= Hi (Rn , Sn−1 )
∼
=He i −1 ( S n −1 ) (8.2.6)
Z for i = n
=
0 otherwise.
Step II: We next show that if the Lemma holds for compact sets K1 ,
K2 and for their intersection K1 ∩ K2 , then it holds for K := K1 ∪ K2 .
Indeed, we have the Mayer-Vietoris sequence
· · · → Hi+1 ( M, M \ (K1 ∩ K2 )) → Hi ( M, M \ K )
φ ψ
−
→ Hi ( M, M \ K1 ) ⊕ Hi ( M, M \ K2 ) −
→ Hi ( M, M \ (K1 ∩ K2 )) → · · ·
φ
0 → Hn ( M, M \ K ) −
→ Hn ( M, M \ K1 ) ⊕ Hn ( M, M \ K2 )
ψ
−
→ Hn ( M, M \ (K1 ∩ K2 )) → . . .
a = 0 ⇐⇒ 0 = φ( a) = ρK1 ( a) ⊕ ρK2 ( a)
⇐⇒ ρK1 ( a) = 0 and ρK2 ( a) = 0
⇐⇒ ρ x ρK1 ( a) = 0 ∀ x ∈ K1 , and ρy ρK2 ( a) = 0 ∀y ∈ K2 (8.2.7)
(since, by assumption, the lemma holds for K1 and K2 )
⇐⇒ ρ x ( a) = 0, ∀ x ∈ K1 ∪ K2 .
ρK
Hi (Rn , Rn \ N ) - Hi (Rn , Rn \ K )
-
ρ∪
iB
i - ρK
Hi (Rn , Rn \ ∪i Bi )
a = ρK ( a′ ) = ρK (ρ∪i Bi ( a′ )) = 0.
Hi ( M, M \ K ) ∼
= Hi (Rn , Rn \ K ). (8.2.8)
Step VI: Finally, note that any compact subset K of M can be written
as a union K = K1 ∪ K2 ∪ . . . ∪ Kr with each Ki as in Step V. Then the
Lemma follows by using Step V, Step II and induction.
Exercises
1. Show that every covering space of an orientable manifold is an
orientable manifold.
(a) Show that if M1 and M2 are closed then there are isomorphisms
Cci ( X ) := Ci ( X, X \ K ) ⊂ Ci ( X ).
[
(8.3.1)
K compact in X
Equivalently,
Cci ( X ) = { φ : Ci ( X ) → Z | ∃ compact K φ ⊂ X
s.t. φ = 0 on chains in X \ K φ }.
δφ(σ) := φ(∂σ),
Hci ( X ) ∼
= lim H i ( X, X \ K ) (8.3.3)
−→
K∈ I
where I := {K ⊂ X | K compact}.
where the direct limit is over the directed set of compact subsets of Rn .
Note that it suffices to let K range over closed balls Bk of integer radius
k centered at the origin since each compact K ⊂ Rn is contained in such
a ball. So we have that
H n (Rn , Rn \ Bk ) ∼
= H n (Rn , Rn \ Bk+1 )
Altogether,
Hci (Rn ) ∼
= lim H i (Rn , Rn \ K ) = lim H i (Rn , Rn \ Bk )
−→ −→
k ∈Z≥0
Z if i = n
=
0 otherwise.
Remark 8.3.7. It follows from the previous example that the cohomol-
ogy with compact support Hc∗ (−) is not a homotopy invariant.
b = X ∪ x̂ be the one point compactification of X.
Remark 8.3.8. Let X
Then
Hci ( X ) ∼ b x̂ ) ∼
= H i ( X, =He i (X
b ). (8.3.4)
For example, Hci (Rn ) ∼=H e i (Sn ). This follows from the following gen-
eral fact. If U is an open subset of a topological space V, with closed
complement Z := V \ U, then there exists a long exact sequence for the
cohomology with compact support
· · · → Hci ( X ) → Hci ( X
b ) → Hci ( x̂ ) → · · ·
where a ∈ C n−i ( X ).
Remark 8.4.2. In view of the definition of the cup product, one can
reformulate the above definition of the cap product as follows: if
σ : ∆n → X is an n-simplex and b ∈ Ci ( X ), then
a ∩ (b ∩ ξ ) = ( a ∪ b) ∩ ξ.
a (∂(b ∩ ξ )) = δa(b ∩ ξ )
= (δa ∪ b)(ξ )
= δ( a ∪ b) − (−1)n−i−1 a ∪ δb (ξ )
= ( a ∪ b)(∂ξ ) − (−1)n−i−1 a(δb ∩ ξ )
= a(b ∩ ∂ξ ) + (−1)n−i a(δb ∩ ξ ).
a ∩ (b ∩ ξ ) = ( a ∪ b) ∩ ξ.
Hence the cap product makes the homology H∗ ( X ) a module over the ring
H∗ ( X ).
poincaré duality 167
The following result states that the cap product ∩ is functorial. Its
proof is a direct consequence of the definition of cap products and is
left as an exercise:
φ ∩ f ∗ ξ = f ∗ (( f ∗ φ) ∩ ξ ) (8.4.7)
∩
H i (Y ) ⊗ Hn (Y ) −−−−→ Hn−i (Y )
∩
H i ( M, M \ K ) ⊗ Hn ( M, M \ K ) −−−−→ Hn−i ( M)
i∗ (µ L ) = µK . (8.4.9)
Proof. The claim follows from the commutativity of the following di-
agram and the uniqueness of µK in Hn ( M, M \ K ) which restricts to
local orientations µ x , ∀ x ∈ K.
µK ∈ Hn ( M, M \ K ) Hn ( M, M \ x )
i∗
µ L ∈ Hn ( M, M \ L)
(i ∗ φ ) ∩ µ L = φ ∩ i ∗ ( µ L ) = φ ∩ µ K , (8.4.10)
Hci ( M) ∼
= lim H i ( M, M \ K ), (8.4.11)
−→
K
where the direct limit on the right-hand side is taken over all compact
subsets K of M. We can now define the Poincaré duality map
∩
Hci ( M) −→ Hn−i ( M) (8.4.12)
i∗
H i ( M, M \ K ) H i ( M, M \ L)
∩µK ∩µ L
Hn−i ( M)
We have now all the necessary ingredients to formulate and prove the
main theorem of this chapter:
poincaré duality 169
φ ∈ Hci ( M ) ∼
= lim H i ( M, M \ K ),
−→
K⊂X
K −compact
and
Z if i = n
Hn−i (Rn ) ≃
0 otherwise.
H n (Rn , Rn \ Bk ) ∼
= Hom( Hn (Rn , Rn \ Bk ); Z).
1 = ak (µ Bk ) = (1 ∪ ak )(µ Bk ) = 1( ak ∩ µ Bk ).
∩µ Bk : H n (Rn , Rn \ Bk ) → H0 (Rn )
Step II: Assuming the theorem holds for opens U, V ⊂ M and for their
intersection U ∩ V, we show that it holds for the union U ∪ V.
For this purpose, we construct a commutative diagram
· · · → H i ( M, M \ (K ∩ L)) → H i ( M, M \ K ) ⊕ H i ( M, M \ L)
→ H i ( M, M \ (K ∪ L)) → · · ·
· · · → H i (U ∩ V, U ∩ V \ K ∩ L) → H i (U, U \ K ) ⊕ H i (V, V \ L)
→ H i (U ∪ V, U ∪ V \ K ∪ L) → · · ·
Taking direct limits over K ⊂ U and L ⊂ V, we get the top long exact
sequence in (8.5.1):
Hi ( M ) = lim
−
→
Hi (Uα ) (8.5.2)
α
and
Hci ( M ) = lim
−
→
Hci (Uα ). (8.5.3)
α
This claim and Poincaré duality for each Uα imply the Poincaré dual-
ity isomorphism for M, since the direct limit of isomorphisms is an
isomorphism. In order to prove the claim, we note that the inclusions
iα : Uα ,→ M induce homomorphisms iα∗ : Hi (Uα ) → Hi ( M) so that
for Uα ,→ Uβ the following diagram commutes:
poincaré duality 171
Hi (Uα ) Hi (Uβ )
Hi ( M )
f : lim
−
→
Hi (Uα ) → Hi ( M).
α
Step IV: We next show that the theorem holds when M is an open
subset of Rn .
If M is convex, then M is homeomorphic to Rn , so the theorem holds
S
by Step I. If M is not convex, then M = k∈Z>0 Vk , with each Vk open
and convex in Rn . By induction and Step II, the theorem holds for the
sets Uk = V1 ∪ · · · ∪ Vk . Note that {Uk }k forms a nested cover of opens
for M, hence the theorem follows by Step III.
form a nested open cover of M. So by Step III, it suffices to show that the
theorem holds for each Uα . But Uα = ∪ β<α Vβ , and the theorem holds
for each Vβ by Step IV. By Step II, Step III, and transfinite induction,
the theorem holds for each Uα , and the claim follows.
172 algebraic topology
defined by the cap product with the fundamental class of M, that is, φ 7→
φ ∩ [ M], is an isomorphism for all i.
Exercises
1. Show that if Mn is a connected, non-compact manifold, then
Hi ( M; Z) = 0 for i ≥ n.
(i) Show that the cup product pairing is nonsingular in the following
sense: for each choice of a Z-basis { β 1 , · · · , β r } of the free abelian
group H n−i ( M; Z)/Torsion, there exists a Z-basis {α1 , · · · , αr } of
H i ( M; Z)/Torsion such that (αi , β j ) = δij . (Hint: Use the Universal
Coefficient Theorem and Poincaré Duality.)
(ii) As an application, re-prove the following facts about the ring struc-
tures on the cohomology of projective spaces:
(a) H ∗ (RPn ; Z2 ) ∼
= Z2 [ x ]/( x n+1 ), | x | = 1,
(b) H ∗ (CPn ; Z) ∼
= Z [ y ] / ( y n +1 ), |y| = 2,
(c) H ∗ (HPn ; Z) ∼
= Z [ w ] / ( w n +1 ), |w| = 4.
poincaré duality 173
( , ) : H 2n ( M; Z)/Torsion ⊗ H 2n ( M; Z)/Torsion → Z
f ∗ : Hn (Sn ; Z) → Hn ( M; Z)
χ( M) = 0.
Proof. Let n = 2k + 1.
If M is oriented, then (with Z-coefficients):
( P.D.) (UCT )
rkHi ( M ) = rkH n−i ( M) = rkHn−i ( M).
So:
2k +1 k
χ( M) = ∑ (−1)i · rkHi ( M) = ∑ (−1)i + (−1)n−i · rkHi ( M) = 0.
i =0 i =0
2k +1
χ( M) := ∑ (−1)n · rkHi ( M; Z)
n =0
(∗) 2k+1
= ∑ (−1)i · dimZ/2 Hi ( M; Z/2)
n =0
= 0,
• a Z-summand of Hi ( M; Z) contributes
Torsion( Hn−1 ( M )) = 0.
Proof. Indeed,
( P.D.)
Torsion( Hi ( M )) = Torsion( H n−i ( M ))
(UCT )
= Ext( Hn−1−i ( M), Z)
= Torsion( Hn−1−i ( M))
Proof. (Sketch)
Define
e := {µ x | x ∈ M, µ x a local orientation of M at x }
M
U (µ B ) = {µ x ∈ M
e | x ∈ B, µ x = ρ x (µ B )},
Hn ( M,
e Me \ µx ) ∼
= Hn (U (µ B ), U (µ B ) \ µ x ) ∼
= Hn ( B, B \ x )
(8.7.1)
∼ Hn ( M, M \ x ).
=
So at the point µ x ∈ M
e there exists a canonical local orientation
ex ∈ Hn ( M,
µ e Me \ µx ) ∼
=Z
(b) The oriented double cover of the Klein bottle K is the 2-torus T 2 .
Proof. The oriented double cover M e can have one or two components.
If M has two components, each is oriented and homeomorphic to M,
e
so M is orientable. Conversely, if M is orientable, it can have exactly
two orientations at each point, each defining a sheet of M.
e
MZ = {α x | x ∈ M, α x ∈ Hn ( M, M \ x ) = Z}.
πZ : MZ → M
U ( B) = {α x | x ∈ B, α x = ρ x (α B ) for α B ∈ Hn ( M, M \ B) ∼
= Z)}
∼
=
with ρ x : Hn ( M, M \ B) → Hn ( M, M \ x ) induced by inclusion as before.
For any k ∈ Z, we then get a subcover Mk ⊂ MZ by selecting ±kµ x in
the fibre above x. So
[
MZ = Mk
k ≥0
with M0 ∼
= M, Mk ∼
= M−k , and Mk ∼
= M,
e for any integer k.
Hn ( M, M \ x; R) ∼
= Hn ( M, M \ x ) ⊗ R ∼
= Z⊗R ∼
= R.
MR = {α x | x ∈ M, α x ∈ Hn ( M, M \ x; R) ∼
= R }.
We are now ready to prove the following result, which shows that
orientability of a closed manifold is reflected in the structure of its
homology:
The proof of Theorem 8.7.9 is based on the Theorem 8.2.7 and Lemma
8.2.10 (which we formulate here with R-coefficients in parts (a) and (b)
below), together with a slight generalization of Theorem 8.2.7 (see part
(c) below) which holds without the orientability assumption:
Note that the proof of part (c) of the above lemma is almost identical
to that of Theorem 8.2.7 (with the uniqueness following from part (b)),
with the only easy modification appearing in Step I of loc.cit. (where the
orientation assumption used in the proof of Theorem 8.2.7 is replaced
by the continuity of the section). We leave the details to the reader.
defined by
α → ( x 7 → α x ),
Otherwise,
Torsion( Hn−1 ( M )) = Z/2.
Remark 8.7.13. By using the universal coefficient theorem for the coho-
mology of a closed n-manifold, we have:
Indeed,
φ(ψ ∩ σ ) = ψ(σ |[vl ,vl +1 ,...,vk+l ] ) · φ(σ |[v0 ,v1 ,...,vl ] ) = ( φ ∪ ψ)(σ). (8.8.3)
h
H l ( M; R) −−−−→ HomR ( Hl ( M; R), R)
∪ψy (ψ∩)∗ y
h
H k+l ( M; R) −−−−→ HomR ( Hk+l ( M; R), R)
poincaré duality 181
∩[ M]
H k ( M; R) × H n−k ( M; R) −→ H n ( M; R) −→ H0 ( M; R) = R (8.8.4)
is defined by
( φ, ψ) 7→ ( φ ∪ ψ) 7→ ( φ ∪ ψ) ∩ [ M].
h ( P.D.)∗
f : H k ( M; R) −→ HomR ( Hk ( M; R), R) −→ HomR ( H n−k ( M; R), R),
1 = φ(α) = (α ∪ β)[ M ].
4n+2
χ( M) = ∑ (−1)i · rk( Hi ( M)).
i =0
Therefore,
χ( M ) ≡ rk( H2n+1 ( M )) (mod 2).
poincaré duality 183
∪ ∩[ M]
H 2n+1 ( M ) × H 2n+1 ( M ) → H 4n+2 ( M) −→ Z
defined by
(α, β) 7→ (α ∪ β) 7→ (α ∪ β) ∩ [ M].
By Poincaré Duality, after moding out by torsion, this pairing is non-
singular. As a result, the matrix A of the cup product pairing is
non-singular and anti-symmetric. By linear algebra, A is similar to a
matrix with diagonal blocks
!
0 −1
−1 0
Therefore,
rk( H 2n+1 ( M )) = rk( A),
which is clearly even.
Hk ( M ) × Hn−k ( M) → Z
defined by
([σ], [η ]) → ♯(σ ∩ η ′ ),
where η ′ is chosen so that it is homologous to η but transversal to σ (so
σ ∩ η ′ is a finite number of points).
Example 8.8.9. Let T be the 2-dimensional torus and S be a meridian
of T. Let M be the pinched torus T/S.
Exercises
1. Let Mg be a closed orientable surface of genus g ≥ 1. Show that for
each non-zero α ∈ H 1 ( M; Z) there exists β ∈ H 1 ( M; Z) with α ∪ β ̸= 0.
Deduce that M is not homotopy equivalent to a wedge sum X ∨ Y of
CW-complexes with non-trivial reduced homology. Do the same for
closed nonorientable surfaces using cohomology with Z2 -coefficients.
(a) if Ux ∼
= Rn , then Hn ( M, M \ x ) ∼
= Hn (Ux , Ux \ x ) ∼
= Z.
(b) if Ux ∼
= Rn+ , then
Hn ( M, M \ x ) ∼
= Hn (Ux , Ux \ x ) ∼
= Hn (Rn+ , Rn+ − {0}) ∼
= 0.
∂M := { x ∈ M | Hn ( M, M \ x ) = 0}.
and
∩µ M
Hci ( M, ∂M) −−
∼
→ Hn−i ( M) (8.9.2)
=
where Hci ( M, ∂M ) := lim Kcompact H i ( M, ( M \ K ) ∪ ∂M ) is the cohomology
−→
K ⊂ M \∂M
with compact support for the pair ( M, ∂M ).
Let us now describe some applications of Poincaré duality for mani-
folds with boundary.
Proof of Proposition 8.9.10. Start with the cohomology long exact se-
quence for the pair (V, M ):
i∗ δ
H n (V; R) - H n ( M; R) - H n+1 (V, M; R)
∼
= ∩[ M] ∼
= ∩[V ]
? i∗ - ?
Hn ( M; R) Hn (V; R)
where i∗ , i∗ are induced by the inclusion i : M = ∂V ,→ V. By exactness,
P.D.
we have that Im i∗ ∼
= ker δ ∼
= ker i∗ , so
is even.
(b) dim(Im i∗ ) = 1
2 dim H n ( M2n ; R).
δ
H 2n ( M; R) - H 2n+1 (V, M; R)
∼
= P.D. ∼
= P.D.
? ?
H0 ( M; R) - H0 (V; R)
Exercises
1. Let X be the cone on CPn . Show that X is a manifold with boundary
if and only if n = 1.
poincaré duality 187
Signature
Definition 8.9.12. Let M be a closed oriented manifold. If dim M = 4k,
the signature σ ( M ) of M is defined to be the signature of the symmetric
non-singular cup product pairing
H 2k ( M; R) × H 2k ( M; R) −→ R
(α, β) 7→ (α ∪ β)[ M]
σ(CP2n ) = 1,
σ(CP2 #CP2 ) = 2.
The signature σ is a cobordism invariant, i.e., if ∂W = M ⊔ − N, then
σ( M) = σ( N ). Here − N denotes the manifold N but with the opposite
orientation.
σ ( M) = r − (2l − r ) = 2r − 2l.
188 algebraic topology
In conclusion, r = l and σ( M) = 0.
Connected Sums
Definition 8.9.16. Let Mn , N n be closed, connected, oriented n-manifolds.
Their connected sum is defined to be
Example 8.9.19. The spaces S2 × S2 and CP2 #CP2 have the same coho-
mology groups,
H 0 = Z, H 2 = Z ⊕ Z = Zα ⊕ Zβ, H 4 = Z,
σ(CP2 #CP2 ) = 2 ̸= 0,
(b) Run an orientation reversing path γ from one CP2 to the other, by
traveling along an orientation reversing path in RP2 .
(c) Enlarge the path to a tube and remove its interior. What is left is a
5-dimensional non-orientable manifold with ∂W = CP2 #CP2 .
basics of homotopy theory 191
9
Basics of Homotopy Theory
f : ( I n , ∂I n ) → ( X, x0 ) / ∼
πn ( X, x0 ) =
(
( I n /∂I n , ∂I n /∂I n )
with ( I n /∂I n , ∂I n /∂I n ) ≃ (Sn , s0 ). So we can also define
πn ( X, x0 ) = g : (Sn , s0 ) → ( X, x0 ) / ∼ .
Proof. First note that since only the first coordinate is involved in this
operation, the same argument used to prove that π1 is a group is valid
192 algebraic topology
I n −1 Figure 9.1: f + g
f g
0 1/2 1 s1
here as well. Then the identity element is the constant map taking all
of I n to x0 and the inverse element is given by
− f ( s1 , s2 , . . . , s n ) = f (1 − s1 , s2 , . . . , s n ).
Figure 9.2: f + g ≃ g + f
f
f g ≃ f g ≃ ≃ g f
g
≃ g f
β γ ([ f ]) = [γ̄ ∗ f ∗ γ]
β γ ([ f ]) = [γ · f ],
x0 Figure 9.4: β γ
x1
x0 x1 f x1 x0
x1
x0
1. γ · ( f + g) ≃ γ · f + γ · g
π1 × π n → π n
(γ, [ f ]) 7→ [γ · f ]
1. (ϕ ◦ ψ)∗ = ϕ∗ ◦ ψ∗ .
( X,
e xe)
:
ϕet
p
/ ( X, x )
ϕt
( S n , s0 )
for all n.
196 algebraic topology
πn ( X ) = πn (Y ) for all n.
By considering homology groups, however, we see that X and Y are
not homotopy equivalent. Indeed, by the Künneth formula, we get that
H5 ( X ) = Z while H5 (Y ) = 0 (since RP3 is oriented while RP2 is not).
h X : πn ( X ) −→ Hn ( X )
defined by
[ f : S n → X ] 7 → f ∗ [ S n ],
where [Sn ] is the fundamental class of Sn . A very important result in
homotopy theory is the following:
I n−1 = {(s1 , . . . , sn ) ∈ I n | sn = 0}
and set
J n−1 = ∂I n \ I n−1 .
Then define the n-th homotopy group of the pair ( X, A) with basepoint x0 as:
f : ( I n , ∂I n , J n−1 ) → ( X, A, x0 ) / ∼
πn ( X, A, x0 ) =
sn J n −1
I n −1
f
( I n , ∂I n , J n−1 ) / ( X, A, x0 )
7
g
(
( D n , S n −1 , s 0 )
πn ( X, A, x0 ) = g : ( D n , Sn−1 , s0 ) → ( X, A, x0 ) / ∼ .
x0
x0 S n −1
A
x0 x0
Dn
Remark 9.2.3. Note that the proposition fails in the case n = 1. Indeed,
we have that
π1 ( X, A, x0 ) = f : ( I, {0, 1}, {1}) → ( X, A, x0 ) / ∼ .
X
A
x0
i
∗ j∗ n ∂
· · · → πn ( A, x0 ) → πn ( X, x0 ) → πn ( X, A, x0 ) −→ πn−1 ( A, x0 ) → · · ·
· · · → π0 ( X, x0 ) → 0,
where the map ∂n is defined by ∂n [ f ] = [ f | I n−1 ] and all others are induced by
inclusions.
Remark 9.2.5. Near the end of the above sequence, where group struc-
tures are not defined, exactness still makes sense: the image of one map
is the kernel of the next, which consists of those elements mapping to
the homotopy class of the constant map.
πn (CX, X, x0 ) ∼
= πn−1 ( X, x0 ).
x0
g
x0 A x0 X
γ A γ
3. πi ( X, x0 ) = 0 for all x0 ∈ X.
In the relative setting, the following are equivalent for any i > 0:
4. πi ( X, A, x0 ) = 0 for all x0 ∈ A.
Xn × I = Xn × {0} ∪ ( An ∪ Xn−1 ) × I ∪ D n × I,
rn : Xn × I → Xn × {0} ∪ ( An ∪ Xn−1 ) × I.
Concatenating these deformation retractions by performing rn over 1 −
1 1
n − 1
, 1 − n , we get a deformation retraction of X × I onto X × {0} ∪
2 2
A × I. Continuity follows since CW complexes have the weak topology
with respect to their skeleta, so a map of CW complexes is continuous
if and only if its restriction to each skeleton is continuous.
The assignment
[ f ] ∈ πi ( X ) 7→ [Σ f ] ∈ πi+1 (ΣX )
defines a homomorphism πi ( X ) → πi+1 (ΣX ), which is an isomorphism for
i < 2n − 1 and a surjection for i = 2n − 1.
πi ( X ) ∼
= πi+1 (C+ , X ) −→ πi+1 (ΣX, C− X ) ∼
= πi+1 (ΣX ),
The assignment
[ f ] ∈ π i ( S n ) 7 → [ Σ f ] ∈ π i +1 ( S n +1 )
Z∼
= π1 ( S1 ) ↠ π2 ( S2 ) ∼
= π3 ( S3 ) ∼
= π4 ( S4 ) ∼
= ···
To show that π1 (S1 ) ∼ = π2 (S2 ), we can use the long exact sequence for
the homotopy groups of a fibration, see Theorem 9.11.8 below. For any
fibration (e.g., a covering map)
F ,→ E −→ B
· · · −→ πi ( F ) −→ πi ( E) −→ πi ( B) −→ πi−1 ( F ) −→ · · · (9.6.1)
π3 ( S2 ) ∼
= π2 ( S2 ) ∼
= Z.
Remark 9.6.4. Unlike the homology and cohomology groups, the ho-
motopy groups of a finite CW-complex can be infinitely generated. This
fact is discussed in the next example.
−1
with Skn
denoting the n-sphere corresponding to the integer k. So for
any n ≥ 2, we have:
π n ( S1 ∨ S n ) = π n ( Skn ),
_
k ∈Z
Proof. Suppose first that there are only finitely many Sαn ’s in the wedge
W n W n n
α Sα . Then we can regard α Sα as the n-skeleton of ∏α Sα . The cell
n 0
structure of a particular Sα consists of a single 0-cell eα and a single
n-cell, eαn . Thus, in the product ∏α Sαn there is one 0-cell e0 = ∏α e0α ,
which, together with the n-cells
( ∏ e0β ) × eαn ,
[
α β̸=α
form the n-skeleton α Sαn . Hence ∏α Sαn \ α Sαn has only cells of di-
W W
mension at least 2n, which by Corollary 9.4.8 yields that the pair
(∏α Sαn , α Sαn ) is (2n − 1)-connected. In particular, as n ≥ 2, we get:
W
πn ( Sαn ) ∼ = πn ∏ Sαn ∼ = ∏ πn (Sαn ) = πn (Sαn ) = Z.
_ M M
α α α α α
πn ∼
= Z[ π1 ] ∼
= Z[Z] ∼
= Z[t, t−1 ],
1 7→ t
− 1 7 → t −1
n 7→ tn ,
The above corollary follows from Whitehead’s theorem and the fol-
lowing relative version of the Hurewicz Theorem 9.10.1 (to be discussed
later on):
f : X = S1 ,→ (S1 ∨ Sn ) ∪ en+1 = Y ( n ≥ 2)
be the inclusion map, with the attaching map for the (n + 1)-cell of
Y described below. We know from Example 9.6.5 that πn (S1 ∨ Sn ) ∼ =
Z[t, t−1 ]. We define Y by attaching the (n + 1)-cell en+1 to S1 ∨ Sn by a
map g : Sn = ∂en+1 → S1 ∨ Sn so that [ g] ∈ πn (S1 ∨ Sn ) corresponds to
the element 2t − 1 ∈ Z[t, t−1 ]. We then see that
f ≃ f 1 , with f 1 ( X1 ∪ A) ⊆ B,
H1
f 1 ≃ f 2 , with f 2 ( X2 ∪ A) ⊆ B,
H2
···
f n−1 ≃ f n , with f n ( Xn ∪ A) ⊆ B,
Hn
and so on. Any finite skeleton is eventually fixed under these homo-
topies.
Define a homotopy H : X × I → Y as
H = Hi on 1 − 2i1−1 , 1 − 21i .
M f := ( X × I ) ⊔ Y/( x, 1) ∼ f ( x ).
f (X)
by the long exact sequence for the homology groups of the pair (Y, X )
that Hn (Y, X ) = 0 for all n.
Since X and Y are simply-connected, we have π1 (Y, X ) = 0. So by
the relative Hurewicz Theorem 9.10.1, the first non-zero πn (Y, X ) is
isomorphic to the first non-zero Hn (Y, X ). So πn (Y, X ) = 0 for all n.
Then, by the homotopy long exact sequence for the pair (Y, X ), we get
that
πn ( X ) ∼
= π n (Y )
for all n, with isomorphisms induced by the inclusion map f . Finally,
Whitehead’s theorem 9.7.2 yields that f is a homotopy equivalence.
= πi (S2 ) × πi (S∞ ), i ≥ 2.
π i (Y ) ∼
πi ( X ) ∼
= π i ( S2 ) ∼
= πi (Y ), i ≥ 2.
9.8 CW approximation
In this section we show that given any space X, there exists a (unique
up to homotopy) CW complex Z and a weak homotopy equivalence
f : Z → X. Such a map f : Z → X is called a CW approximation of X.
(r, f 0 )
( f , f 0 ) : ( Z, Z0 ) −→ ( Mg , Z0 ) −→ ( X, X0 )
πi ( X ) ∼
= πi ( A ) ∼
= πi ( Z ) where the isomorphisms are induced by f
since the following diagram commutes,
f
ZO /X
O
A
id /A
πn ( A) ↠ πn ( Z ) ↣ πn ( X )
W f := M f /{{ a} × I ∼ pt, ∀ a ∈ A}
A = Zn ⊆ Zn+1 ⊆ · · ·
f
( Z, A) −−−−→ ( X, A)
hy g
y
f′
( Z ′ , A′ ) −−−−→ ( X ′ , A′ )
commutes up to homotopy.
and g = id in the above lemma twice, and conclude that there are two
maps h0 : Z → Z ′ and h1 : Z ′ → Z, such that f ◦ h1 ≃ f ′ (rel. A)
and f ′ ◦ h0 ≃ f (rel. A). In particular, f ◦ (h1 ◦ h0 ) ≃ f (rel. A) and
f ′ ◦ (h0 ◦ h1 ) ≃ f ′ (rel. A). The uniqueness in Proposition 9.8.7 then
implies that h1 ◦ h0 and h0 ◦ h1 are homotopic to the respective identity
maps (rel. A).
ZG 2
1
? Z
A / Z0 /X
· · · −→ Z2 −→ Z1 −→ Z0 −→ X
XG 3
2
X
?
X / X1
Y = X ∪ A CA
q : Y −→ Y/CA = X/A
where the first and second maps are induced by the inclusion of pairs,
the second map is an isomorphism by the long exact sequence of the
pair (Y, CA)
n +1
Lemma 9.9.4. Assume n ≥ 2. If X = ( Sαn ) ∪
W S
α β eβ is obtained
n enβ+1
W
from by attaching (n + 1)-cells
α Sα via basepoint-preserving maps
ϕβ : Snβ → α Sαn , then
W
Sαn )/⟨ϕβ ⟩ = ( Z) / ⟨ ϕ β ⟩.
_ M
πn ( X ) = πn (
α α
Proof. Consider the following portion of the long exact sequence for
the homotopy groups of the n-connected pair ( X, α Sαn ):
W
Theorem 9.9.6. For any n ≥ 1 and any group G (which is assumed abelian
if n ≥ 2) there exists an Eilenberg-MacLane space K ( G, n).
∂
· · · → πn+2 ( Xn+2 , Xn+1 ) → πn+1 ( Xn+1 ) → πn+1 ( Xn+2 ) → 0.
Corollary 9.9.7. For any sequence of groups { Gn }n∈N , with Gn abelian for
n ≥ 2, there exists a space X such that πn ( X ) ∼
= Gn for any n.
we have that
f ∗ ([iα ]) = [ f ◦ iα ] = [ f |Sαn ] = ψ([iα ]).
Proof. First, since all hypotheses and assertions in the statement deal
with homology and homotopy groups, if we prove the statement for
a CW approximation of X (or ( X, A)) then the results will also hold
for the original space (or pair). Hence, we assume without loss of
generality that X is a CW complex and ( X, A) is a CW-pair.
Secondly, the relative case can be reduced to the absolute case. In-
deed, since ( X, A) is (n − 1)-connected and that A is 1-connected,
Lemma 9.9.3 implies that πi ( X, A) = πi ( X/A) for i ≤ n, while
Hi ( X, A) = He i ( X/A) always holds for CW-pairs.
In order to prove the absolute case of the theorem, let x0 be a 0-cell
in X. Since X, hence also ( X, x0 ), is (n − 1)-connected, Corollary 9.8.5
tells us that we can replace X by a homotopy equivalent CW complex
with (n − 1)-skeleton a point, i.e., Xn−1 = x0 . In particular, H
ei (X ) = 0
for i < n. For showing that πn ( X ) ∼ = Hn ( X ), we may disregard any
cells of dimension greater than n + 1 since these have no effect on πn
or Hn . Thus we may assume that X has the form ( α Sαn ) ∪ β enβ+1 . By
W S
Recall the Hurewicz Theorem has been already used for proving the
important Corollary 9.7.3. Here we give another important application
of Theorem 9.10.1:
ge0
X /E
?
p
get
X /B
gt
220 algebraic topology
7? E
g
p
fe
A /Z /B
f
Remark 9.11.6. The homotopy lifting property with respect to the pair
( X, A) is the lift extension property for ( X × I, X × {0} ∪ A × I ).
Remark 9.11.7. The homotopy lifting property with respect to a disk
D n is equivalent to the homotopy lifting property with respect to the
pair ( D n , ∂D n ), since the pairs ( D n × I, D n × {0}) and ( D n × I, D n ×
{0} ∪ ∂D n × I ) are homeomorphic. This implies that a fibration has
the homotopy lifting property with respect to all CW pairs ( X, A). Indeed,
the homotopy lifting property for disks is in fact equivalent to the
homotopy lifting property with respect to all CW pairs ( X, A). This
can be easily seen by induction over the skeleta of X, so it suffices to
construct a lifting get one cell of X \ A at a time. Composing with the
characteristic map D n → X of a cell then gives the reduction to the case
( X, A) = ( D n , ∂D n ).
Theorem 9.11.8 (Long exact sequence for homotopy groups of a fibra-
tion). Given a fibration p : E → B, points b ∈ B and e ∈ F := p−1 (b), there
∼
=
is an isomorphism p∗ : πn ( E, F, e) −→ πn ( B, b) for all n ≥ 1. Hence, if B is
path-connected, there is a long exact sequence of homotopy groups:
p∗
· · · −→ πn ( F, e) −→ πn ( E, e) −→ πn ( B, b) −→ πn−1 ( F, e) −→ · · ·
· · · −→ π0 ( E, e) −→ 0
p1 p2
B
commutes. Such a map f is called a fiber homotopy equivalence if f is both
fiber-preserving and a homotopy equivalence, i.e., there is a map g : E2 → E1
such that f and g are fiber-preserving and f ◦ g and g ◦ f are homotopic to
the identity maps by fiber-preserving maps.
Definition 9.11.10 (Fiber Bundle). A map p : E → B is a fiber bundle
with fiber F if, for any point b ∈ B, there exists a neighborhood Ub of b with
a homeomorphism h : p−1 (Ub ) → Ub × F so that the following diagram
commutes:
p−1 (Ub )
h / Ub × F
p pr
" }
Ub
0 1
−1
S1 p = S1 q = ··· = S1 p
! " !
S2n+1 ⊂ S2n+3 ⊂ ... ⊂ S∞
CPn ⊂ CPn+1 ⊂ ... ⊂ CP∞
basics of homotopy theory 223
S1 ,→ S∞ −→ CP∞
i.e.,
CP∞ = K (Z, 2),
as already mentioned in our discussion about Eilenberg-MacLane
spaces.
Remark 9.11.17. As we will see later on, for any topological group
πG
G there exists a “universal fiber bundle” G ,→ EG −→ BG with EG
contractible, classifying the space of (principal) G-bundles. That is, any
G-bundle π : E → B over a space B is determined by (the homotopy
class of) a classifying map f : B → BG by pull-back: π ∼ = f ∗ πG :
E EG ≃ {pt}
π πG
B / BG
f
From this point of view, CP∞ can be identified with the classifying
space BS1 of (principal) S1 -bundles.
Example 9.11.18. By letting n = 1 in the fibration of Example 9.11.16,
the corresponding bundle
S1 ,→ S3 −→ CP1 ∼
= S2 (9.11.1)
is called the Hopf fibration. The long exact sequence of homotopy group
for the Hopf fibration gives: π2 (S2 ) ∼
= π1 (S1 ) and πn (S3 ) ∼
= πn (S2 ) for
∞
all n ≥ 3. Together with the fact that CP = K (Z, 2), this shows that S2
and S3 × CP∞ are simply-connected CW complexes with isomorphic
homotopy groups, though they are not homotopy equivalent as can be
easily seen from cellular homology.
Example 9.11.19. A fiber bundle similar to that of Example 9.11.16 can
be obtained by replacing C with the quaternions H, namely:
S3 ,→ S4n+3 −→ HPn .
(Note that S4n+3 can be identified with the unit sphere in Hn+1 .) In
particular, by letting n = 1 we get a second Hopf fiber bundle
S3 ,→ S7 −→ HP1 ∼
= S4 . (9.11.2)
224 algebraic topology
S7 ,→ S15 −→ S8 (9.11.3)
Here we regard S15 as the unit sphere in the 16-dimensional vector space
O2 , and the projection map S15 −→ S8 = O ∪ {∞} is (z0 , z1 ) 7→ z0 z1−1
(just like for the other Hopf bundles). There are no fiber bundles
with fiber, total space and base spheres, other than those provided by
the Hopf bundles of (9.11.1), (9.11.2) and (9.11.3). Finally, note that
there is an “octonion projective plane” OP2 obtained by glueing a cell
e16 to S8 via the Hopf map S15 → S8 ; however, there is no octonion
analogue of RPn , CPn or HPn for higher n, since the associativity of
multiplication is needed for the relation (z0 , · · · , zn ) ∼ λ(z0 , · · · , zn ) to
be an equivalence relation.
U (n − 1) ,→ U (n) → S2n−1
A 7→ Ax,
p : Vn (Rk ) −→ Gn (Rk )
given by
{v1 , . . . , vn } 7→ span{v1 , . . . , vn }.
Lemma 9.12.6. The projection p is a fiber bundle with fiber Vn (Rn ) = O(n).
To conclude this discussion, we have shown that for k > n, there are
fiber bundles:
/ Vn (Rk ) / Gn (Rk )
O(n) (9.12.1)
A similar method gives the following fiber bundle for all triples
m < n ≤ k:
Vn−m (Rk−m )
p
/ Vn (Rk ) / Vm (Rk ) (9.12.2)
{ v1 , . . . , v n } / { v1 , . . . , v m }
O ( n − 1) / O(n) / S n −1 (9.12.4)
!
/ A 0
A
0 1
B / Bu
with u ∈ Sn−1 some fixed unit vector. In particular, this identifies Sn−1
as an orbit (or homogeneous) space:
= O(n)⧸O(n − 1).
S n −1 ∼
Vn−1 (Rk−1 ) / Vn (Rk ) / S k −1 . (9.12.5)
By using the long exact sequence for bundle (9.12.5) and induction on
n, it follows readily that Vn (Rk ) is (k − n − 1)-connected.
Remark 9.12.9. The long exact sequence of homotopy groups for the
bundle (9.12.4) shows that πi (O(n)) is independent of n for n large. We
call this the stable homotopy group πi (O). Bott Periodicity shows that
πi (O) is periodic in i with period 8. Its values are:
i 1 2 3 4 5 6 7 8
πi (O) Z/2 Z/2 0 Z 0 0 0 Z
Definition 9.12.10.
∞ ∞
Vn (R∞ ) := Vn (Rk ) Gn (R∞ ) := Gn (Rk )
[ [
k =1 k =1
basics of homotopy theory 227
Proof. By using the bundle (9.12.5) for k → ∞, we see that πi (Vn (R∞ )) =
0 for all i. Using the CW structure and Whitehead’s Theorem 9.7.2
shows that Vn (R∞ ) is contractible.
Alternatively, we can define an explicit homotopy ht : R∞ → R∞ by
with Vn (C∞ ) contractible. As we will see later on, this means that
Vn (C∞ ) is the classifying space for rank-n complex vector bundles. We
also have a fiber bundle similar to (9.12.4)
/ U (n) / S2n−1 ,
U ( n − 1) (9.12.9)
= U (n)⧸U (n − 1).
S2n−1 ∼
Many of these fiber bundles will become essential tools in the next
chapter for computing (co)homology of matrix groups, with a view
towards classifying spaces and characteristic classes of manifolds.
A / Ef /9 B
h.e. fibration
i p
f
πi ( B ) ∼
= πi−1 (ΩB) (9.13.1)
for all i.
The isomorphism (9.13.1) suggests that the Hurewicz Theorem 9.10.1
can also be proved by induction on the degree of connectivity. Indeed,
if B is n-connected then ΩB is (n − 1)-connected. We’ll give the details
of such an approach by using spectral sequences.
· · · −→ Ω2 B −→ ΩF −→ ΩE −→ ΩB −→ F −→ E −→ B
9.14 Exercises
f ∗ : [ Z, X ] → [ Z, Y ] and f ∗ : [Y, Z ] → [ X, Z ] ,
5. (Extension Lemma)
Given a CW pair ( X, A) and a map f : A → Y with Y path-connected,
show that f can be extended to a map X → Y if πn−1 (Y ) = 0 for all n
such that X \ A has cells of dimension n.
16. Show that if there were fiber bundles Sn−1 → S2n−1 → Sn for all n,
then the groups πi (Sn ) would be finitely generated free abelian groups
computable by induction, and non-zero if i ≥ n ≥ 2.
17. Let U (n) be the unitary group. Find πk (U (n)) for k = 1, 2, 3 and
n ≥ 2.
10
Spectral Sequences. Applications
E∗r+ 1 r
,∗ = H∗ ( E∗,∗ ).
In more detail, we have abelian groups { Erp,q } and maps (called “differentials”)
+1
Erp,q = Erp,q = · · · = E∞
p,q .
232 algebraic topology
dr
q+r−1 Erp−r,q+r−1
dr
q Erp,q
dr
..
.
0
0 ··· p−r p
( Er , dr ) ⇛ H∗ ,
D0,n
D1,n−1 /D0,n
Dn−1,1 /Dn−2,2
Hn /Dn−1,1
spectral sequences. applications 233
Hn ∼ E∞
M
= p,q ,
p+q=n
• If E∞
p,q = 0, for all p + q = n, then Hn = 0.
• If Hn = 0, then E∞
p,q = 0 for all p + q = n.
converging to H∗ ( E).
Remark 10.1.5. Fix some coefficient group K. Then, since B and F are
connected, we have:
2 = H ( B; H ( F; K)) = H ( F; K)
• E0,q 0 q q
Definition 10.1.6. The spectral sequence of the above theorem shall be referred
to as the Leray-Serre spectral sequence of a fibration, and any ring of coefficients
can be used.
E2p,q = H p ( B; Hq ( F )),
H∗ ( F )
H∗ ( B)
and recall that the path space PX is contractible. Note that the loop
space ΩX is connected, since π0 (ΩX ) ∼= π1 ( X ) = 0. Moreover, since
π1 ( X ) = 0, the Leray-Serre spectral sequence (10.1.1) for the path
fibration has the E2 -page given by
E2p,q = H p ( X, Hq (ΩX )) ⇛ H∗ ( PX ).
π2 ( X ) ∼
= π1 (ΩX ) ∼
= H1 (ΩX ),
where the first isomorphism follows from the long exact sequence of
homotopy groups for the path fibration, and the second isomorphism
is the abelianization since π2 ( X ), hence also π1 (ΩX ), is abelian. So it
remains to show that we have an isomorphism
H1 (ΩX ) ∼
= H2 ( X ). (10.2.2)
Consider the E2 -page of the Leray-Serre spectral sequence for the path
fibration. We need to show that
d2 : E2,0
2 2
= H2 ( X ) → E0,1 = H1 (ΩX )
spectral sequences. applications 235
is an isomorphism.
H∗ (ΩX )
E2
H1 (ΩX )
d2
H∗ ( X )
Z H1 ( X ) H2 ( X )
πn ( X ) ∼
= πn−1 (ΩX ) ∼
= Hn−1 (ΩX ),
where the first isomorphism follows from the long exact sequence
of homotopy groups for the path fibration, and the second is by the
induction hypothesis, as already mentioned. So it suffices to show that
we have an isomorphism
Hn−1 (ΩX ) ∼
= Hn ( X ). (10.2.3)
E2p,q = H p ( X, Hq (ΩX ))
∼
= H p ( X ) ⊗ Hq (ΩX ) ⊕ Tor( H p−1 ( X ), Hq (ΩX ))
=0
H∗ (ΩX )
Hn−1 (ΩX )
.. dn
.
H∗ ( X )
...
Z H1 ( X ) H2 ( X ) Hn−1 ( X ) Hn ( X )
In this section, we give some more details about the Leray-Serre spectral
sequence. We begin with some general considerations about spectral
sequences.
Start off with a chain complex C∗ with a bounded increasing filtration
F C∗ , i.e., each F p C∗ is a subcomplex of C∗ , F p−1 C∗ ⊆ F p C∗ for any p,
•
( F p C∗ , F p−1 C∗ , F p−2 C∗ ).
Moreover, we have
Theorem 10.3.1.
E∞ p
p,q = F H p+q (C∗ ) /F
p −1
H p+q (C∗ )
F p C∗ ( E) := C∗ (π −1 ( B p )),
By excision,
ep
H∗ (π −1 (e p ), π −1 (∂e p )) ∼
= H∗ (e p × F, ∂e p × F )
∼
= H∗ ( D p × F, S p−1 × F )
∼
= H∗− p ( F )
∼
= H p ( D p , S p−1 ; H∗− p ( F )),
map of the long exact sequence of the triple ( B p , B p−1 , B p−2 ). By cellular
homology, this is exactly a description of the boundary map of the CW-
chain complex of B with coefficients in Hq ( F ), hence
E2p,q = H p ( B, Hq ( F )).
238 algebraic topology
E2p,q = H p ( B; Hq ( F )),
i π
Theorem 10.3.3. Let F ,→ E → B be a fibration with π1 ( B) = 0 (or π1 ( B)
acts trivially on H∗ ( F )) and π0 ( E) = 0. Then, there is a first quadrant
spectral sequence with E2 -page
E2p,q = H p ( B, Hq ( F ))
such that E∞
p,q = D p,q /D p−1,q+1 .
n-th diagonal of E∞
D0,n
D1,n−1 /D0,n
Dn−1,1 /Dn−2,2
Hn ( E)/Dn−1,1
spectral sequences. applications 239
p p +1
H p ( B) = E2p,0 ⊇ ker d2p,0 = E3p,0 ⊇ ker d3p,0 = E4p,0 ⊇ . . . ⊇ ker d p,0 = E p,0
[ O
=
..
.O
E∞
p,0
O
π∗ =
H p ( E)/D p−1,1 H p ( E)
O
onto
H p ( E)
H p ( E) ↠ E∞ 2
p,0 ⊆ E p,0 = H p ( B ), (10.3.1)
2
Hq ( F ) = E0,q / / E0,q
3
= Hq ( F )/Im(d2 ) // ... / / E q +2
0,q
=
..
.
=
∞
E0,q
i∗
=
D0,q H q ( E)
_
)
Hq ( E)
(c)
hnB
πn ( B) / Hn ( B) = E2 ⊇ . . . ⊇ n
En,0
n,0
l.e.s. ∂ dn
hnF−1
π n −1 ( F ) / Hn−1 ( F ) = E2 // ... / / En
0,n−1 0,n−1
hnX
πn ( X ) / Hn ( X ) = E2 = . . . = En
n,0 n,0
∼
= ∂ ∼
= dn
∼
πn−1 (ΩX )
= / Hn−1 (ΩX ) = E2 = . . . = E n
−1 0,n−1 0,n−1
hnΩX
−1
The Hurewicz homomorphism hnΩX is an isomorphism by the inductive
hypothesis, ∂ is an isomorphism by the homotopy long exact sequence
associated to the path fibration for X, and dn is an isomorphism by the
spectral sequence argument used in the proof of the Hurewicz theorem.
Therefore, hnX : πn ( X ) → Hn ( X ) is an isomorphism since the diagram
commutes.
Remark 10.4.1. It can also be shown inductively that under the assump-
tions of the Hurewicz theorem,
is an epimorphism.
0 /A /B /C /0
We need the following easy fact which guarantees that in the Leray-
Serre spectral sequence of the path fibration we have Enp,q ∈ C .
∼
= ∂ ∼
= mod C dn
∼
πn−1 (ΩX ) n−1 / Hn−1 (ΩX ) = E0,n
= mod C 2 n
−1 = . . . = E0,n−1
hΩX
−1
Specifically, hnΩX is an isomorphism mod C by the inductive hypothesis,
∂ is an isomorphism by the long exact sequence associated to the path
fibration, and dn is an isomorphism mod C by a spectral sequence
argument similar to the one used in the proof of the Hurewicz theorem.
Therefore, hnX is an isomorphism mod C since the diagram commutes.
E2p,q = H p (Y, Hq ( F )) ⇛ H∗ ( X ),
and
E2p,0 = H p (Y ).
E2 = · · · = E ∞ .
Hn ( X ) = Dn,0 ⊇ Dn−1,1 ⊇ · · · ⊇ 0
H∗ ( B)
n
0 E 2 = · · · = E n +1
0
H∗ ( B)
In particular,
+1
Enp,q = · · · = E2p,q
for any ( p, q), and
0
, q ̸= 0, n
∞
E p,q = ker(dn+1 : Enp,0 +1
→ Enp−+1
n−1,n ) ,q = 0 (10.5.1)
+1
coker(dn+1 : Enp+ n +1
n+1,0 → E p,n ) , q = n.
We first need to compute the homology of the loop space ΩSn for n > 1.
Proof. Consider the Leray-Serre spectral sequence for the path fibration
(with π1 (Sn ) = π0 (ΩSn ) = 0)
ΩSn ,→ PSn ≃ ∗ → Sn ,
with E2 -page
H (ΩSn ) , p = 0, n
q
E2p,q = H p (Sn ; Hq (ΩSn )) =
0 , otherwise
H∗ (ΩSn )
E2 = · · · = E n
Hi (ΩSn ) Hi (ΩSn )
.. ..
. 0 .
H1 (ΩSn ) H1 (ΩSn )
... H∗ (Sn )
0 n
E2 = . . . = E n .
Since E∞ n
p,q = 0 for all ( p, q ) ̸ = (0, 0), all nonzero entries in E (except at
the origin) have to be killed in En+1 . In particular,
dnn,q : En,q
n n
−→ E0,q + n −1
are isomorphisms.
H∗ (ΩSn )
E2 = · · · = E n
H2n−2 (ΩSn )
..
. dn
Hn (ΩSn ) 0
Hn−1 (ΩSn ) Hn−1 (ΩSn )
0 0
.. dn ..
. .
0 0
0Z ... n Z = H0 (ΩSn )
isomorphisms
Hq (ΩSn ) ∼
= Hq+n−1 (ΩSn )
as desired.
We can now give a new proof of the Suspension Theorem for homo-
topy groups.
g∗
π n −1 ( S n −1 ) - πn−1 (ΩSn )
[id] 7→ [ g ◦ id] = [ g]
From the above claim, we obtain that Hi (ΩSn , Sn−1 ) = 0, for i < 2n − 2,
together with the exact sequence
∼
=
→ H2n−2 (ΩSn , Sn−1 ) → 0
0 → Z = H2n−2 (ΩSn ) −
From the homotopy long exact sequence of the pair (ΩSn , Sn−1 ), we
then get πi (ΩSn ) ∼
= πi (Sn−1 ) for i < 2n − 3 and the exact sequence
so that
p,q
E∞ = D ⧸D p+1,q−1 .
p,q
n-th diagonal of E∞
H n ( E)/D1,n−1
D1,n−1 /D2,n−2
D n−1,1 /D n,0
D n,0
dr ( x • y) = dr ( x ) • y + (−1)deg( x) x • dr (y)
248 algebraic topology
and
0,q 0,q 0,q 0,q
H q ( E) ↠ E∞ = Eq+1 ⊂ Eq ⊂ · · · ⊂ E2 = H q ( F ) (10.7.2)
Recall that for a space of finite type, the (co)homology groups are
finitely generated. By using the universal coefficient theorem in coho-
mology, we have the following useful result:
Proposition 10.7.3. Suppose that F ,→ E → B is a fibration with F connected
and assume that π1 ( B) = 0 (or π1 ( B) acts trivially on the fiber cohomology).
If B and F are spaces of finite type (e.g., finite CW complexes), then for a field
K of coefficients we have:
p,q
E2 = H p ( B; K) ⊗K H q ( F; K).
H ∗ ( E; K) ∼
= H ∗ ( B; K) ⊗K H ∗ ( F; K).
E2 = · · · = E∞ ,
i.e., that all differentials d2 , d3 , etc., vanish. Indeed, since we work with
field coefficients, all extension problems encountered in passing from
E∞ to H ∗ ( E; K) are trivial, i.e.,
p,q
H n ( E; K) ∼
M
= E∞ .
p+q=n
p,0
since K is a field, and d2 is already zero on E2 since we work with a
first quadrant spectral sequence. Since d2 is a derivation with respect
to (10.7.3), we conclude that d2 = 0 and E3 = E2 . The same argument
applies to d3 and, continuing in this fashion, we see that the spectral
sequence collapses (degenerates) at E2 , as desired.
S1 ,→ S∞ ≃ ∗ → CP∞ .
H ∗ ( S1 )
E2
Z
d2
H ∗ (CP∞ )
Z 0 Z
250 algebraic topology
H ∗ ( S1 )
E2
x 0 xy
d2 d2
H ∗ (CP∞ )
1 0 y 0 y2
d2 ( a) = rx. (10.8.2)
H ∗ ( S1 )
E2
a 0 ax 0 ax2 ax n
d2 d2
... H ∗ (CPn )
1 0 x 0 x2 xn
252 algebraic topology
0 0 0 0 0 Z
...
Z 0 Z/r 0 Z/r Z/r
Theorem 10.9.1. If n ≥ 3,
K (πk , k) ,→ Yk → Yk−1
Y1 ko Y2 oi · · · Yn−1 o YO n
c
H4 (Y4 ) = H5 (Y4 ) = 0.
Let us now consider the homology spectral sequence for the fibration
(10.9.1). By the Hurewicz theorem,
0 p = 1, 2
H p (K (Z, 3); Z) =
Z p = 3
0 q = 1, 2, 3
Hq (K (π4 , 4); Z) =
π4 ( S3 ) q = 4.
H∗ (K (π4 , 4))
π4
d5
0
H∗ (K (Z, 3)
Z 0 0 Z H4 H5
Since H4 (Y4 ) = 0 = H5 (Y4 ), all entries on the fourth and fifth diagonals
of E∞ are zero. The only differential that can affect π4 (S3 ) = E0,4 2 =
5
· · · = E0,4 is
d5 : H5 (K (Z, 3), Z) −→ π4 (S3 ),
and by the previous remark, this map has to be an isomorphism (note
2 = H ( K (Z, 3), Z) can be affected only by d5 , and this
also that E5,0 5
element too has to be killed at E∞ ). Hence
π4 ( S3 ) ∼
= H5 (K (Z, 3), Z). (10.9.2)
and note that, since PK (Z, 3) is contractible, we have πi (ΩK (Z, 3)) ∼ =
πi+1 (K (Z, 3)), i.e., ΩK (Z, 3) ≃ K (Z, 2) = CP∞ . Since each H j (CP∞ ) is
a finitely generated free abelian group, the universal coefficient theorem
yields that
will sooner or later kill off all the non-zero elements in the spectral
sequence.
H ∗ (CP∞ )
5 0 E2 = E3
4 a2
d3
3 0
·2
2 a as
d3 d3
1 0 ∼
=
1 0 0 s 0 0 y = s2
0 H ∗ (K (Z, 3))
0 1 2 3 4 5 6
Note that E31,0 = E21,0 = H 1 (K (Z, 3)) is never touched by any differen-
tial, so
H 1 (K (Z, 3)) = E∞ 1,0
= 0.
Moreover, since d2 = 0, we also have that
The only differential that can affect ⟨ a⟩ = E20,2 = E30,2 is d0,2 0,2 3,0
3 : E3 → E3 ,
3,0
so there must be an element s ∈ E3 that kills off a, i.e., d3 ( a) = s. On
the other hand, since E33,0 is only affected by d3 and it must be killed
256 algebraic topology
so d0,4
3 : E3
0,4
→ E33,2 is given by multiplication by 2. In particular,
E4 = 0. Next notice that H 4 (K (Z, 3)) = E34,0 and H 5 (K (Z, 3)) = E35,0
0,4
Similarly, H 6 (K (Z, 3)) = E36,0 and ⟨ as⟩ = E33,2 are only affected by
d3 . Since d3 ( a2 ) = 2as, we have ker(d3 : ⟨ as⟩ = E33,2 → E36,0 ) =
Im(d3 : E30,4 → E33,2 = ⟨ as⟩) = ⟨2as⟩ ⊆ ⟨ as⟩, and hence H 6 (K (Z, 3)) =
Im(d3 : E33,2 → E36,0 ) ∼
= ⟨ as⟩ / ⟨2as⟩ = Z/2.
In view of the above calculations, we get by the universal coefficient
theorem that
H5 (K (Z, 3)) = Z/2. (10.9.4)
Corollary 10.9.4.
π4 (S2 ) = Z/2.
Proof. This follows from Theorem 10.9.2 and the long exact sequence
of homotopy groups for the Hopf fibration S1 ,→ S3 → S2 .
Whitehead tower
Let X be a connected CW complex, with πq = πq ( X ) for any q ≥ 0.
· · · −→ Xn −→ Xn−1 −→ · · · → X0 = X
such that
(a) Xn is n-connected
spectral sequences. applications 257
(b) πq ( Xn ) = πq ( X ) for q ≥ n + 1
Theorem 10.10.3.
π5 ( S3 ) ∼
= Z/2.
Proof. Consider the Whitehead tower for X = S3 . Since S3 is 2-
connected, we have in the notation of Definition 10.10.1 that X =
X1 = X2 . Let πi := πi (S3 ), for any i ≥ 0. We have fibrations
K ( π4 , 3) / X4
K ( π3 , 2) / X3
S3
258 algebraic topology
π5 ( S3 ) ∼
= π 5 ( X4 ) ∼
= H5 ( X4 ).
Similarly,
π4 ( S3 ) ∼
= π 4 ( X3 ) ∼
= H4 ( X3 ).
Once again we are reduced to computing homology groups. Using the
universal coefficient theorem, we will deduce the homology groups
from cohomology.
Consider now the cohomology spectral sequence for the fibration
CP∞ ,→ X3 → S3 .
p,q
In particular, E2 = 0 unless p = 0, 3 and q is even.
H ∗ (CP∞ )
E2 = E3
4 x2
d3
3 0
·2
2 x xu
d3
1 0 ∼
=
1 0 0 u
0 H ∗ ( S3 )
0 1 2 3
p,q
Since E2 = 0 for q odd, we have d2 = 0, so E2 = E3 . In addition, for
r ≥ 4, dr = 0. So E4 = E∞ .
Since X3 is 3-connected, we have by Hurewicz that H 2 ( X3 ) =
3
H ( X3 ) = 0, so all entries on the second and third diagonals of
E∞ = E4 are 0. This implies that d0,2 3 : E30,2 = Z → E33,0 = Z is
an isomorphism. Let H ∗ (CP∞ ) = Z[ x ] with x of degree 2, and let u be
a generator of H 3 (S3 ). Then we have d3 ( x ) = u. By the Leibnitz rule,
d3 x n = nx n−1 dx = nx n−1 u, and since x n generates E30,2n and x n−1 u
generates E33,2n−2 , the differential d0,2n
3 is given by multiplication by n.
This completely determines E4 = E∞ , hence the integral cohomology
and (by the universal coefficient theorem) homology of X3 is easily
computed as:
q 0 1 2 3 4 5 6 7 ··· 2k 2k + 1 ···
H q ( X3 ) Z 0 0 0 0 Z/2 0 Z/3 ··· 0 Z/k ···
Hq ( X3 ) Z 0 0 0 Z/2 0 Z/3 0 ··· Z/k 0 ···
spectral sequences. applications 259
K (π4 , 3) ,→ X4 → X3 ,
with E2 -page
H∗ (K (Z/2, 3))
5 Z/2
4 0
3 Z/2
d6
2 0 d4
1 0
Z 0 0 0 Z/2 0 Z/3
0 H∗ ( X3 )
0 1 2 3 4 5 6
Proof of part (a). The case k = 0 is easy since πi (S1 ) is in fact trivial
for i > 1. For k > 0, recall Serre’s theorem 10.4.2, according to which
a simply-connected finite CW complex has finitely generated homo-
topy groups. In particular, the groups πi (S2k+1 ) are finitely generated
abelian for all i > 1. Therefore, πi (S2k+1 ) (i > 1) is finite if it is a torsion
group.
In what follows we show that
πi (S2k−1 ) ∼
= πi+2 (S2k+1 ) mod torsion, (10.11.1)
and part (a) of the theorem follows then by induction. The key to
proving the isomorphism (10.11.1) is the fact that
The relative version of the Hurewicz mod torsion Theorem 10.4.5 then
tells us that
πi (Ω2 S2k+1 , S2k−1 ) = 0 mod torsion
for all i, so again by the homotopy long exact sequence of the pair
we get that πi (S2k−1 ) ∼
= πi (Ω2 S2k+1 ) ∼
= πi+2 (S2k+1 ) mod torsion, as
desired.
Thus, it remains to show that the generator β : S2k−1 → Ω2 S2k+1 of
π2k−1 (Ω2 S2k+1 ) induces an isomorphism on H∗ (−; Q). The bulk of the
argument amounts to showing that Hi (Ω2 S2k+1 ; Q) = 0 for i ̸= 2k − 1,
which we do by computing Hi (Ω2 S2k+1 ; Q)∨ = H i (Ω2 S2k+1 ; Q) with
the help of the cohomology spectral sequence for the path fibration
Ω2 S2k+1 ,→ ∗ → ΩS2k+1 . The E2 -page is given by
p,q
E2 = H p (ΩS2k+1 ; H q (Ω2 S2k+1 ; Q)) ⇛ H ∗ (∗; Q),
spectral sequences. applications 261
p,q
and since the total space of the fibration is contractible, we have E∞ = 0
0,0 ∼
unless p = q = 0, in which case E∞ = Q.
It is a simple exercise (using the path fibration ΩS2k+1 ,→ ∗ → S2k+1 )
to show that
H ∗ (ΩS2k+1 ; Q) ∼
= Q[e], deg e = 2k.
Hence,
p,q
E2 = H p (ΩS2k+1 ; H q (Ω2 S2k+1 ; Q))
∼ H p (ΩS2k+1 ; Q) ⊗Q H q (Ω2 S2k+1 ; Q)
=
2kj,0 ∼
has possibly non-trivial columns only at multiples p of 2k, with E2 =
Q = ⟨e ⟩. This implies that d2 , d3 , . . . , d2k−1 are all zero, hence E2 = E2k .
j
H ∗ (Ω2 S2k+1 ; Q)
E2 = · · · = E2k
2k − 1 ω 0 eω 0
d2k d2k
0 1 0 e 0 e2
... H ∗ (ΩS2k+1 ; Q)
0 ... 2k ... 4k
2k,0 ∼ 0,2k −1 ∼
Since E2k = H 2k (ΩS2k+1 ) = ⟨e⟩ and E2k = H 2k−1 (Ω2 S2k+1 ) are
−1 0,2k −1 −1
only affected by d0,2k
2k : E2k 2k,0
→ E2k , we must have that d0,2k2k is an
2k,0 2k,0 0,2k −1 0,2k−1
isomorphism in order for E2k+1 = E∞ and E2k+1 = E∞ to be zero.
So H 2k−1 (Ω2 S2k+1 ) ∼
= Q = ⟨ω ⟩, with d2k (ω ) = e. As a consequence,
2jk,2k −1
E2k = H 2jk (ΩS2k+1 ; Q) ⊗Q H 2k−1 (Ω2 S2k+1 ) = ⟨e j ⟩ ⊗Q ⟨ω ⟩ = ⟨e j ω ⟩
S2k−1 ,→ E −
→ S2k
π
such that
πi ( E ) ∼
= πi (S4k−1 ) (mod torsion). (10.11.2)
Assuming for now that such a fibration exists, then since by part (a) we
have that
finite i ̸= 4k − 1
πi (S4k−1 ) = ,
Z i = 4k − 1
we deduce that
finite i ̸= 4k − 1
πi ( E ) =
Z ⊕ finite i = 4k − 1.
as desired.
spectral sequences. applications 263
E = T0 S2k .
Therefore, the page E2 has only four non-trivial entries at ( p, q) = (0, 0),
(2k, 0), (0, 2k − 1), (2k − 1, 2k), and all these entries are isomorphic to
Z.
H∗ (S2k−1 )
E2 = · · · = E2k
Z Z
d2k
H∗ (S2k )
Z Z
Clearly, the differentials d2 , d3 , . . . , d2k−1 are all zero, as are the dif-
ferentials d2k+1 , . . . . The only possibly non-zero differential in the
spectral sequence is d2k 2k 2k 2 2k
2k,0 : E2k,0 → E0,2k −1 . Thus, E = · · · = E and
E2k+1 = · · · = E∞ . Therefore, the space E has the desired homology if
and only if
d2k
2k,0 ̸ = 0.
264 algebraic topology
π2k (S2k )
∂ / π2k−1 (S2k−1 )
h ∼
= ∼
= h
d2k
H2k (S2k ) / H2k−1 (S2k−1 )
H ∗ (U (n); Z) = ΛZ [ x1 , · · · , x2n−1 ].
U (n − 1) ,→ U (n) → S2n−1 .
deg( xi ) = i.
d1 = · · · = d2n−2 = 0,
so
E2 = · · · = E2n−1 .
Furthermore, higher differentials starting with d2n are also zero (since
either their domain or target is zero), so
E2n = · · · = E∞ .
d2n−1 = 0.
(Here, x2n−1 denotes the generator of H ∗ (S2n−1 ).) Thus, E2n−1 = E2n ,
so in fact the spectral sequence degenerates at the E2 -page, i.e.,
E2 = · · · = E∞ .
as desired.
266 algebraic topology
hence
H ∗ (SU (n)) = ΛZ [ x3 , . . . , x2n−1 ]
with deg xi = i.
10.13 Exercises
1. Show that πi (ΣRP2 ) are finitely generated abelian groups for any
i ≥ 0. (Hint: Use Theorem 10.4.5, with C the category of finitely
generated 2-groups.
χ ( E ) = χ ( B ) · χ ( F ).
6. Prove that H5 (K (π4 , 3)) = Z/2. (Hint: consider the two fibrations
K (Z/2, 2) = ΩK (Z/2, 3) ,→ ∗ → K (Z/2, 3), and RP∞ = K (Z/2, 1) ,→
∗ → K (Z/2, 2). Then compute H∗ (K (Z/2, 2)) via the spectral sequence
of the second fibration, and use it in the spectral sequence of the first
fibration to compute H∗ (K (Z/2, 3)).)
K (Z, n − 1) ,→ ∗ → K (Z, n)
and induction.)
15. Show that the p-torsion in πi (S3 ) appears first for i = 2p, in which
case it is Z/p. (Hint: use the Whitehead tower of S3 , the homology
spectral sequence of the relevant fibration, together with Hurewicz mod
C p , where C p is the class of torsion abelian groups whose p-primary
subgroup is trivial.)
16. Where does the 7-torsion appear first in the homotopy groups of
Sn ?
fiber bundles. classifying spaces. applications 269
11
Fiber bundles. Classifying spaces. Applications
Definition 11.1.1 (Atlas for a fiber bundle with group G and fiber F).
Given a continuous map π : E → B, an atlas for the structure of a fiber bundle
with group G and fiber F on π consists of the following data:
a) an open cover {Uα }α of B,
π pr1
# |
Uα
270 algebraic topology
commutes,
π −1 (Uα ∩ Uβ )
hα hβ
v (
(Uα ∩ Uβ ) × F / (Uα ∩ Uβ ) × F
h β ◦ h−
α
1
is given by
( x, m) 7→ ( x, gβα ( x ) · m).
(By the effectivity of the action, if such maps gαβ exist, they are unique.)
Definition 11.1.3 (Fiber bundle with group G and fiber F). A structure
of a fiber bundle with group G and fiber F on π : E → B is a maximal atlas
for π : E → B.
Example 11.1.4.
map.
Lemma 11.1.5. The transition functions gαβ satisfy the following properties:
−1
(b) g βα ( x ) = gαβ ( x ), for all x ∈ Uα ∩ Uβ .
(c) gαα ( x ) = eG .
fiber bundles. classifying spaces. applications 271
gαβ ( x ) g βγ ( x ) = gαγ ( x )
for all x ∈ Uα ∩ Uβ ∩ Uγ .
Note that (hα ◦ h− 1 −1
β ) ◦ ( h β ◦ hα ) = id, which translates into
Example 11.1.7.
1. Fiber bundles with fiber F = Rn and group G = GL(n, R) are called
rank n real vector bundles. For example, if M is a differentiable real
n-manifold, and TM is the set of all tangent vectors to M, then
π : TM → M is a real vector bundle on M of rank n. More precisely,
∼
=
if φα : Uα → Rn are trivializing charts on M, the transition functions
for TM are given by gαβ ( x ) = d( φα ◦ φ− 1
β ) φβ (x) .
272 algebraic topology
Theorem 11.1.8. A fiber bundle has the homotopy lifting property with respect
to all CW complexes (i.e., it is a Serre fibration). Moreover, fiber bundles over
paracompact spaces are fibrations.
1. the diagram
fˆ
E′ /E
π′ π
f
B′ /B
commutes, i.e., π ◦ fˆ = f ◦ π ′ .
2. if {(Uα , hα )}α is a trivializing atlas of π and {(Vβ , Hβ )} β is a trivializing
atlas of π ′ , then the following diagram commutes:
Hβ fˆ
(Vβ ∩ f −1 (Uα )) × F o / π −1 (Uα ) / Uα × F
−1 hα
π′ (Vβ ∩ f −1 (Uα ))
π′ π
pr1 pr1
( z
Vβ ∩ f −1 (Uα )
f
/ Uα
in G, where gα′ α are transition functions for π and g ββ′ are transition functions
for π ′ ,
Proof. Exercise.
ˆ ′
a map of bundles ( f , f ) : π → π exists if and only if there exist continuous
maps {dαβ } as above, satisfying (11.1.1).
Proof. Exercise.
d βα ( x ) = d− 1
αβ ( x )
to get:
d β′ α′ ( x ) = g β′ β ( x ) d βα ( x ) gαα′ ( x ).
So {d βα } are as in Definition 11.1.9 and satisfy (11.1.1). Theorem 11.1.12
implies that there exists a bundle map ĝ : E → E′ over idB .
We claim that ĝ is the inverse fˆ−1 of fˆ, and this can be checked
locally as follows:
fˆ ĝ
( x, m) 7→ ( x, dαβ ( x ) · m) 7→ ( x, d βα ( x ) · (dαβ ( x ) · m))
= ( x, d βα ( x )dαβ ( x ) ·m)
| {z }
eG
= ( x, m).
So ĝ ◦ fˆ = id E′ . Similarly, fˆ ◦ ĝ = id E
274 algebraic topology
f ∗ E := {( x, e) ∈ X × E | f ( x ) = π (e)},
f ∗E E e
f ∗π π
X B
f
x f (x)
Theorem 11.1.15.
hα
( f ∗ π )−1 ( f −1 (Uα )) π −1 (Uα ) Uα × F
f −1 (Uα ) Uα
f
We have
Define
k α : ( f ∗ π )−1 ( f −1 (Uα )) −→ f −1 (Uα ) × F
by
( x, e) 7→ ( x, pr2 (hα (e))).
Then it is easy to check that k α is a homeomorphism (with inverse
k− 1 −1
α ( x, m ) = ( x, hα ( f ( x ), m )), and in fact the following assertions hold:
fiber bundles. classifying spaces. applications 275
As a consequence, we have:
ct∗ i∗ E / i∗ E /E
ct∗ i∗ π i∗ π π
B
ct / {b} i /B
>
id B
π∼
= (id B )∗ π ∼
= ct∗ i∗ π.
Proposition 11.1.19. If
f˜
E′ /E
π′ π
f
B′ /B
fˆ
f ∗E /) E
π′
π′ π
f
B′ /B
Example 11.1.20. We can now show that the set of isomorphism classes
of bundles over Sn with group G and fiber F is isomorphic to πn−1 ( G ).
Indeed, let us cover Sn with two contractible sets U+ and U− obtained
by removing the south, resp., north pole of Sn . Let i± : U± ,→ Sn be
the inclusions. Then any bundle π over Sn is trivial when restricted
∗ π ∼ U × F. In particular, U provides a trivializing
to U± , that is, i± = ± ±
cover (atlas) for π, and any such bundle π is completely determined by
the transition function g± : U+ ∩ U− ≃ Sn−1 → G, i.e., by an element
in πn−1 ( G ).
f ∗ : B ( X, G, F, ρ) −→ B X ′ , G, F, ρ
As we will see later on, the fiber F doesn’t play any essential role in the
classification of fiber bundle, and in fact it is enough to understand the
set
P ( X, G ) := B ( X, G, G, mG )
Proof. We will define the action locally over a trivializing chart for π.
Let Uα be a trivializing open in X with trivializing homeomorphism
∼
=
hα : π −1 (Uα ) → Uα × G. We define a right action on G on π −1 (Uα ) by
π −1 (Uα ) × G → π −1 (Uα ) ∼
= Uα × G
(e, g) 7→ e · g := h− 1
α ( π ( e ) , pr2 ( hα ( e )) · g )
h− 1 −1
α ( π ( e ) , pr2 ( hα ( e )) · g ) = h β π (e) , pr2 h β (e) · g . (11.2.1)
After applying hα and using the transition function gαβ for π (e) ∈
Uα ∩ Uβ , (11.2.1) becomes
P ( X, G ) → B ( X, G, F, ρ)
g · (e, f ) 7→ (e · g−1 , g · f ).
Let
E ×G F := E × F⧸G
E×F (11.2.2)
pr1 $
} E × F⧸
E G
π
ω
! z
X
ω −1 (Uα ) ∼ π −1 (Uα ) × F⧸
= (e, f ) ∼ (e · g−1 , g · f )
= Uα × G × F⧸(u, h, f ) ∼ (u, hg−1 , g · f ).
∼
Let us define
k α : ω −1 (Uα ) → Uα × F
by
[(u, h, f )] 7→ (u, h · f ).
fiber bundles. classifying spaces. applications 279
Proposition 11.2.8. If
fb
E′ /E
π′ π
f
X′ /X
fb×G id F
E′ × G F / E ×G F
π′ π
f
X′ /X
f ∗E
? / g∗ E
~
X
280 algebraic topology
fˆ
E1 / E2
π1 π2
f
X /Y
Since G acts on the right of E1 and E2 , we also get an action on the left
of E2 by g · e2 := e2 · g−1 . Then we get an associated bundle of π1 with
fiber E2 , namely
ω := π1 ×G E2 : E1 ×G E2 −→ X.
fb
V×G U×G
π1 π2
f
V U
We define a section σ in
( V × G ) × G (U × G )
σ ω
[e1 · g, fb(e1 · g)] = [e1 · g, fb(e1 ) · g] = [e1 · g, g−1 · fb(e1 )] = [e1 , fb(e1 )].
where σ (π1 (e1 )) = [(e1 , e2 )]. Note that this is an equivariant map
because
[e1 · g, e2 · g] = [e1 · g, g−1 · e2 ] = [e1 , e2 ],
bi0 pr
/E / E0
b1
E0
π0 π π0
X
i0 pr1
/ X×I /X
X × { 0}
σ0
/ E ×G E0
_ 9
s
ω
X×I
id / X×I
f ∗E / H∗ E H
b
/E
:
f ∗π g∗ E H∗ π π
X × {0}
i0
g∗ π / X×I H /Y
:
i1 pr1
, !
X × {1} X
E /G
π π′
X
ct / point
π −1 (Uα ) × G⧸ ∼ Uα × G × G⧸ ∼
∼= (u, g1 , g2 ) ∼ (u, g1 g−1 , gg2 ) = Uα × G,
X = A ∪ϕ en ,
σ0 ( x ) = ( x, τ0 ( x )) ∈ en × F,
with τ0 : ∂en ∼
= Sn−1 → F. Since πn−1 ( F ) = 0, τ0 extends to a map
τ : e → F which can be used to extend σ0 over en by setting
n
σ ( x ) = ( x, τ ( x )).
fb
E0 = f ∗ EG −−−−→ EG
π π
y 0 y G
f
X = X × {0} −−−−→ BG
gb
E0 ∼
= E1 = g∗ EG −−−−→ EG
π π
y 0 y G
g
X = X × {1} −−−−→ BG
where we regard gb as defined on E0 via the isomorphism π0 ∼
= π1 . By
putting together the above diagrams, we have a commutative diagram
ω : ( E0 × I ) ×G EG → X × I.
B(Sn , G, F, ρ) ∼
= P (Sn , G ) ∼
= [Sn , BG ] = πn ( BG ) ∼
= π n −1 ( G ),
fiber bundles. classifying spaces. applications 285
where the last isomorphism follows from the homotopy long exact
sequence for πG , since EG is contractible.
′ − fˆ
EG −−−→ EG
π ′ π
y G y G
′ − f
BG −−−→ BG
π G = g∗ π G
′
= g∗ f ∗ π G = ( f ◦ g)∗ π G .
Example 11.3.5. Recall from Section 9.12 that we have a fiber bundle
O(n) / Vn (R∞ ) / Gn (R∞ ), (11.3.1)
T : [ X, K ( G, n)] −→ H n ( X, G )
[ f ] 7→ f ∗ (α)
where α ∈ H n (K ( G, n), G ) ∼
= Hom( Hn (K ( G, n), Z), G ) is given by the
inverse of the Hurewicz isomorphism G = πn (K ( G, n)) → Hn (K ( G, n), Z).
the complex line (i.e., rank-one) bundles. Since we also have that
CP∞ = K (Z, 2), we get:
cohomology class
c1 (π ) := f π∗ (c)
∼
=
called the first Chern class of π. The bijection P ( X, S1 ) −→ H 2 ( X, Z) is
then given by π 7→ c1 (π ), so complex line bundles on X are classified
by their first Chern classes.
fiber bundles. classifying spaces. applications 287
11.4 Exercises
c1 ( π ) = d2 ( a ),
(b) What is the first Chern class of the sphere (or unit) bundle of the
tangent bundle TS2 ?
12
Vector Bundles. Characteristic classes.
Cobordism. Applications.
Proposition 12.1.1.
H ∗ ( BU (n); Z) ∼
= Z [ c1 , · · · , c n ] ,
with deg ci = 2i
H ∗ (U ( n ) ; Z ) ∼
= ΛZ [ x1 , · · · , x2n−1 ].
Then using the Leray-Serre spectral sequence for the universal U (n)-
bundle, and using the fact that EU (n) is contractible, yields the desired
result.
Alternatively, the functoriality of the universal bundle construction
yields that for any subgroup H < G of a topological group G, there
is a fibration G/H ,→ BH → BG. In our case, consider ! U (n − 1) as a
A 0
subgroup of U (n) via the identification A 7→ . Hence, there
0 1
exists fibration
U (n)/U (n − 1) ∼
= S2n−1 ,→ BU (n − 1) → BU (n).
ci (π ) := f π∗ (ci ) ∈ H 2i ( X; Z).
fb
f ∗E /E / EU (n)
f ∗π π πU ( n )
f fπ
Y /X / BU (n)
ci ( f ∗ π ) = ( f π ◦ f )∗ ci
= f ∗ ( f π∗ ci )
= f ∗ ci ( π ) .
c(π ) = c0 (π ) + c1 (π ) + · · · cn (π ) = 1 + c1 (π ) + · · · cn (π ) ∈ H ∗ ( X; Z),
π1 ⊕ π2 : = ∆ ∗ ( π1 × π2 ),
c ( π1 ⊕ π2 ) = c ( π1 ) ∪ c ( π2 ).
Indeed, by taking the product of the universal bundles for U (n) and
U (m), we get a U (n) × U (m)-bundle over BU (n) × BU (m), with total
space EU (n) × EU (m):
By using the Künneth formula, one can show (e.g., see Milnor’s book,
p.164) that:
ω ∗ c k = ∑ ci × c j .
i + j=k
Therefore,
= ∑ c i ( π1 ) ∪ c j ( π2 ).
i+ j=k
Here, we use the fact that the classifying map for π1 × π2 , regarded as
a U (n + m)-bundle is ω ◦ ( f π1 × f π2 ).
Since the trivial bundle has trivial Chern classes in positive degrees,
we get
Proposition 12.2.1.
H ∗ ( BO(n); Z/2) ∼
= Z/2 [w1 , · · · , wn ] ,
with deg wi = i.
O(n)/O(n − 1) ∼
= Sn−1 ,→ BO(n − 1) → BO(n),
wi (π ) := f π∗ (wi ) ∈ H i ( X; Z/2).
vector bundles. characteristic classes. cobordism. applications. 293
w(π ) = 1 + w1 (π ) + · · · wn (π ) ∈ H ∗ ( X; Z/2),
f ∗ TN = TM ⊕ ν, (12.3.1)
294 algebraic topology
f ∗ w ( N ) = w ( M ) ∪ w ( ν ). (12.3.2)
w ( ν ) = w ( M ) −1 ∪ f ∗ w ( N ).
~
π1 π2
X
be a linear monomorphism of vector bundles, i.e., in local coordinates, i is
given by U × Rn → U × Rm (n ≤ m), (u, v) 7→ (u, ℓ(u)v), where ℓ(u)
is a linear map of rank n for all u ∈ U. Then there exists a vector bundle
π1⊥ : E1⊥ → X so that π2 ∼= π1 ⊕ π1⊥ .
Theorem 12.3.5.
w(RPm ) = (1 + x )m+1 , (12.3.3)
w(RP9 )−1 = 1 + x2 + x4 + x6 .
w1 (π ) = i∗ w1 = i∗ x = x,
TRPm ⊕ E 1 ∼
=π ⊕ ·{z
· · ⊕ π}, (12.3.4)
|
m+1 times
TSm ⊕ Eν ∼
= TRm+1 |Sm = E m+1 ∼
=E 1
· · ⊕ E }1 ,
⊕ ·{z
|
m+1 times
with E m+1 the trivial bundle of rank m + 1 on Sm , i.e., the Whitney sum
of m + 1 trivial line bundles E 1 on Sm , each of which is generated by
the global non-zero section y 7→ dxd |y , i = 1, · · · , m + 1.
i
Let a : Sm → Sm be the antipodal map, with differential da : TSm →
TSm . Let γ : (−ϵ, ϵ) → Sm , γ(0) = y, v = γ′ (0) ∈ Ty Sm . Then
da(v) = dt d
( a ◦ γ(t))|t=0 = −γ′ (0) = −v ∈ Ta(y) Sm . Therefore da is an
involution on TSm , commuting with a, and hence
d d ∼ n
Sm × R/da ∼
= Sm × R/(y, t ) ∼ (−y, −t ) = S ×Z/2 R,
dxi dxi
which is the associated bundle of π with fiber R. So,
E 1 /da ∼
= π.
Remark 12.3.13. Note that RP3 = ∼ SO(3) is a Lie group, so its tangent
bundle is trivial. In this case, once can conclude directly that w(RP3 ) =
1, but this fact can also be seen from formula (12.3.3).
Boundary Problem
For a closed smooth manifold Mn , let µ M ∈ Hn ( M, Z/2) be the funda-
mental class. We will associate to M certain Z/2-invariants, called its
Stiefel-Whitney numbers.
Definition 12.3.14. Let α = (α1 , . . . , αn ) be a tuple of non-negative integers
such that ∑in=1 iαi = n. Set
i∗ TW ∼
= TM ⊕ ν1 ,
w k ( M ) = i ∗ w k (W ) ,
∂ ( µW ) = µ M .
w ( α ) ( M ) = ⟨ w [ α ] ( M ), µ M ⟩
= ⟨i∗ w[α] (W ), ∂µW ⟩
= ⟨δ(i∗ w[α] (W )), µW ⟩
= ⟨0, µW ⟩
= 0,
Example 12.3.18. The real projective space RP2k is not a boundary, for
any integer k ≥ 0. Indeed, the total Stiefel-Whitney class of RP2k is
2k + 1 2k + 1 2k
w(RP2k ) = (1 + x )2k+1 = 1 + x+···+ x
1 2k
= 1 + x + · · · + x2k
ωR ⊗ C ∼
= ω ⊕ ω.
FR ⊗ C ∼
= Eigen(i ) ⊕ Eigen(−i ),
π⊗C ∼
= π ⊗ C.
ck (ω ) = (−1)k · ck (ω ),
for any k = 1, · · · , n.
S2k−1 ,→ BU (k − 1) → BU (k ).
In fact,
ck = d2k ( a),
ck (π ⊗ C) = ck (π ⊗ C) = (−1)k ck (π ⊗ C).
p( M) := p( TM).
p( M ) := p(( TM)R ).
were E 1 is the trivial complex line bundle on CPn and γ is the complex
line bundle associated to the principle S1 -bundle S1 ,→ S2n+1 → CPn .
Then γ is classified by the inclusion
/ S∞
S2n+1
CPn / CP∞ = BU (1)
c( TCPn ) = (1 − c)n+1 .
Therefore,
CPn ,→ R2n+k ,
( TCPn )R ⊕ νk ∼
= TR2n+k |CPn ∼
= E 2n+k , (12.4.3)
p(CPn ) · p(νk ) = 1.
vector bundles. characteristic classes. cobordism. applications. 303
Therefore, we get
p(νk ) = p(CPn )−1 . (12.4.4)
And by the definition of the Pontryagin classes, we know that if
pi (νk ) ̸= 0, then i ≤ 2k .
Example 12.4.11. In this example, we use Pontrjagin classes to show that
CP2 does not embed in R5 . First,
p(CP2 ) = (1 + c2 )3 = 1 + 3c2 ,
Ωn −→ Z, [ M ] 7→ p(α) ( M ).
Example 12.5.11. The above remark and Example 12.5.9 show that
CP2n does not have any orientation-reversing diffeomorphism. How-
ever, CP2n+1 has an orientation-reversing diffeomorphism induced by
complex conjugation.
vector bundles. characteristic classes. cobordism. applications. 305
We start by noting that both CP4 and CP2 × CP2 are compact oriented
8-manifolds which are not boundaries. We calculate the Pontrjagin
numbers of these two spaces. First,
where × denotes the external product. Let c1 and c2 denote the genera-
tors of the second integral cohomology of the two CP2 factors. Then:
Hence, p1 (CP2 × CP2 ) = 3(c21 + c22 ) and p2 (CP2 × CP2 ) = 9c21 c22 . There-
fore, the Pontrjagin numbers of CP2 × CP2 corresponding to the parti-
tions α1 and α2 are computed by (here we use the fact that c41 = 0 = c42 ):
H 2k ( M; Q) × H 2k ( M; Q) → Q,
σ= ∑ aα p(α) (12.6.1)
α∈ I
p1 (CP2 ) = 3c2 . Hence p(1) (CP2 ) = 3, and (12.6.2) implies that 1 = 3a,
or a = 31 . Therefore, for any closed oriented 4-manifold M4 we have
that
1 1
σ ( M ) = ⟨ p1 ( M ), µ M ⟩ = p(1) ( M ) ∈ Z.
3 3
Example 12.6.3. On closed oriented 8-manifolds, the signature is com-
puted by (12.6.1) as
and
1 = σ (CP2 × CP2 ) = 18a(2,0) + 9a(0,1) . (12.6.5)
Solving for a(2,0) and a(0,1) in (12.6.4) and (12.6.5), we get:
1 7
a(2,0) = − , a(0,1) = .
45 45
Altogether, the signature of a closed oriented manifold M8 is computed
by the following formula:
1
σ ( M8 ) = ⟨7p2 ( M) − p1 ( M)2 , µ M ⟩. (12.6.6)
45
= π3 ( S3 × S3 ) ∼
π3 (SO(4)) ∼ = Z ⊕ Z.
1
σ( M) = 7p(0,1) ( M ) − p(2,0) ( M ) .
45
Moreover, one can show that:
Lemma 12.7.2. p(2,0) ( M ) = 4(i − j)2 = 4(2i − 1)2 .
Note that σ ( M ) = ±1 since H 4 ( M; Z) = Z, and let us fix the
orientation according to which σ ( M ) = 1. Our assumption that X was
diffeomorphic to S7 leads now to a contradiction, since
4(2i − 1)2 + 45
p(0,1) ( M ) =
7
is by definition an integer for all i, which is contradicted e.g., for i = 2.
So far (for i = 2 and j = −1), we constructed a space X which is
homotopy equivalent to S7 , but which is not diffeomorphic to S7 . In
fact, one can further show the following:
12.8 Exercises
Bibliography
James F. Davis and Paul Kirk. Lecture notes in algebraic topology, vol-
ume 35 of Graduate Studies in Mathematics. American Mathematical
Society, Providence, RI, 2001.
Index