Download as pdf or txt
Download as pdf or txt
You are on page 1of 168

Christian Bär

GEOMETRIE IN POTSDAM

Algebraic Topology
Lecture Notes, Winter Term 2016/17

Version of February 23, 2017


© Christian Bär 2017. All rights reserved.

Picture on title page taken from www.sxc.hu.


Contents
Preface 1

1 Set Theoretic Topology 3


1.1 Typical problems in topology . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Some basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Hausdorff spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.5 Quotient spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 Product spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2 Homotopy Theory 17
2.1 Homotopic maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 The fundamental group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 The fundamental group of the circle . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 The Seifert-van Kampen theorem . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.5 The fundamental group of surfaces . . . . . . . . . . . . . . . . . . . . . . . . 52
2.6 Higher homotopy groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

3 Homology Theory 79
3.1 Singular homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.2 Relative homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.3 The Eilenberg-Steenrod axioms and applications . . . . . . . . . . . . . . . . 87
3.4 The degree of a continuous map . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.5 Homological algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.6 Proof of the homotopy axiom . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
3.7 Proof of the excision axiom . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
3.8 The Mayer-Vietoris sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
3.9 Generalized Jordan curve theorem . . . . . . . . . . . . . . . . . . . . . . . . 126
3.10 CW-complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
3.11 Homology of CW-complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
3.12 Betti numbers and the Euler number . . . . . . . . . . . . . . . . . . . . . . . 140
3.13 Incidence numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
3.14 Homotopy versus homology . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
3.15 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

Sources list for pictures 157

i
Contents

Bibliography 159

Index 161

ii
Preface
These are the lecture notes of an introductory course on algebraic topology which I taught
at Potsdam University during the winter term 2016/17. The aim was to introduce the basic
tools from homotopy and homology theory. Choices concerning the material had to be made.
Since time was too short for a reasonable discussion of cohomology theory after homology
had been treated, I decided to skip cohomology altogether and instead included more material
from homotopy theory than is often done. In particular, there is a detailed discussion of higher
homotopy groups and the long exact sequence for Serre fibrations.

The necessary prerequisites of the students were rather modest. The course contains a quick
introduction to set theoretic topology but a certain acquaintance with these concepts was
certainly helpful. Familiarity with basic algebraic notions like rings, modules, linear maps etc.
was assumed.

These notes are based on the lecture notes of a course I taught in 2010. I am grateful to Volker
Branding who wrote a first draft of those lecture notes and produced most of the pstricks-figures
and to Ramona Ziese who improved those notes considerably.

Potsdam, February 2017

Christian Bär

1
1 Set Theoretic Topology

1.1 Typical problems in topology


Topology is rough geometry. For example, a sphere is topologically the same as a cube, even
though the sphere is smooth and curved while the cube is piecewise flat and has corners. In
more precise mathematical terms this means that they are homeomorphic. On the other hand,
the sphere is different from a torus even topologically.

Here are four versions of a typical question that one asks in mathematics. Fix the integers
1 ≤ n < m.
1.) Does there exist a linear isomorphism ϕ : Rm → Rn ?
No, since linear algebra tells us that isomorphic vector spaces have equal dimension.
2.) Does there exist a diffeomorphism ϕ : Rm → Rn ?
No, otherwise the differential dϕ(0) : Rm → Rn would be a linear isomorphism.
3.) Does there exist a homeomorphism ϕ : Rm → Rn ?
We cannot answer that question yet, but we will develop the necessary tools to find the
answer.
4.) Does there exist a bijective map ϕ : Rm → Rn ?
Yes. We will now explicitly construct an example for such a map in the case n = 1 and
m = 2.

Example 1.1. Since the exponential function maps R bijectively onto R+ = (0, ∞) it suffices to
construct a bijective map ϕ : R+ → R+ × R+ .

3
1 Set Theoretic Topology

Given x > 0 write x as infinite decimal fraction.

x = . . . a3 a2 a1, b1 b2 b3 . . .

Here a j ∈ {0, 1, . . . , 9} and almost all a j are equal to 0. The b j are blocks of the form 0 . . . 0c j
with c j ∈ {1, . . . , 9}. This representation of x is unique. Now define ϕ(x) = (ϕ1 (x), ϕ2 (x))
where
ϕ1 (x) = . . . a5 a3 a1, b1 b3 b5 . . . ϕ2 (x) = . . . a6 a4 a2, b2 b4 b6 . . .
One easily checks that ϕ maps R+ bijectively onto R+ × R+ .
Let us evaluate the construction for an example. Let x = 1987, 30500735 . . .. Then we can read
off the a j and the b j as

a1 = 7, a2 = 8, a3 = 9, a4 = 1, a j = 0 for j ≥ 5
b1 = 3, b2 = 05, b3 = 007, b4 = 3, b5 = 5, . . .

Hence ϕ1 (x) = 97, 30075 . . . and ϕ2 (x) = 18, 053 . . .

Remark 1.2. In the continuous world counter-intuitive things can happen which are not possible
in the differentiable world. For example, there exist continuous maps

ϕ : [0, 1] → [0, 1] × [0, 1]

which are surjective. Such a map is called a plane-filling curve.

Example 1.3. The first example goes back to Peano [6]. The following example was shortly
after given by Hilbert [3] and is known as the Hilbert curve. It is defined by

ϕ(x) := lim ϕ n (x)


n→∞

where the ϕ n are defined recursively as indicated by Hilbert in the pictures:

4
1.1 Typical problems in topology

ϕ1 ϕ2 ϕ3

ϕ4 ϕ5 ϕ6

Remark 1.4. This cannot happen for smooth curves due to a Theorem by Sard which states that
for smooth ϕ : [0, 1] → R2 the image ϕ([0, 1]) ⊂ R2 is a zero set.

Many typical problems in topology are of the following form:

1.) Given two spaces, are they homeomorphic? To show that they are, construct a homeomor-
phism. To show that they are not, find topological invariants, which are different for given
spaces.

2.) Classify all spaces in a certain class up to homeomorphisms.

3.) Fixed point theorems.

Example 1.5 (Classification theorem for surfaces). The classification theorem for surfaces
states that each orientable compact connected surface is homeomorphic to exactly one in the
following infinite list:

5
1 Set Theoretic Topology

S2 (sphere), genus 0

T 2 (torus), genus 1

F2 (Pretzel surface), genus 2

F3 (true pretzel), genus 3

b
b
b

The genus is the number of “holes” in the surface. We will give a more precise definition later. So
the classification theorem says that to each g ∈ N0 there exists an orientable compact connected
surface Fg with genus g and each orientable compact connected surface is homeomorphic to
exactly one Fg .

Example 1.6 (Brouwer fixed point theorem). Any continuous map f : D n → D n with
D n = {x ∈ Rn |k xk ≤ 1} has a fixed point, i.e., there exists an x ∈ D n such that

f (x) = x.

We give a proof for n = 1. Consider the continuous function g : [−1, 1] → R defined by


g(x) := f (x) − x. Now | f | ≤ 1 implies

g(−1) ≥ −1 − (−1) = 0, g(1) ≤ 1 − 1 = 0.

By the intermediate value theorem we can find an x with g(x) = 0 which is equivalent to
f (x) = x.

6
1.2 Some basic definitions

1.2 Some basic definitions


First of all let us recall the following

Definition 1.7. A subset U ⊂ Rn is called open


iff
∀x ∈ U ∃r > 0 : B(x, r) ⊂ U
r
where B(x, r) = {y ∈ Rn | d(x, y) = k x − yk < r}. b x

The set of open subsets of Rn is called the standard topology of Rn . We write

TRn := {U ⊂ Rn open} ⊂ P(Rn ).

One easily checks

Proposition 1.8. TRn has the following properties

(i) ∅, Rn ∈ TRn
S
(ii) U j ∈ TRn , j ∈ J ⇒ j ∈J U j ∈ TRn

(iii) U1, U2 ∈ TRn ⇒ U1 ∩ U2 ∈ TRn

The importance of the concept of open subsets comes from the fact that one can characterize
continuous maps entirely in terms of open subsets. Namely, a map f : Rn → Rm is continuous
iff for each U ∈ TRm the preimage f −1 (U ) is open, f −1 (U ) ∈ TRn . This motivates taking open
subsets axiomatically as the starting point in topology.

Definition 1.9. A topological space is a pair (X, TX ) with TX ⊂ P(X ) such that

(i) ∅, X ∈ TX
S
(ii) U j ∈ TX , j ∈ J ⇒ j ∈J U j ∈ TX

(iii) U1, U2 ∈ TX ⇒ U1 ∩ U2 ∈ TX

7
1 Set Theoretic Topology

Definition 1.10. Let (X, TX ) be a topological space. Elements of TX are called open subsets
of X . Subsets of the form X \ U with U ∈ TX are called closed.

Examples 1.11. 1.) X arbitrary set, TX = P(X ) (discrete topology). All subsets of X are open
and they are also all closed.

2.) X arbitrary set, TX = {∅, X } (coarse topology). Only X and ∅ are open subsets of X . They
are also the only closed subsets.

3.) Let (X, TX ) be a topological space, let Y ⊂ X be an arbitrary subset. The induced topology
or subspace topology of Y is defined by TY := {U ∩ Y | U ∈ TX }.

4.) Let (X, d) be a metric space. Then we can imitate the construction of the standard topology
on Rn and define the induced metric as the set of all U ⊂ X such that for each x ∈ U there
exists an r > 0 so that B(x, r) ⊂ U. Here B(x, r) = {y ∈ X | d(x, y) < r} is the metric ball
of radius r centered at x.

Remark 1.12. Let two metrics d1 and d2 be given on a set X . If d1 and d2 are equivalent, i.e.,
∃ C ≥ 0 with
C −1 d2 (x, y) ≤ d1 (x, y) ≤ Cd2 (x, y) ∀x, y ∈ X,
then d1 and d2 induce the same topology.

Remark 1.13. Openness is a relative concept. If Y ⊂ X be an arbitrary subset of a topological


space X , then Y is in general not an open subset of X but it is always open as a subset of itself
(w.r.t. the induced topology). The same remark applies to closed subsets.

It is now clear how to define continuous maps between topological spaces.

Definition 1.14. Let (X, TX ) and (Y, TY ) be topological spaces. Then a map f : X → Y is
called continuous iff ∀ U ∈ TY : f −1 (U ) ∈ TX .

Example 1.15. If X carries the discrete topology then every map f : X → Y is continuous.

Example 1.16. If Y carries the coarse topology then every map f : X → Y is continuous.

Remark 1.17. In general, a continuous bijective map need not have a continuous inverse. For
example, let #X > 1 and consider f = id X : (X, Tdiscrete ) → (X, Tcoarse ).

Remark 1.18. f : X → Y is continuous iff f −1 (A) ⊂ X is closed for all closed subsets A ⊂ Y .

8
1.3 Compactness

Definition 1.19. Let X and Y be topological spaces. A bijective continuous map f : X → Y is


called a homeomorphism iff f −1 : Y → X is again continuous. If there exists a homeomorphism
f : X → Y then X and Y are called homeomorphic. We then write X ≈ Y .

Remark 1.20. If f : X → Y and g : Y → Z are continuous then the composition g ◦ f : X → Z


is also continuous.

1.3 Compactness

Definition 1.21. Let X be a topological space. A subset Y ⊂ X is called compact iff for
any collection {Ui }i ∈I , Ui ⊂ X open, with Y ⊂ ∪i ∈I Ui there exist i 1, . . . , i n ∈ I such that
Y ⊂ Ui 1 ∪ · · · ∪ Ui n .

Examples 1.22. 1.) Finite sets are always compact. Namely, et Y = {y1, . . . , yn } and let {Ui }i ∈I
be an open cover of Y . Then choose i j such that yi ∈ Ui j . Then {Ui 1 , . . . , Ui n } still covers Y .

2.) If X carries the discrete topology then a subset Y ⊂ X is compact if and only if it is finite.
We have already seen that finite sets are always compact. Conversely, let Y be a compact
subset of X . We cover Y by {{y} | y ∈ Y }. These one-point sets are open because X carries
the discrete topology. Since Y is compact this cover must be finite and therefore #Y < ∞.

3.) If X carries the coarse topology then every Y ⊂ X is compact

4.) A subset Y ⊂ Rn is compact iff Y is closed and bounded (Heine-Borel theorem).

Example 1.23. In particular, R is not compact. We can see this directly by looking at the open
cover {(x − 1, x + 1)|x ∈ R} of R. Since all intervals in this cover have length 2, any finite
subcover can cover only a bounded subset of R.

Proposition 1.24. Let X be a compact topological space. Let Y ⊂ X be a closed subset. Then
Y is compact.

Proof. Let {Ui }i ∈I be an open cover of Y . Put U := X \ Y ∈ TX . Then {U, Ui }i ∈I is an open


cover of X . Since X is compact, there exist i1, . . . , i n such that X ⊂ U ∪ Ui 1 ∪ · · · ∪ Ui n and we
conclude that Y ⊂ Ui 1 ∪ · · · ∪ Ui n . 

9
1 Set Theoretic Topology

Proposition 1.25. Let f : X → Y be continuous. Let K ⊂ X be compact. Then f (K ) ⊂ Y is


compact.

Proof. Let {Ui }i ∈I be an open cover of f (K ), i.e. f (K ) ⊂ ∪i ∈I Ui . Then we have

K ⊂ f −1 ( f (K )) ⊂ f −1 (∪i ∈I Ui ) = ∪i ∈I f −1 (Ui ) .
| {z }
open

Since K is compact there exist i 1, . . . , i n such that for

K ⊂ f −1 (Ui 1 ) ∪ · · · ∪ f −1 (Ui n ) = f −1 (Ui 1 ∪ · · · ∪ Ui n )

and we conclude that

f (K ) ⊂ f ( f −1 (Ui 1 ∪ · · · ∪ Ui n )) ⊂ Ui 1 ∪ · · · ∪ Ui n .

1.4 Hausdorff spaces

Definition 1.26. A topological space X is called Hausdorff iff ∀x 1, x 2 ∈ X with x 1 , x 2


∃Ui ⊂ X open with x i ∈ Ui , such that U1 ∩ U2 = ∅.

The Hausdorff property says that any two distinct points b b

can be separated by disjoint open neighborhoods. x1 x2

U1 U2

Examples 1.27. 1.) Spaces with discrete topology are Hausdorff spaces, simply put Ui = {x i }.

2.) Let #X ≥ 2. Then X with the coarse topology is not a Hausdorff space.

3.) If the topology of X is induced by a metric d, then X is Hausdorff. Namely, for x 1 , x 2 put
r := d(x 1, x 2 ) > 0. Then the open balls Ui = B(x i , r/2) separate x 1 and x 2 .

4.) Let X be Hausdorff and let Y ⊂ X any subset. Then Y with its induced topology is again
Hausdorff.

10
1.5 Quotient spaces

Proposition 1.28. Let X be a Hausdorff space. Let Y ⊂ X be a compact subset. Then Y is a


closed subset.

Proof. Let p ∈ X \ Y . Then for every y ∈ Y the Hausdorff property tells us that there exist open
subsets Uy, p, Vy, p ⊂ X such that y ∈ Uy, p , p ∈ Vy, p and Uy, p ∩ Vy, p = ∅. Since Y is compact
there exist y1, . . . , yn ∈ Y such that Y ⊂ Uy1, p ∪ · · · ∪ Uy n, p . Now put Vp := ∩nj=1 Vy j , p . Since
Vp is a finite intersection of open subsets it is open itself. Moreover, p ∈ Vp . Now

Y ∩ Vp ⊂ (Uy1, p ∪ · · · ∪ Uy n, p ) ∩ Vp ⊂ (Uy1, p ∩ Vy1, p ) ∪ · · · ∪ (Uy n, p ∩ Vy n, p ) = ∅,

hence Vp ⊂ X \ Y . Therefore

X \ Y = ∪ p ∈X \Y Vp ⊂ X is open.

Thus Y ⊂ X is closed. 

Corollary 1.29. Let X be compact, Y Hausdorff. If f : X → Y is continuous and bijective,


then f is a homeomorphism.

Proof. We have to show that ∀A ⊂ X closed f (A) ⊂ Y is closed. Now let A ⊂ X be closed.
Since X is compact, A is compact and also the image f (A) is compact. Since Y is a Hausdorff
space we conclude that f (A) ⊂ Y is closed. 

1.5 Quotient spaces


Let X be a topological space. Let ∼ be an equivalence relation on X . For any x ∈ X let [x] be
the equivalence class of x. Denote the set of equivalence classes by

X/∼ := [x] | x ∈ X

Let π : X → X/∼ , x 7→ [x]. Define U ⊂ X/∼ to be open iff π −1 (U ) ⊂ X is open. One can
easily check that this defines a topology on X/∼ . Observe that π : X → X/∼ is continuous by
definition.
We now state the universal property of the quotient topology: For any topological space Y and
for any maps f : X → Y and f¯ : X/∼ → Y such that the diagram
f
X /Y
⑤⑤=

π ⑤⑤⑤
 ⑤⑤ f¯
X/∼

commutes, f is continuous iff f¯ is continuous.

11
1 Set Theoretic Topology

Example 1.30
Let X = [0, 1] and let the equivalence relation be given by x ∼ x ∀x ∈ X and 0 ∼ 1.

This equivalence relation identifies the end points


of the interval. We expect to obtain a topological
space homeomorphic to the circle. 0 1

Indeed, we can construct such a homeomorphism. Consider f : X → S1 given by


f (x) = (cos(2πx), sin(2πx)). Since f (0) = f (1) there exists a unique f¯ : X/∼ → S1 such
that the diagram
f
[0, 1] / S1
① ;
①①①
π ①
①①
 ①① f¯
[0, 1]/∼
commutes. From the universal property we know that f¯ is continuous because f is continuous.
Moreover, f¯ : X/∼ → S1 is bijective. Since X is compact (Heine Borel) π(X ) = X/∼ is
compact as well. Since R2 is Hausdorff, S1 is also Hausdorff. Hence Corollary 1.29 applies and
f¯ : X/∼ → S1 is a homeomorphism.

If X is a topological space and Y ⊂ X a subset then X/Y := X/∼ where

x 1 ∼ x 2 ⇔ x 1 = x 2 or x 1, x 2 ∈ Y .

This equivalence relation identifies all points in Y to one point and performs no further identifi-
cations. Example 1.30 is of this form.

Example 1.31. R/[0,1] is homeomorphic to R.

Example 1.32. R/(0,∞) is not a Hausdorff space because the points [0] and [1] cannot be sepa-
rated.

1.6 Product spaces


Let X1, . . . , X n be topological spaces. We set

X := X1 × · · · × X n

12
1.7 Exercises

Definition 1.33. A subset U ⊂ X is called open (for the product topology) iff for all
p = (p1, . . . , pn ) ∈ U there exist open subsets Ui ⊂ X i with pi ∈ Ui and U1 × · · · × Un ⊂ U.

It is easy to check that this defines a topology on X .

Examples 1.34. 1.) If all X i carry the discrete topology then X carries the discrete topology.
Namely, the sets {(p1, . . . , pn )} are open, hence all subsets of the product space are open.

2.) If all X i carry the coarse topology then X carries the coarse topology.

3.) Rn × Rm ≈ Rn+m . To see this it is convenient to use the maximum metric on Rn to


characterize the standard topology.

Now we list some important properties of the product topology:


1.) The projection maps πi : X → X i , πi (x) = x i , are continuous because for any open subset
Ui ⊂ X i
πi−1 (Ui ) = X1 × · · · × X i−1 × Ui × X i+1 × · · · × X n
is open in X .

2.) Fix x i ∈ X i , i = {1, . . . , i 0 − 1, i 0 + 1, . . . , n}. Then the map ι : X i 0 → X is continuous where


ι(ξ) = (x 1, . . . , x i 0 −1, ξ, x i 0 +1, . . . , x n ).

3.) Universal property: for all topological spaces Y and for all maps

f = ( f 1, . . . , f n ) : Y → X = X1 × · · · × X n

the map f is continuous if and only if all f i : Y → X i are continuous.

4.) If all X i are compact then X is also compact.

5.) If all X i are Hausdorff then X is also Hausdorff.

1.7 Exercises
1.1. Determine all topologies on the set {1, 2, 3} and investigate which ones are homeomorphic.

1.2. Let W n = [−1, 1] × · · · × [−1, 1] ⊂ Rn and D n = {x ∈ Rn | k xk ≤ 1}. Show that D n and


W n are homeomorphic.

1.3. Let T be a topology on R which contains all half-open intervals (x, y] and [x, y), x < y.
Show that T is the discrete topology.

1.4. Show that (0, 1) and [0, 1]

13
1 Set Theoretic Topology

a) are not homeomorphic when equipped with the standard topologies induced by R;

b) are homeomorphic when equipped with the discrete topology.

1.5. Let D n be as in Exercise 1.2. Let x, y ∈ D̊ n , i.e., k xk, k yk < 1. Construct a homeomor-
phism ϕ : D n → D n with ϕ(x) = y and ϕ(z) = z for all z ∈ ∂D n , i.e., kzk = 1.

1.6. Let X be a topological space, let “∼” be an equivalence relation on X . On page 11 we


defined that U ⊂ X/ ∼ is called open if and only if π −1 (U ) ⊂ X is open where π : X → X/ ∼ is
the standard projection.
a) Show that X/ ∼ equipped with this system of open sets is a topological space.

b) Show that the universal property on page 11 determines the topology on X/ ∼ uniquely, i.e.,
if T is a topology on X/ ∼ such that π : (X, TX ) → (X/ ∼, T ) is continuous and if the
universal property holds for (X/ ∼, T ) then T is the topology defined above.

1.7. Let D n and W n be as in Exercise 1.2. Show:


a) D n /∂D n ≈ S n ;

b) W n /∂W n ≈ S n .

1.8. Let X = R/Q, i.e., the quotient space of R under the equivalence relation x ∼ y iff x − y ∈ Q.
Show that X has uncountably many elements but carries the coarse topology.

1.9. Let X = R2 /Z2 , i.e., the quotient space of R2 under the equivalence relation x ∼ y iff
x − y ∈ Z2 . Let Y = S1 × S1 with the product topology. Show that X and Y are homeomorphic.

1.10. Let X be a topological space. One obtains the cone C X over X by considering the cylinder
X × [0, 1] and then passing to the quotient C X = (X × [0, 1])/ ∼ where the equivalence relation
∼ is given by (x, t) ∼ (x ′, t ′ ) if and only if (x, t) = (x ′, t ′ ) or t = t ′ = 1. Show:
a) CS n ≈ D n+1 .

b) If X is compact so is C X .

c) If X is Hausdorff so is C X .

1.11. Let X be a topological space. One obtains the suspension ΣX of X as the quotient
ΣX = (X × [0, 1])/ ∼ where the equivalence relation ∼ is given by (x, t) ∼ (x ′, t ′ ) if and only if
(x, t) = (x ′, t ′ ) or t = t ′ = 1 or t = t ′ = 0.
If furthermore f : X → Y is a map then f × id : X × [0, 1] → Y × [0, 1] induces a map
Σ f : ΣX → ΣY . Show:
a) If f is continuous so is Σ f .

14
1.7 Exercises

b) ΣD n ≈ D n+1 .

c) ΣS n ≈ S n+1 .

15
2 Homotopy Theory

2.1 Homotopic maps

Definition 2.1. Let X and Y be topological spaces. Let A ⊂ X be a subset. Two continuous
maps f 0, f 1 : X → Y are called homotopic relative to A iff there exists a continuous map
F : X × [0, 1] → Y such that

(i) F (x, 0) = f 0 (x) ∀x ∈ X

(ii) F (x, 1) = f 1 (x) ∀x ∈ X

(iii) F (a, t) = f 0 (a) ∀a ∈ A, ∀t ∈ [0, 1]

In symbols, f 0 ≃ A f 1 . The map F is then called a homotopy relative to A.

Remark 2.2. f 0 ≃ A f 1 ⇒ f 0 A = f 1 A

Remark 2.3. If A = ∅, then we say “ f 0 and f 1 are homotopic” instead of “ f 0 and f 1 are
homotopic relative to ∅”. Similarly, we write “ f 0 ≃ f 1 ” instead of “ f 0 ≃∅ f 1 ”.

Examples 2.4. 1.) Let f 0, f 1 : X → Rn be continuous with f 0 A = f 1 A . Put


F : X × [0, 1] → Rn with F (x, t) := t f 1 (x) + (1 − t) f 0 (x). Then F is a homotopy from
f 0 to f 1 relative to A, hence f 0 ≃ A f 1 .

2.) Let f 0 : Rn → Y be continuous. Put F (x, t) := f 0 ((1 − t)x), then f 0 ≃ const map.

3.) Lef f = Exp : R → S1 ∈ C, Exp(x) = e2πi x . From the previous example we know
f ≃ const map, but we will shortly see that f ;Z const map.
Given two topological spaces X, Y we set

C (X, Y ) := { f : X → Y | f is continuous}

Lemma 2.5. Let X, Y be topological spaces, let A ⊂ X and let ϕ ∈ C (A, Y ). Then “≃ A ” is
an equivalence relation on { f ∈ C (X, Y ) | f | A = ϕ}.

17
2 Homotopy Theory

Proof. a) “≃ A ” is reflexive:
f ≃ A f , because we can put F (x, t) := f (x).

b) “≃ A ” is symmetric:
Let f ≃ A g. We have to show g ≃ A f . Let F : X × [0, 1] → X be a homotopy relative to
A from f to g. Put G(x, t) := F (x, 1 − t). This is a homotopy relative to A from g to f ,
therefore g ≃ A f .

c) “≃ A ” is transitive:
Let f ≃ A g and g ≃ A h. We have to show f ≃ A h. Let F : X × [0, 1] → Y be homotopy
relative to A from f to g and let G : Y × [0, 1] → X be homotopy relative to A from g to h.
Then put H : X × [0, 1] → Y ,


 F (x, 2t),
H (x, t) := 
0 ≤ t ≤ 1/2
 G(x, 2t − 1),
 1/2 ≤ t ≤ 1

This is a homotopy relative to A from f to h, thus f ≃ A h. 

Lemma 2.6. Let X, Y, Z be topological spaces. Let A ⊂ X, B ⊂ Y be subsets. Let f 0, f 1 ∈


C (X, Y ) be such that f 0 ≃ A f 1, f i (A) ⊂ B and let g0, g1 ∈ C (Y, Z ) be such that g0 ≃ B g1 .
Then g0 ◦ f 0 ≃ A g1 ◦ f 1 .

Proof. Let F : X × [0, 1] → Y be homotopy relative to A from f 0 to f 1 and G : Y × [0, 1] → Z


be homotopy relative to B from g0 to g1 Then H : X × [0, 1] → Z is a homotopy relative to A
from g0 ◦ f 0 to g1 ◦ f 1 where H (x, t) = G(F (x, t), t). 

Definition 2.7. A map f ∈ C (X, Y ) is called a homotopy equivalence iff there exists a g ∈
C (Y, X ) such that g◦ f ≃ id X and f ◦g ≃ idY If there exists a homotopy equivalence f : X → Y
then X and Y are homotopy equivalent. In symbols, X ≃ Y .

Remark 2.8. This defines an equivalence relation on the class of topological spaces.

Remark 2.9. Obviously, homeomorphic implies homotopy equivalent, in short,

X ≈ Y ⇒ X ≃ Y.

Example 2.10. Euclidean space is homotopy equivalent to a point, Rn ≃ {0}. Namely, put
f : {0} → Rn, f (0) = 0, and g : Rn → {0}, g(x) = 0. Then g ◦ f = id {0} and f ◦ g ≃ idRn by
Example 2.4.1. Remark 2.8 implies Rn ≃ Rm for all n, m ∈ N.

18
2.1 Homotopic maps

Definition 2.11. A topological space is called contractible iff it is homotopy equivalent to


{point}.

Definition 2.12. Let A ⊂ X and let ι : A → X be the inclusion map. Then A is called

1.) a retract of X iff there exists r ∈ C (X, A) such that r A = id A , i.e. r ◦ ι = id A . Then r is
called a retraction from X to A.

2.) a deformation retract of X iff there exists a retraction r : X → A such that ι ◦ r ≃ id X .

3.) a strong deformation retract of X iff there exists a retraction r : X → A such that
ι ◦ r ≃ A id X .

Examples 2.13. 1.) Let X any topological space, let A = {x 0 } ⊂ X consist of just one point.
Then r : X → A, r (x) = x 0 , is a retraction, hence A is a retract of X . The one-pointed set A
is a deformation retract of X iff X is contractible.
2.) Let X = Rn+1 \ {0} and A = S n . Consider the map r : X → A with r (x) = k xx k .
The composition ι ◦ r : Rn+1 \ {0} → Rn+1 \ {0} then satisfies ι ◦ r = k xx k . The map
F ∈ C (Rn+1 \ {0} × [0, 1], Rn+1 \ {0}) given by
x
F (x, t) = t x + (1 − t)
k xk
is continuous and satisfies
F (x, 0) = (ι ◦ r)(x), F (x, 1) = id(x)
for all x ∈ Rn+1 \ {0} and
F (a, t) = a, a ∈ Sn
We thus conclude that ι ◦ r ≃S n idRn+1\{0} , hence S n is a strong deformation retract of
Rn+1 \ {0}.

The difference between a deformation retract and a strong deformation retract is rather subtle.

Example 2.14
We consider the comb space given by
( 1 )
X = (x, y) ∈ R2 | 0 ≤ y ≤ 1 and (x = 0 or x = for some n ∈ N)
( ) n
2
∪ (x, y) ∈ R | y = 0 and 0 ≤ x ≤ 1

The space X is a bounded and closed subset of R2 , hence compact. Let


the set A be given by A = {(0, 1)}.

19
2 Homotopy Theory

A
1

......

Z
1 1 1
0 4 3 2 1
First we show that A is a deformation retract of X . The map F : X × [0, 1] → X given by
F (x, y, t) := (x, (1 − t)y) is continuous and satisfies

F (x, y, 0) = (x, y), F (x, y, 1) = (x, 0) .

Therefore id X ≃ Inclusion Z →X ◦ π where π : X → [0, 1] × {0} =: Z is the projection π(x, y) =


(x, 0). Moreover, we have π ◦ Inclusion Z →X = id Z . This shows that π is a homotopy equivalence
between X and Z. Hence X ≃ Z ≈ [0, 1] ≃ {pt}, which means that X is contractible. Therefore
A is a deformation retract of X .
Now we show that A is not a strong deformation retract of X . Suppose it were, then there would
exist a continuous map G : X × [0, 1] → X , such that for all t ∈ [0, 1] and all (x, y) ∈ X

G(x, y, 0) = (x, y), G(0, 1, t) = (0, 1).

Since X × [0, 1] is compact the map G would be uniformly continuous. Therefore for ε = 1/2
we can find δ > 0 such that
G(x, y, t) − G(x ′, y ′, t ′ ) < 1
2
whenever |x − x ′ | < δ, |y − y ′ | < δ and |t − t ′ | < δ.

Now choose n so large that n1 < δ and consider A


! 1
1 
(x, y) = , 1 , x ′, y ′ = (0, 1), t = t ′ .
n

Then
1
G , 1, t − G(0, 1, t) < ,
!
1
∀t ∈ [0, 1] .
n 2
Hence G( n1 , 1, t) ∈ B((0, 1), 12 ) for all t ∈ [0, 1]. Z
1
0 n 1

20
2.2 The fundamental group

On the other hand the mapping t 7→ G( n1 , 1, t) is a continuous path in X from ( n1 , 1) to (0, 1) and
must for some t take values in Z.

Remark 2.15. We have the following scheme of implications:

A is a strong deformation retract of X


A is a deformation retract of X



:
A is a retract of X ; A≃X

That both possible implications in the bottom row do not hold in general can be seen by
counterexamples. Let A = {x 0 } be a point in X and let X be not contractible. Then A is a retract
of X but A and X are not homotopically equivalent. This is a counterexample for the implication
“ ⇒ ”.
A counterexample for the other direction “ ⇐ ” is given by X = [0, 1] × [0, 1] ⊂ R2 and
A = comb space. Then X and A are contractible, hence X ≃ A, but one can show that there is
no retraction X → A.

2.2 The fundamental group

Definition 2.16. Let X be a topological space and let x 0, x 1, x 2 ∈ X . Put

Ω (X ; x 0, x 1 ) := {ω ∈ C ([0, 1], X ) | ω(0) = x 0, ω(1) = x 1 } and


Ω (X ; x 0 ) := Ω (X ; x 0, x 0 ) .

Elements of Ω(X ; x 0, x 1 ) are called paths and elements of Ω(X ; x 0 ) loops.


For ω ∈ Ω(X ; x 0, x 1 ) and η ∈ Ω(X ; x 1, x 2 ) define ω ⋆ η ∈ Ω(X ; x 0, x 2 ) by


ω(2t),
(ω ⋆ η)(t) = 
0 ≤ t ≤ 1/2,
 η(2t − 1),
 1/2 ≤ t ≤ 1.

Moreover, we consider ω−1 ∈ Ω(X ; x 1, x 0 ) with ω−1 (t) = ω(1 − t) and ε x0 ∈ Ω(X ; x 0 ) where
ε x0 (t) = x 0 .

21
2 Homotopy Theory

x1
X

η x0 x2
ω

0 1 0 1
t → 2t t → 2t − 1

0 1

Definition 2.17. For ω ∈ Ω(X ; x 0 ) denote by [ω] the homotopy class of ω relative to {0, 1}.
Then
π1 (X ; x 0 ) := {[ω] | ω ∈ Ω(X ; x 0 )}
is called the fundamental group of (X, x 0 ).

Lemma 2.18. For ω, ω ′, η, η ′, ζ ∈ Ω(X, x 0 ) we have

(i) If ω ≃ {0,1} ω ′ and η ≃ {0,1} η ′, then ω ⋆ η ≃ {0,1} ω ′ ⋆ η ′;

(ii) ε x0 ⋆ ω ≃ {0,1} ω ≃ {0,1} ω ⋆ ε x0 ;

(iii) ω ⋆ ω−1 ≃ {0,1} ε x0 ≃ {0,1} ω−1 ⋆ ω;

(iv) (ω ⋆ η) ⋆ ζ ≃ {0,1} ω ⋆ (η ⋆ ζ ).

Proof. The proof will be given graphically. In the following diagrams we draw the domain of
the required homotopies. The horizontal axis denotes the loop parameter whereas the vertical
axis represents the deformation parameter. The interpolations are piecewise linear. The red area

22
2.2 The fundamental group

gets mapped to x 0 .

(i) We run the loop parameter with twice the speed.

ω′ η′

ω η
(ii) The first diagram shows ε x0 ≃ {0,1} ω, the second proves that ω ≃ {0,1} ε x0 .

ω ω

ε x0 ω ω ε x0

(iii) The statement ω ⋆ ω−1 ≃ {0,1} ε x0 is proven by the following diagram


ε x0

ω(1 − s)
s

ω ω−1

23
2 Homotopy Theory

For any s ∈ [0, 1] the corresponding ”blue“ line segment gets mapped to ω(1 − s). To prove the
statement ε x0 ≃ {0,1} ω−1 ⋆ ω interchange the roles of ω and ω−1 .

(iv) The last assertion follows from


ω η ζ

ω η ζ 

Lemma 2.18 has the following implications:

(i) ⇒ [ω] · [η] := [ω ⋆ η] is well defined;

(ii) ⇒ [ε x0 ] · [ω] = [ω] = [ω] · [ε x0 ];

(iii) ⇒ [ω] · [ω−1 ] = [ε x0 ] = [ω−1 ] · [ω];

(iv) ⇒ ([ω] · [η]) · [ζ] = [ω] · ([η] · [ζ]).

Hence π1 (X ; x 0 ) together with “ · ” is a group with neutral element 1 := [ε x0 ] and inverse element
[ω]−1 = [ω−1 ].
To any topological space with a preferred point we have associated a group, the fundamental
group of the space with that point. Now we consider continuous maps. Let f ∈ C (X, Y ) and
x 0 ∈ X . We put f (x 0 ) =: y0 ∈ Y . If ω ≃ {0,1} ω ′ and H is a homotopy between them relative to
{0, 1} then f ◦H is a homotopy between f ◦ω and f ◦ω ′ relative to {0, 1}. Hence f ◦ω ≃ {0,1} f ◦ω ′ .
Therefore we have a well-defined map f # : π1 (X ; x 0 ) → π1 (Y ; y0 ), f # ([ω]) = [ f ◦ ω].

Lemma 2.19. (i) The map f # : π1 (X ; x 0 ) → π1 (Y ; y0 ) is a group homomorphism.

(ii) It has the functorial properties


a) ( f ◦ g)# = f # ◦ g# ;
b) (id X )# = idπ 1 (X ;x0 ) .

(iii) If f ≃ {x0 } f ′ then f # = f #′.

24
2.2 The fundamental group

Proof. (i) From the definitions we have f ◦ (ω ⋆ η) = ( f ◦ ω) ⋆ ( f ◦ η) and hence


f # ([ω] · [η]) = f # ([ω ⋆ η]) = [ f ◦ (ω ⋆ η)] = [( f ◦ ω) ⋆ ( f ◦ η)] = f # ([ω]) · f # ([η]).

(ii) Assertion b) being obvious we compute a):


( f ◦ g)# ([ω]) = [( f ◦ g) ◦ ω] = [ f ◦ (g ◦ ω)] = f # ([g ◦ ω]) = f # (g# ([ω])) = ( f # ◦ g# )([ω]) .

(iii) By Lemma 2.6, f ≃ {0,1} f ′ implies f ◦ ω ≃ {0,1} f ′ ◦ ω. We conclude


f # ([ω]) = [ f ◦ ω] = [ f ′ ◦ ω] = f #′ ([ω])
which proves the statement. 

Corollary 2.20. If f : X → Y is a homeomorphism with f (x 0 ) = y0 then

f # : π1 (X, x 0 ) → π1 (Y, y0 )

is a group isomorphism.

Proof. We only use the functorial properties:


( f −1 )# ◦ ( f # ) = ( f −1 ◦ f )# = (id X )# = idπ 1 (X ;x0 )
and similarly one sees that ( f # ) ◦ ( f −1 )# = idπ 1 (Y ;y0 ) . Thus f ♯ is an isomorphism with ( f # ) −1 =
( f −1 )# . 

Now we want to deal with the question to what extent π1 (X ; x 0 ) depends on the choice of x 0 .

X
To study this question let x 0, x 1 ∈ X and assume ω
there exists a path γ ∈ Ω(X ; x 0, x 1 ). If such a
path does not exist π1 (X ; x 0 ) and π1 (X ; x 1 ) are x1
not related. γ
x0

Look at the map Φγ : Ω(X ; x 1 ) → Ω(X ; x 0 ) where ω →7 (γ ⋆ ω) ⋆ γ −1 . Applying


Lemma 2.18 twice we know that if ω ≃ {0,1} ω ′ then γ ⋆ ω ≃ {0,1} γ ⋆ ω ′ and hence
(γ ⋆ ω) ⋆ γ −1 ≃ {0,1} (γ ⋆ ω ′ ) ⋆ γ −1 . Thus the map
Φ̂γ : π1 (X ; x 1 ) → π1 (X ; x 0 ),
[ω] 7→ [Φγ (ω)] = [(γ ⋆ ω) ⋆ γ −1 ],
is well defined.

25
2 Homotopy Theory

Proposition 2.21. Let X be a topological space, assume x 0, x 1 ∈ X and γ, γ ′ ∈ Ω(X ; x 0, x 1 ).


Then

1. The map Φ̂γ : π1 (X ; x 1 ) → π1 (X ; x 0 ) is a group isomorphism.

2. If γ ≃ {0,1} γ ′ then Φ̂γ = Φ̂γ′ .

3. For β ∈ Ω(X ; x 1, x 2 ) we have Φ̂γ⋆β = Φ̂γ ◦ Φ̂β .

4. For the constant path we have Φ̂ε x0 = idπ 1 (X ;x0 ) .

5. For any [ω] ∈ π1 (X ; x 1 ) we have Φ̂γ′ ([ω]) = κ · Φ̂γ ([ω]) · κ −1 where κ = [γ ′ ⋆ γ −1 ] ∈


π1 (X ; x 0 ).

Proof. a) Assertions 2., 3., and 4. follow directly from Lemma 2.18 and the definitions.

b) The map Φ̂γ is a group homomorphism because



f g
Φ̂γ [ω] · [η] = Φ̂γ ([ω ⋆ η])
= (γ ⋆ (ω ⋆ η)) ⋆ γ −1
f g
γ ⋆ ((ω ⋆ (γ −1 ⋆ γ)) ⋆ η) ⋆ γ −1
f g
=

= (γ ⋆ ω) ⋆ γ −1 ⋆ (γ ⋆ η) ⋆ γ −1
= Φ̂γ ([ω]) · Φ̂γ ([η]).

The map Φ̂γ is bijective, because


3. 2. 4.
Φ̂γ ◦ Φ̂γ−1 = Φ̂γ⋆γ−1 = Φ̂ε x0 = idπ 1 (X ;x0 ) .

and similarly Φ̂γ−1 ◦ Φ̂γ = idπ 1 (X ;x1 ) . This proves 1.

c) We compute
f g f g f g
κ · Φ̂γ ([ω]) · κ −1 = γ ′ ⋆ γ −1 · γ ⋆ ω ⋆ γ −1 · γ ⋆ (γ ′ ) −1
f g
= γ ′ ⋆ γ −1 ⋆ γ ⋆ ω ⋆ γ −1 ⋆ γ ⋆ (γ ′ ) −1
f g
= γ ′ ⋆ ω ⋆ (γ ′ ) −1
= Φ̂γ′ ([ω]).

26
2.2 The fundamental group

Proposition 2.22. Let X, Y be topological spaces and x 0 ∈ X . Let f , g ∈ C (X, Y ) and let
H : X × [0, 1] → Y be a homotopy from f to g. Define η ∈ Ω(Y ; f (x 0 ), g(x 0 )) by

η(s) := H (x 0, s).

Then the following diagram commutes:

π1 (X ; x 0P)
PPP
g#
PPPf#
PPP
 PP(
π1 (Y ; g(x 0 )) / π1 (Y ; f (x 0 ))
Φ̂η

Proof. Let [ω] ∈ π1 (X ; x 0 ) and put F : [0, 1] × [0, 1] → Y with F (t, s) := H (ω(t), s).

Y
g◦ω

η
s F

f ◦ω
t

The deformation indicated in the figure yields a homotopy re-


lative to {0, 1} from f ◦ ω to η ⋆ (g ◦ ω) ⋆ η −1 . We conclude
that

f # [ω] = [ f ◦ ω] = [η ⋆ (g ◦ ω) ⋆ η −1 ] = Φ̂η (g# ([ω])) .

| |
1 2
3 3


Now we can improve Corollary 2.20 and show that homotopy equivalent spaces have isomorphic
fundamental groups.

27
2 Homotopy Theory

Theorem 2.23. Let f : X → Y be a homotopy equivalence. Then

f # : π1 (X ; x 0 ) → π1 (Y ; f (x 0 ))

is an isomorphism for all x 0 ∈ X .

Proof. Let g : Y → X be a homotopy inverse of f , i.e., g ◦ f ≃ id X and f ◦ g ≃ idY . We know


by Proposition 2.22 that for a suitable η ∈ Ω(X ; x 0, f (g(x 0 )))

g# ◦ f # = (g ◦ f )# = Φ̂η ◦ (id X )# = Φ̂η ◦ idπ 1 (X ;x0 ) = Φ̂η .

Hence g# ◦ f # is an isomorphism. In particular, f # is injective. Similarly, we can show that


f # ◦ g# is an isomorphism, hence f # is surjective. Therefore f # is an isomorphism. 

Corollary 2.24. If A ⊂ X is a deformation retract then the inclusion ι : A → X induces an


isomorphism ι# : π1 (A, x 0 ) → π1 (X ; x 0 ) for any x 0 ∈ A.

Proof. If A ⊂ X is a deformation retract then the map ι : A → X is a homotopy equivalence.


Theorem 2.23 yields the claim. 

Example 2.25. If X is contractible then the one-point set A = {x 0 } ⊂ X is a deformation retract.


Hence π1 (X ; x 0 ) ≃ π1 (A; x 0 ) = {[ε x0 ]} = {1}. Thus contactible spaces have trivial fundamental
group.

2.3 The fundamental group of the circle


Recall the map Exp : R → S1 ⊂ C where Exp(t) = e2πit .

Lemma 2.26. Let t 0 ∈ R and z0 = Exp(t 0 ) ∈ S1 . Then for all f ∈ C (S1, S1 ) with f (1) = z0
there exists a unique fˆ ∈ C (R, R) with fˆ(0) = t 0 such that the following diagram commutes:


R /R

Exp Exp
 
S1 / S1
f

Proof. a) First we show uniqueness of the map fˆ.

28
2.3 The fundamental group of the circle

Assume that besides fˆ there is a second map f˜ ∈ C (R, R) with the same proper-
ties as fˆ. The equality of Exp(x) = Exp(x ′ ) is equivalent to x − x ′ ∈ Z. Since
Exp( fˆ(t)) = f (Exp(t)) = Exp( f˜(t)) we deduce that fˆ(t) − f˜(t) ∈ Z for all t ∈ R. Since both
f˜, fˆ are continuous it follows that fˆ − f˜ is constant. Finally, we know that fˆ(0) = t 0 = f˜(0),
hence f˜ = fˆ.

b) Now we show existence of fˆ.


Since S1 is compact, f is uniformly continuous. Since Exp is also uniformly continuous, the
composition f ◦ Exp : R → S1 is uniformly continuous. Hence there exists ε > 0 such that
f (Exp(I)) ⊂ S1 is contained in a semi-circle for every interval I ⊂ R with length ≤ ε.
For any closed semi-circle C ⊂ S1 the pre-image Exp−1 (C) ⊂ R is the disjoint union of
compact intervals of length 21 . More precisely,
[" 1
#
−1
Exp (C) = t 1 + k, t 1 + k + .
k ∈Z
2

Moreover, the restriction of Exp to each of these intervals is a homeomorphism onto its image,
" #
Exp [t 1 +k, t 1 +k+ 1 ]
1 ≈
: t 1 + k, t 1 + k + −→ C.
2 2
Its inverse can written down explicitly in terms of logarithms but we will not need this.
Now we construct fˆ step by step.

Since f (Exp([0, ε])) is contained in a closed semi-circle C0 we can define fˆ on [0, ε] by

fˆ := (Exp | I0 ) −1 ◦ f ◦ Exp,

1
where I0 ⊂ R is the compact interval of length 2 with Exp(I0 ) = C0 and t 0 ∈ I0 . This insures
that

fˆ(0) = (Exp | I0 ) −1 ◦ f ◦ Exp(0) = (Exp | I0 ) −1 ◦ f (1) = (Exp | I0 ) −1 (z0 ) = t 0 .

Put t 1 := fˆ(ε).
Next, f (Exp([ε, 2ε])) is contained in a closed semi-circle C1 and we define fˆ on [ε, 2ε] by

fˆ := (Exp | I1 ) −1 ◦ f ◦ Exp

where I1 ⊂ R is the compact interval of length 21 with Exp(I1 ) = C1 and t 1 ∈ I1 . This insures
that the two definitions of fˆ at ε agree so that we obtain a continuous function fˆ : [0, 2ε] → R.
Repeating this procedure infinitely many times we can extend fˆ continuously to [0, ∞) → R.
The extension to the left is done similarly so that we obtain a continuous function fˆ : R → R.
Commutativity of the diagram holds by construction. 

29
2 Homotopy Theory

Definition 2.27. For a map f ∈ C (S1, S1 ) a map fˆ ∈ C (R, R) for which the diagram in
Lemma 2.26 commutes is called a lift of f .

Example 2.28. Consider the map f n : S1 → S1 with f n = z n, n ∈ Z. Then we have



f n Exp(t) = Exp(t) n = Exp(nt) .

Hence the map fˆn given by fˆn (t) = nt is a lift of f n .

Definition 2.29. For f ∈ C (S1, S1 ) we call

deg( f ) := fˆ(1) − fˆ(0)

the degree of f , where fˆ is a lift of the map f .

Remark 2.30
1.) We have seen that different lifts of f differ by an additive constant in Z. Hence the degree
deg( f ) is well defined, independently of the choice of lift fˆ.

2.) The degree deg( f ) is an integer because


     
Exp fˆ(1) = f Exp(1) = f (1) = f Exp(0) = Exp fˆ(0)

Hence we have deg( f ) = fˆ(1) − fˆ(0) ∈ Z.

3.) The map t 7→ fˆ(t + 1) − fˆ(t) is continuous and takes values in Z, by the same argument as
above. We conclude that fˆ(t + 1) − fˆ(t) = deg( f ) for all t ∈ R.

4.) For k ∈ Z we compute

fˆ(t 0 + k) − fˆ(t 0 ) = fˆ(t 0 + k) − fˆ(t 0 + (k − 1))


+ fˆ(t 0 + (k − 1)) − fˆ(t 0 + (k − 2))
+··· +
+ fˆ(t 0 + 1) − fˆ(t 0 )
(3)
= k deg( f ) .

5.) For f , g ∈ C (S1, S1 ) and lifts fˆ, ĝ we compute


     
Exp fˆ + ĝ (t) = Exp fˆ(t) Exp ( ĝ(t)) = f Exp(t) g Exp(t) .

Hence fˆ + ĝ is a lift of f g and we get the following formula for the degree

deg( f g) = fˆ(1) + ĝ(1) − fˆ(0) + ĝ(0) = deg( f ) + deg(g) .

30
2.3 The fundamental group of the circle

6.) For f , g ∈ C (S1, S1 ) and lifts fˆ, ĝ we compute


     
Exp fˆ ĝ(t) = f Exp ĝ(t) = f g Exp(t)

and therefore fˆ ◦ ĝ is a lift of f ◦ g. For the degree of f ◦ g this means


    (4)
deg( f ◦ g) = fˆ ĝ(1) − fˆ ĝ(0) = fˆ ĝ(0) + deg(g) − fˆ ĝ(0) = deg(g) deg( f ) ,

hence deg( f ◦ g) = deg(g) deg( f ).


7.) Let f ∈ C (S1, S1 ) with deg( f ) , 0. We show that f must then be surjective.
Namely, let fˆ be a lift of f . Then deg( f ) = fˆ(1) − fˆ(0) is an integer, not equal to 0. Then,
if fˆ(1) > fˆ(0), the whole interval I := [ fˆ(0), fˆ(1)] must be contained in the image of fˆ by
the intermediate value theorem. If fˆ(1) < fˆ(0) this holds for I := [ fˆ(1), fˆ(0)]. In either
case I is an interval of length at least 1, hence Exp(I) = S1 . We conclude

S1 = Exp(I) ⊂ Exp(im( fˆ)) = im( f ).

Thus f is onto.

Example 2.31. For the map f n : S1 → S1 with f n (z) = z n a lift is given by fˆ(t) = nt so that its
degree is deg( f n ) = fˆn (1) − fˆn (0) = n.

Lemma 2.32. Let f , g ∈ C (S1, S1 ). If f ≃ g, then we have deg( f ) = deg(g).

Proof. Let F : S1 × [0, 1] → S1 be a homotopy from f to g. Since S1 × [0, 1] is compact the


map F is uniformly continuous. Hence, there exists a δ > 0 such that

|F (z, s) − F (z, s ′ )| < 1

whenever z ∈ S1 and s, s ′ ∈ [0, 1] with |s − s ′ | < δ. For such s, s ′ the map


F (z, s)
z 7→ : S1 → S1
F (z, s ′ )

 (·, s)  and not surjective because −1 is not contained in the image. Hence, by Remark 2.30,
is continuous
deg FF(·, s ′ ) = 0. We now compute using 2.30.5:
! !
F (·, s) ′ F (·, s)
deg(F (·, s)) = deg · F (·, s ) = deg + deg F (·, s ′ ) = deg F (·, s ′ ) .
F (·, s ′ ) F (·, s ′ )
We see inductively that

deg( f ) = deg F (·, 0) = deg F (·, s1 ) = · · · = deg F (·, 1) = deg(g)

where 0 = s0 < s1 < · · · < sr = 1 is a partition of the unit interval [0, 1] satisfying
|si+1 − si | < δ. 

31
2 Homotopy Theory

Corollary 2.33. Let f ∈ C (S1, S1 ) be such that f = g S 1 where g ∈ C (D 2, S1 ) then


deg( f ) = 0.

Proof. The map F ∈ C (S1 × [0, 1], S1 ), F (z, s) := g(sz), defines a homotopy from a constant
map to f . By Lemma 2.32 we conclude that deg( f ) = deg(const) = 0. 

Let us give several applications of the concept of the degree.

Theorem 2.34 (Fundamental theorem of algebra). Every non-constant complex polyno-


mial has a root.

Proof. Suppose we are given a non-constant polynomial

p(z) = an z n + an−1 z n−1 + · · · + a1 z + a0, a j ∈ C, an , 0, n ≥ 1.

Since dividing by an does not change the roots, we assume without loss of generality that an = 1.
Now assume that p has no roots. Then the map f : C → S1 with f (z) = | p(z)p(z) | is a well-defined

continuous map. To compute deg( f S 1 ) consider

s n p(z/s) z n + san−1 z n−1 + · · · + s n a0


F (z, s) := = .
|s n p(z/s)| |z n + san−1 z n−1 + · · · + s n a0 |

The map F ∈ C (S1 × [0, 1], S1 ) is a homotopy from f n (z) = z n to f S 1 . Computing its degree
we find deg( f S 1 ) = deg( f n ) = n ≥ 1.
On the other hand, f is a continuous map defined on all of C, hence Corollary 2.33 implies
deg( f S 1 ) = 0. This is a contradiction, thus p must have a root. 

Lemma 2.35. Suppose the map f ∈ C (S1, S1 ) satisfies f (−z) = − f (z) for all z ∈ S1 . Then
the degree deg( f ) is odd.

Proof. Lef fˆ be a lift of the map f . We compute


          
f − Exp(t) = f Exp 21 Exp(t) = f Exp t + 21 = Exp fˆ t + 12 .

Moreover,
        
− f Exp(t) = − Exp fˆ(t) = Exp 21 Exp fˆ(t) = Exp fˆ(t) + 21 .

32
2.3 The fundamental group of the circle

    
From f (− Exp(t)) = − f (Exp(t)) we conclude Exp fˆ t + 21 = Exp fˆ(t) + 21 and hence
 
fˆ t + 12 − fˆ(t) + 21 =: k (t)

is an integer for every t. Due to the continuity of the map, k (t) it is constant, k (t) = k. Hence
fˆ(t + 12 ) − fˆ(t) = k + 12 for all t ∈ R. Now we can compute
      
deg( f ) = fˆ(1) − fˆ(0) = fˆ(1) − fˆ 21 + fˆ 21 − fˆ(0)
= (k + 12 ) + (k + 21 ) = 2k + 1

which proves the assertion. 

Theorem 2.36 (Borsuk-Ulam). Let f ∈ C (S2, R2 ) satisfy f (−x) = − f (x) for all x ∈ S2 .
Then f has a zero.

Proof. Assume that the map f has no zero. Then the map g : S2 → S1 with g(x) := k ff (x) (x)
k
is defined and continuous. Moreover, g satisfies g(−x) = −g(x) for all x ∈ S 2 . Now define a
p
map G : D 2 → S1 by G(y) = g(y, 1 − k yk 2 ). The map G ∈ C (D 2, S1 ) has the property that
G S 1 = g|S 1 . By Corollary 2.33 we know that deg(g S 1 ) = 0. On the other hand we know by
Lemma 2.35 that deg(g S 1 ) is odd, which gives a contradiction. 

Corollary 2.37. Let f ∈ C (S2, R2 ). Then there exists a point x 0 ∈ S2 with f (−x 0 ) = f (x 0 ).

Proof. Put g(x) := f (x) − f (−x). Then the map g ∈ C (S2, R2 ) satisfies g(−x) = −g(x)
for all x ∈ S2 . Hence by the Borsuk-Ulam Theorem 2.36 there exists an x 0 ∈ S2 with
0 = g(x 0 ) = f (x 0 ) − f (−x 0 ), which proves the theorem. 

Remark 2.38. In particular, the map f ∈ C (S2, R2 ) cannot be injective. Thus the sphere S2
cannot be homeomorphic to a subset of R2 . This also shows that R3 cannot be homeomorphic to
R2 .

Now suppose you have a sandwich consisting of bread,


ham, and cheese. You want to share it evenly with your
friend. Can you cut the sandwich into two pieces such
that each piece contains the same amount of bread, the
same amount of ham and the same amount of cheese?
The following theorem tells us that it is possible.

33
2 Homotopy Theory

Theorem 2.39 (Ham-Sandwich-Theorem). Let A, B, C ⊂ R3 be open and bounded. Then


there exists an affine hyperplane H ⊂ R3 such that each of the sets is divided into pieces of
equal volume.

Proof. The proof consists of four steps.

a) For each x ∈ S2 and t ∈ R we define the affine H x, t


hyperplane

H x, t := {y ∈ R3 | hy, xi = t} . tx • +
H x, t
S2
•x
It is clear that H−x,−t = H x, t . We define the
half space H x,+ by
t
0•

+ 3
H x, t := {y ∈ R | hy, xi ≥ t}.

H x, t ′
H x, t
b) Now fix x ∈ S2 . Look at the function
a x : R → R defined by
+
a x (t) := vol(A ∩ H x, t ). B(0, R A )

It satisfies a−x (t) + a x (−t) = vol(A) A


and is monotonically decreasing.
Since A is bounded there exists
R A > 0 such that A ⊂ B(0, R A ). |t − t ′ |

For t < t ′ ∈ R we have


|a x (t ′ ) − a x (t)| = vol(A ∩ (H x,
+ + 2 ′
t \ H x, t ′ )) ≤ πR A · |t − t | .

Thus a x is Lipschitz continuous.

vol(A)
αx
Moreover, a x (t) = vol(A) for t ≪ 0
and a x (t) = 0 for t ≫ 0.

t −x t +x

34
2.3 The fundamental group of the circle

c) By continuity and monotonicity the pre-image of any value under α x is a closed interval. In
particular, a−1 − +
x (vol(A)/2) = [t x , t x ].
t − +t +
Now put α(x) := x 2 x . Hence H x,α(x) divides A into two pieces of equal volume. Moreover,
α(−x) = −α(x) for all x ∈ S2 and it is not hard to check that α is continuous. Similarly,
define the functions β for B and γ for C.

d) Consider the map f ∈ C (S2, R2 ) with f (x) = (α(x) − β(x), α(x) − γ(x)). We have

f (−x) = (α(−x) − β(−x), α(−x) − γ(−x))


= (−α(x) + β(x), −α(x) + γ(x))
= − f (x).

Thus the Borsuk-Ulam Theorem 2.36 applies and there exists a point x 0 ∈ S2 with

(0, 0) = f (x 0 ) = (α(x 0 ) − β(x 0 ), α(x 0 ) − γ(x 0 )).

Hence α(x 0 ) = β(x 0 ) = γ(x 0 ). Thus the hyperplance H x0,α(x0 ) does the job. 

Remark 2.40. The ham-sandwich theorem is optimal in the sense that it fails for more than three
sets in R3 .

Example 2.41. A ball B(x, r) ⊂ R3 with center x is cut into two halves of equal volume exactly
by those planes that contain x. If you choose four balls in R3 in such a way that their centers are
not contained in one plane, then no plane will cut them all into halves of equal volume.

In the following we want to use the concept of degree to determine π1 (S1 ; 1).

Exp !ω
a) For ω ∈ Ω(S1 ; 1) and I = [0, 1] consider the fol- S1 a❈o I / S1
❈❈ ④=
lowing diagram: ❈❈≈ ④④④
❈❈ ④④
Exp ❈  ④④ ω
I/∂I
Here ω and Exp are the continuous maps induced on the quotient space. They exist because
ω(0) = ω(1) and Exp(0) = Exp(1), compare Section 1.5. By Example 1.30 we know that
Exp : I/∂I → S1 is a homeomorphism. Now put

f ω := ω ◦ (Exp) −1 ∈ C S1, S1

and define deg(ω) := deg( f ω ). We have obtained a map

deg : Ω(S1 ; 1) → Z.

35
2 Homotopy Theory

b) Now suppose ω ≃ {0,1} ω ′ . We choose a homotopy F : I × I → S1 from ω to ω ′ relativ to


{0, 1}. Then the map G : S1 × I → S1 defined by G(z, s) := F ((Exp) −1 (z), s) is a homotopy
from f ω to f ω′ , hence f ω ≃ f ω′ . Therefore
deg(ω) = deg( f ω ) = deg( f ω′ ) = deg(ω ′ ) .
Hence we get a well-defined map
deg : π1 (S1 ; 1) → Z where [ω] 7→ deg(ω).

c) This map is surjective because for n ∈ Z we can consider the map ω(t) = Exp(nt). The
commutative diagram
z7→z n = f n (z)
S1 f◆◆ / S1
O
◆◆◆
◆◆◆ t 7→ω(t )=Exp(nt )=Exp(t ) n

t 7→Exp(t ) ◆◆◆

I
shows f ω (z) = z n = f n (z) and we get deg(ω) = deg( f n ) = n.
d) Now let ω, ω ′ ∈ Ω(S1 ; 1) and consider the map
f ω⋆ω′ (Exp(t)) = ω ⋆ ω ′ (t)

ω(2t),
= 
0 ≤ t ≤ 1/2
ω ′ (2t − 1), 1/2 ≤ t ≤ 1


 f ω (Exp(2t)),
= 
0 ≤ t ≤ 1/2
 f ω′ (Exp(2t − 1)), 1/2 ≤ t ≤ 1

Let fˆω , fˆω′ be lifts of f ω , f ω′ with fˆω′ (0) = fˆω (1). Then we have

 Exp( fˆω (2t)),
f ω⋆ω′ (Exp(t)) = 
0 ≤ t ≤ 1/2
 Exp( fˆω′ (2t − 1)), 1/2 ≤ t ≤ 1



!
= Exp   fˆω (2t), 0 ≤ t ≤ 1/2
 f ω′ (2t − 1),{z 1/2 ≤ t ≤ }
ˆ 1
|
=:g (t )

Note that the map g(t) is continuous because of fˆω′ (0) = fˆω (1). Thus g(t) is a lift of f ω⋆ω′ .
Now we compute the degree of ω ⋆ ω ′.
deg(ω ⋆ ω ′ ) = g(1) − g(0)
= fˆω′ (1) − fˆω (0)
= fˆω′ (1) − fˆω′ (0) + fˆω (1) − fˆω (0)
= deg(ω ′ ) + deg(ω)
Hence the map deg : π1 (S1 ; 1) → (Z, +) is a group homomorphism.

36
2.3 The fundamental group of the circle

e) Finally we compute its kernel. Let ω ∈ Ω(S1 ; 1) with deg(ω) = 0. Let fˆω be the lift of f ω
with fˆω (0) = 0. Since 0 = deg(ω) = fˆω (1) − fˆω (0) we have fˆω (1) = fˆω (0) = 0. Next
consider the continuous map F : I × I → S1 with F (t, s) := Exp(s fˆω (t)). It satisfies:

F (t, 0) = 1 = ε 1 (t) ,
F (t, 1) = f ω (Exp(t)) = ω(t) ,
F (0, s) = Exp(s · 0) = 1 ,
F (1, s) = Exp(s · fˆω (1)) = Exp(s 0) = 1 .

We conclude that ω ≃ {0,1} ε 1 , hence [ω] = [ε 1 ] = 1 ∈ π1 (S1 ; 1). Therefore the kernel is
trivial and the map deg : π1 (S1 ; 1) → Z is injective.

We summarize the discussion in the following

Theorem 2.42. The map deg : π1 (S1 ; 1) → Z is a group isomorphism.

Example 2.43. We already know that S1 is a strong deformation retract of C \ {0}. Hence the
inclusion ι : S1 → C \ {0} induces an isomorphism

ι# : π1 (S1 ; 1) → π1 (C \ {0}; 1)

and therefore π1 (C \ {0}; 1)  Z.

S1 = ∂D 2
Example 2.44
We show that S1 = ∂D 2 is not a retract of D 2 . Suppose the map D2
r : D 2 → S1 is a retraction and denote the inclusion by ι : S1 → D 2 .

Then we would have r ◦ ι = idS 1 , hence r# ◦ ι# = (r ◦ ι)# = (idS 1 )# = idπ 1 (S 1 ;1) . We then get a
contradiction because of the following diagram
ι# r#
Z  π(S1 ; 1) / π(D 2 ; 1)  {1} / π(S 1 ; 1)  Z
:

idπ 1
1 (S ;1)

As a corollary we get a proof of Brouwer’s fixed point theorem in dimension two. See page 92
for the theorem in general dimensions.

37
2 Homotopy Theory

Theorem 2.45 (Brouwer’s fixed point theorem in 2 dimensions). Let f : D 2 → D 2 be a


continuous map. Then f has a fixed point, i.e., there exists an x ∈ D 2 such that f (x) = x.

Proof. Assume that f ∈ C (D 2, D 2 ) has no fixed point. Then f (x) , x for all x ∈ D 2 so that we
can consider the half line emanating from f (x) through x. We let r (x) be its intersection point
with ∂D 2 = S1 as indicated in the picture.

x
f (x) • r (x)

This yields a retraction r : D 2 → S1 and we get a contradiction. 

Example 2.46. We show that the system of equations


1
1+ sin(x)y − 2x = 0,
2
1 x2
cos(x)y + − 2y = 0, (2.1)
2 2
has a solution. We rewrite this as a fixed point equation and apply the Brouwer fixed point
theorem. To do this we put
!
1 sin(x)y cos(x)y x 2
f (x, y) := + , +
2 4 4 4

Fixed points of f (x, y) are then the same as solutions to the above system of equations (2.1). To
apply Brouwer’s fixed point theorem we have to show that f (D 2 ) ⊂ D 2 . Let x, y ∈ D 2 . Then
!2 !2
2 1 sin(x)y cos(x)y x 2
k f (x, y)k = + + +
2 4 4 4
1 sin(x)y sin(x) y 2 2 cos(x) 2 y2 cos(x)yx 2 x 4
= + + + + +
4 4 16 16 8 16
1 sin(x)y y2 cos(x)yx 2 x 4
= + + + +
4 4 16 8 16
1 |y| y2 |y|x 2 x 4
≤ + + + +
4 4 16 8 16
1 1 1 1 1
≤ + + + +
4 4 16 8 16

38
2.3 The fundamental group of the circle

3
=
4
≤ 1.

Example 2.47. Consider f n ∈ C (S1, S1 ) with f n (z) = z n . We check that the diagram
deg
π1 (S1 ; 1) 
/Z

( f n )# n·
 deg 
π1 (S1 ; 1) 
/Z

commutes. Let [ω] ∈ π1 (S1 ; 1). We compute


 
deg ( f n )# ([ω]) = deg [ f n ◦ ω] = deg( f n ◦ ω)

= deg( f n ◦ f ω ) = deg( f n ) deg( f ω ) = n deg [ω] .

Remark 2.48. From Exercise 2.5 we know that for X1, X2 and x 1 ∈ X1, x 2 ∈ X2 we have

π1 (X1 × X2 ; (x 1, x 2 ))  π1 (X1 ; x 1 ) × π1 (X2 ; x 2 ) .

For the two-dimensional torus T 2 = S1 × S1 we get π1 (T 2 ; x)  Z × Z = Z2 . More generally, we


S1 × {z
get inductively for the n-torus T n = | · · · × S}1 that π1 (T n ; x)  Zn . In particular, T n ; T m
n
for n , m.

Definition 2.49. Let X be a topological space.

1.) We call X connected iff X and ∅ are the only subsets of X which are both open and closed.

2.) We call X path-connected iff for all x 1, x 2 ∈ X there exists a path ω ∈ Ω(X ; x 1, x 2 ).

3.) Let X be path-connected. Then X is called simply-connected (or 1-connected) iff


π1 (X ; x 0 ) = {1} for some (and hence all) x 0 ∈ X .

Example 2.50. The interval [0, 1] is connected. To see this let I ⊂ [0, 1] be open and closed
and assume that I is neither empty nor all of [0, 1]. Then there exists t 0 ∈ I and t 1 ∈ [0, 1] \ I.
W.l.o.g. let t 0 < t 1 , the other case being analogous. We put T := sup(I ∩ [0, t 1 )). Then
0 ≤ t 0 ≤ T ≤ t 1 ≤ 1. Since I is closed T ∈ I.
If T = 1 then T = t 1 which contradicts t 1 < I. If T < 1 then there exists ε > 0 such that
[T, T + ε) ⊂ I because I is open. This contradicts the maximality of T.

Remark 2.51. If X is 1-connected then it is path-connected by definition but the converse is not
true. For example, S1 is path-connected but not 1-connected.

39
2 Homotopy Theory

Remark 2.52. If X is path-connected then X is connected.

Proof. Let X be path-connected and let U ⊂ X be open and closed, U , ∅. We show U = X .


Since U is non-empty we can find x 1 ∈ U. Let x 2 ∈ X be any point.
Since X is path-connected there is a path ω ∈ Ω(X ; x 1, x 2 ). Then I := ω−1 (U ) is an open and
closed subset of [0, 1]. Since ω(0) = x 1 ∈ U we have 0 ∈ I and hence I is non-empty. Thus
I = [0, 1] because [0, 1] is connected. We conclude 1 ∈ I and hence x 2 = ω(1) ∈ U. This shows
that U contains all points of X . 

Again, the converse implication does not hold in general. Consider for example the space

X := {(t, sin(1/t)) | t > 0} ∪ {(0, s) | −1 ≤ s ≤ 1}

Then X is connected but not path-connected.

Remark 2.53. Let X be path-connected. Then the following are equivalent (see Exercise 2.2):
(i) X is simply connected;

(ii) Every ω ∈ C (S1, X ) is homotopic to a constant map;

(iii) Every ω ∈ C (S1, X ) has a continuous extension to a map D 2 → X .

2.4 The Seifert-van Kampen theorem


The Seifert-van Kampen theorem will allow us to compute the fundamental group of spaces
which are built out of simpler spaces whose fundamental groups we already know. We start with
an excursion to group theory.

Definition 2.54. Let G be a group. A subgroup H ⊂ G is called normal iff

g · H = H · g for all g ∈ G .

40
2.4 The Seifert-van Kampen theorem

The condition on a subgroup of being normal can be reformulated in various ways. It is equivalent
to any of the following:

1.) g · H · g−1 = H for all g ∈ G;

2.) g · H · g−1 ⊂ H for all g ∈ G;

3.) g · h · g−1 ∈ H for all g ∈ G and h ∈ H.

Example 2.55. Consider the cartesian product of two groups G = G1 ×G2 = {(g1, g2 ) | g j ∈ G j }
with componentwise multiplication. Then {(g1, 1) | g1 ∈ G1 }  G1 is a normal subgroup of G
because
 
(g1, g2 ) · (g̃1, 1) · (g1, g2 ) −1 = g1 g̃1 g1−1, g2 1g2−1 = g1 g̃1 g1−1, 1 .

Example 2.56. Let ϕ : G → K be a group homomorphism. Then H = ker(ϕ) is a normal


subgroup because for h ∈ ker(ϕ) and g ∈ G we have

ϕ(g · h · g−1 ) = ϕ(g) · ϕ(h) ·ϕ(g) −1 = 1 ,


|{z}
1

hence g · h · g−1 ∈ ker(ϕ).

Remark 2.57. If H ⊂ G is a normal subgroup then G/H is again a group via

(g · H ) · (g̃ · H ) = (g g̃) · H.

Normality of H ensures that this multiplication is well defined. The group H is then the kernel of
G → G/H with g 7→ g · H. Thus the normal subgroups are exactly those which arise as kernels
of group homomorphisms.

Now let S ⊂ G be any subset. Then


\
N (S) := H
H ⊂G normal subgroup,
H ⊃S

is the smallest normal subgroup containing S. We call N (S) the normal subgroup generated by
S.

Example 2.58. N (∅) = {1}.

41
2 Homotopy Theory

Definition 2.59
Let G1 and G2 be groups. A group G is called free product G1 ❆
of G1 and G2 iff there exist homomorphisms i j : G j → G such i1 ⑥⑥⑥ ❆❆ ϕ1
❆❆

that for all groups H and for all homomorphisms ϕ j : G j → H ~ ⑥⑥ ⑥
ϕ 1 ⋆ϕ 2
❆❆
there exists a unique homomorphism G `❆ /H
❆❆ ⑥>
❆❆ ⑥⑥⑥
❆ ⑥⑥ϕ
ϕ1 ⋆ ϕ2 : G → H i 2 ❆❆ ⑥⑥ 2
G2
such that the diagram commutes.

Remark 2.60. Here we have characterized free products by


their universal property. This universal property implies for G1
i1
⑦ ❇❇❇❇❇❇❇
example that the maps i j : G j → G are injective. ⑦⑦⑦ ❇❇❇id❇
⑦⑦ ❇❇❇❇
⑦ ❇
Namely, choose H = G1 , ϕ1 = idG1 and ϕ2 (g2 ) = 1 for all ~ ⑦ id ⋆ϕ 2
/ G1
G `❅
g2 ∈ G2 . The diagram now tells us that the map i1 must be ❅❅ ⑤⑤>
❅❅ ⑤⑤
injective because the identity is injective. Similarly, we see that ❅
i 2 ❅❅ ⑤⑤ ϕ
i 2 is injective. ⑤⑤ 2
G2

Remark 2.61. The free product of G1 and G2 is unique up to isomorphism.

G1
Namely, let G ′ be another free product of G1 and G2 with i1 ⑥ ❇❇❇ i ′
⑥⑥ ❇❇1
i ′j : G j → G ′ the corresponding homomorphisms. By the ⑥ ⑥ ❇❇
~⑥⑥ ′ ′
i 1 ⋆i 2
universal property of G with H = G ′ and ϕ j = i ′j we get the G `❆ / G′
❆❆ ⑤>
following commutative diagram: ❆❆ ⑤⑤

❆ ⑤⑤i ′
i 2 ❆❆ ⑤
⑤ 2
G2

G1
i ′1 ⑤ ❆❆❆ i
⑤⑤⑤ ❆❆1
⑤ ⑤ ❆❆
~
⑤ ⋆i
Interchanging the roles of G and G ′ we get another commutative G ′ `❇
i 1 2 /G
diagram: ❇❇ ⑥ ⑥>
❇❇ ⑥⑥
❇ ⑥⑥ i 2
i ′2 ❇❇ ⑥⑥
G2

42
2.4 The Seifert-van Kampen theorem

G1 ❆
i1 ⑥⑥⑥ ′ ❆❆❆i 1
i1
⑥⑥ ❆❆
Combining both diagrams we obtain ~⑥⑥i ′1 ⋆i ′2  i 1 ⋆i 2 ❆
G `❆ / G′ /G
❆❆ O >
❆❆ ′ ⑥⑥⑥
❆i ⑥⑥
i 2 ❆❆2 ⑥⑥⑥ i 2
G2

G1 ❆
i1 ⑥ ❆❆ i
On the other hand, this diagram commute as well. ⑥⑥⑥ ❆❆1
⑥ ❆❆
~⑥
⑥ id
G `❆ /G
❆❆ ⑥ ⑥>
❆❆ ⑥⑥
❆ ⑥⑥ i 2
i 2 ❆❆ ⑥⑥
G1
By the uniqueness of the induced homomorphisms we have

(i1 ⋆ i 2 ) ◦ i1′ ⋆ i 2′ = idG

and similarly

i1′ ⋆ i 2′ ◦ (i1 ⋆ i 2 ) = idG′ .
Hence the map i 1 ⋆ i 2 : G → G ′ is a group isomorphism with inverse i 1′ ⋆ i2′ .
Next we show the existence of the free product of two groups by a direct construction. For this
purpose let G1 and G2 be groups. For formal reasons we assume without loss of generality that
G1 ∩ G2 = ∅.1 We define
 
G1 ⋆ G2 := (x 1, . . . x n ) | n ∈ N0, x j ∈ G1 \ {1G1 )} ∪ (G2 \ {1G2 } such that

if x i ∈ G1 then x i+1 ∈ G2 or conversely .

The group multiplication in G1 ⋆ G2 is then inductively defined as

(x 1, . . . , x n ) · (x n+1, . . . , x n+m )




(x 1, . . . , x n−1, x n · x n+1, x n+2, . . . , x n+m ) if (x n , x n+1 ∈ G1 or x n, x n+1 ∈ G2 )







and x n · x n+1 , 1,
:= 


(x 1, . . . , x n−1 ) · (x n+2, . . . , x n+m ) if (x n , x n+1 ∈ G1 or x n, x n+1 ∈ G2 )






and x n · x n+1 = 1,

 (x , . . . , x , x , . . . , x
 1 n n+1 n+m ) otherwise.

This turns G1 ⋆G2 into a group with neutral element the empty sequence (). The inverse element
for (x 1, . . . x n ) is given by (x −1 −1
n , . . . , x 1 ) because of

−1  −1 
(x 1, . . . , x n ) · x −1
n , . . . , x1 = (x 1, . . . , x n−1 ) · x −1
n−1, . . . , x 1 = · · · = (x 1 )(x −1
1 ) = ().
1 Otherwise replace G2 by an isomorphic group which is disjoint to G1 .

43
2 Homotopy Theory

Now consider the map i j : G j → G1 ⋆ G2 given by



 (x),
i j (x) = 
x,1
 (),
 x=1
For ϕ j : G j → H homomorphisms we put
(ϕ1 ⋆ ϕ2 )(x 1, . . . , x n ) := ϕi 1 (x 1 ) · ϕi 2 (x 2 ) · . . . · ϕi n (x n )
where i j is chosen such that x j ∈ G i j .

Remark 2.62
1.) The subset i j (G j ) ⊂ G1 ⋆ G2 is a subgroup isomorphic to G j . The intersection i1 (G1 ) ∩
i2 (G2 ) = {1} is trivial. The union i 1 (G1 ) ∪ i2 (G2 ) generates i 1 (G1 ) ⋆ i 2 (G2 ) as a group.
2.) If G1 = {1} then G1 ⋆ G2 = i 2 (G2 )  G2 . Similarly, if G2 = {1} then G1 ⋆ G2  G1 .
3.) If G1 , {1} and G2 , {1} then we may choose x ∈ G1 \ {1} and y ∈ G2 \ {1}. Then
(x), (x, y), (x, y, x), (x, y, x, y), . . .
yields infinitely many pairwise different elements in G1 ⋆ G2 , hence |G1 ⋆ G2 | = ∞ (even if
|G j | < ∞). In addition, we have
(x) · (y) = (x, y) , (y, x) = (y) · (x) ,
hence the group G1 ⋆ G2 is not abelian, even if G1 and G2 are.

Remark 2.63. Usually one identifies G1 with i1 (G1 ) and G2 with i2 (G2 ) and writes
x 1 · x 2 · . . . · x n instead of (x 1, x 2, . . . , x n ).

Example 2.64. Consider G1 = Z/2Z = {1, −1} and G2 = Z/2Z = {1′, −1′ }. Elements of
Z/2Z ⋆ Z/2Z are for example
a = (−1) · (−1′ ) · (−1) · (−1′ )
b = (−1′ ) · (−1) · (−1′ ) · (−1) · (−1′ ) .
Now we calculate
a · b = (−1) · (−1′ ) · (−1) ·(−1′ ) · (−1′ ) ·(−1) · (−1′ ) · (−1) · (−1′ )
| {z }
1′

= (−1) · (−1 ) ·(−1) · (−1) ·(−1′ ) · (−1) · (−1′ )
| {z }
1
= (−1) ·(−1′ ) · (−1′ ) ·(−1) · (−1′ )
| {z }
1′
= (−1) · (−1) ·(−1′ )
| {z }
1
= (−1′ ) .

44
2.4 The Seifert-van Kampen theorem

Now we are ready to return to topology.

Proposition 2.65. Let X be a topological space and let U, V ⊂ X be open such that U ∪V = X .
Let x 0 ∈ U ∩ V . Furthermore, assume that U, V and U ∩ V are path-connected. Then X is
path-connected and the map

i # ⋆ j# : π1 (U; x 0 ) ⋆ π1 (V ; x 0 ) → π1 (X ; x 0 )

is onto where i : U → X and j : V → X are the corresponding inclusion maps.

Proof. First of all we note that the space X is path-connected because each point in X lies in U
or in V and can therefore be connected to x 0 by a path.
The statement of the proposition is equivalent to saying that
 
i# π1 (U; x 0 ) ∪ j# π1 (V ; x 0 )
generates π1 (X ; x 0 ) as a group. Now let [ω] ∈ π1 (X ; x 0 ) and subdivide the unit interval I = [0, 1]
by 0 = t 0 ≤ t 1 · · · < t n = 1 such that ω([t i , t i+1 ]) ⊂ U or ⊂ V .

X
ω0 ω2

U
x0
U ∩V η1 η2 η3 U ∩V
ω(t 1 ) ω(t 2 ) ω(t 3 )

ω1

By removing subdivision points if necessary we can assume that if ω([t i−1, t i ]) ⊂ U then
ω([t i , t i+1 ]) ⊂ V or conversely. Reparametrize ωi (t) = ω((1 − t)t i + tt i+1 ). Then ωi ∈
Ω(U or V ; ω(t i ), ω(t i+1 )) and in particular ω(t i ) ∈ U ∩ V . Since by assumption U ∩ V is
path-connected there exist η j ∈ Ω(U ∩ V ; x 0, ω(t j )). Then
ω0 ⋆ η 1−1, η 1 ⋆ ω1 ⋆ η 2−1, η 2 ⋆ ω2 ⋆ η 3−1, . . . , η n−1 ⋆ ω n−1

45
2 Homotopy Theory

are loops with base point x 0 contained entirely in U or V . We then calculate


  
ω0 ⋆ η 1−1 ⋆ η 1 ⋆ ω1 ⋆ η 2−1 ⋆ η 2 ⋆ ω2 ⋆ η 3−1 ⋆ · · · ⋆ (η n−1 ⋆ ω n−1 )
≃ {0,1} ω0 ⋆ ω1 ⋆ · · · ⋆ ω n−1 ≃ {0,1} ω
in X . Denoting the homotopy class of a loop ω in X , U, V , or U ∩ V by [ω] X , [ω]U , [ω]V , and
[ω]U∩V respectively, we have
     
[ω] X = ω0 ⋆ η 1−1 X · η 1 ⋆ ω1 ⋆ η 2−1 X · . . . · η n−1 ⋆ ω n−1 X
and it follows that
        
[ω] X = i# ω0 ⋆ η 1−1 U · j# η 1 ⋆ ω1 ⋆ η 2−1 V · i # η 2 ⋆ ω2 ⋆ η 3−1 U . . .
    f  
= (i# ⋆ j# ) ω0 ⋆ η 1−1 U · η 1 ⋆ ω1 ⋆ η 2−1 V · η 2 ⋆ ω2 ⋆ η 3−1 U . . .
| {z }
∈π 1 (U;x 0 )⋆π 1 (V ;x 0 )

which proves the assertion. 

Corollary 2.66. Let X is a topological space and let U, V ⊂ X be open such that U ∪ V = X
and U ∩ V , ∅. If U and V are 1-connected and U ∩ V is path-connected, then the space X is
1-connected.

Proof. Since
{1} = {1} ⋆ {1} = π1 (U; x 0 ) ⋆ π1 (V ; x 0 ) → π1 (X ; x 0 )
is onto we get that π1 (X ; x 0 ) = {1}. 

Example 2.67. Consider X = S n and put U = S n \ {e1 }. The stereographic projection yields
a homeomorphism U → Rn . Hence U is 1-connected. Similarly, V = S n \ {−e1 } is also
1-connected. Now
U ∩ V = S n \ {e1, −e1 } ≈ Rn \ {0}.
For n ≥ 2 the space U ∩ V is path-connected. Corollary 2.66 shows that S n is simply connected
for n ≥ 2.
Recall that we know already that this is not true for n = 1 because π1 (S1 ; 1)  Z.

To determine π1 (X ; x 0 ) more precisely we compute the kernel of the homomorphism


i# ⋆ j# : π1 (U; x 0 ) ⋆ π1 (V ; x 0 ) → π1 (X ; x 0 ).
i /X
UO O
Consider the inclusion maps i ′
: U ∩ V → U and : U ∩ V → V . j′
i′ j
Clearly the diagram on the right commutes. This implies
U ∩V /V
j′
j# ◦ j#′ = i # ◦ i#′ .

46
2.4 The Seifert-van Kampen theorem

For α ∈ π1 (U ∩ V ; x 0 ) we calculate
    −1   
1 = i # i#′ (α) · j# j#′ (α) = i # i #′ (α) · j# j#′ (α) −1 = (i # ⋆ j# ) i #′ (α) · j#′ (α) −1 .
| {z }
∈π 1 (U;x 0 )⋆π 1 (V ;x 0 )

Hence i#′ (α) · j#′ (α) −1 ∈ ker(i # ⋆ j# ) and it follows that


 
N i#′ (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 ) ⊂ ker(i# ⋆ j# ) .

Theorem 2.68 (Seifert-van Kampen). Let X be a topological space and let U, V ⊂ X be


open subsets such that U ∪ V = X and x 0 ∈ U ∩ V . Let U, V and U ∩ V be path connected.
Then the map i# ⋆ j# induces an isomorphism
π1 (U; x 0 ) ⋆ π1 (V ; x 0 )
′   π1 (X ; x 0 ).
N i# (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 )

Proof. It remains to show that


 
ker(i# ⋆ j# ) ⊂ N i#′ (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 ) .

Let ω1, ω3, · · · ∈ Ω(U; x 0 ) and ω2, ω4, · · · ∈ Ω(V ; x 0 ) be such that

1 = i # ⋆ j# [ω1 ]U · [ω2 ]V · [ω3 ]U · . . .
= i # ([ω1 ]U ) · j# ([ω2 ]V ) · i # ([ω3 ]U ) · . . .
= [ω1 ] X · [ω2 ] X · [ω3 ] X · . . .
= [ω1 ⋆ ω2 ⋆ ω3 ⋆ . . . ] X
s

1 −
Then there exists a homotopy H : [0, 1] ×
[0, 1] → X such that

H (t, 0) = (ω1 ⋆ ω2 ⋆ ω3 ⋆ . . . )(t),


H (t, 1) = x 0,
H (0, s) = H (1, s) = x 0 .

In the picture the red area gets mapped to 0 − | | |


x0 . ω1 ω2 ωN
| |
t
0 1

47
2 Homotopy Theory

Now subdivide [0, 1] × [0, 1] further such b

that H maps each closed subsquare entirely v


to U or to V .

| | |
ω1 ω2 ωN

u
>

Considering the edges of the subsquares we l r


get homotopies ∧ ∧

>
d

and hence the relation


d ⋆ r ≃ {0,1} l ⋆ u (2.2)
resulting in
(H ◦ d) ⋆ (H ◦ r) ≃ {0,1} (H ◦ l) ⋆ (H ◦ u) in U or in V .
For each vertex v in this subdivision of [0, 1] × [0, 1] choose η v ∈ Ω(X ; x 0, H (v)) in such a way
that η v ∈ Ω(W ; x 0, H (v)) if H (v) ∈ W where W = U, V or U ∩ V . This is possible because
x 0 ∈ W and W is path-connected by assumption. If H (v) = x 0 choose η v = ε x0 . For each edge
with endpoints v0 and v1 we obtain a loop in U or in V by

η v0 ⋆ (H ◦ c) ⋆ η −1
v 1 ∈ Ω(U or V ; x 0 ) .

Now look at one row of the subdivision

48
2.4 The Seifert-van Kampen theorem

u1 u2 uN
> > >
ε x0 = l 1 ∧ ∧ ∧ ∧ ∧ r N = ε x0
> > >
d1 d2 dN

We find that
Di := η d i (0) ⋆ (H ◦ d i ) ⋆ η −1
d i (1) ∈ Ω(U or V ; x 0 ).
Similarly we define L i , Ri , Ui ∈ Ω(U or V ; x 0 ) and we conclude that by (2.2)
[Di ]Wi · [Ri ]Wi = [L i ]Wi · [Ui ]Wi ∈ π1 (Wi ; x 0 ) (2.3)
where Wi = Uor V . We now compute in π1 (U; x 0 ) ⋆ π1 (V ; x 0 ) modulo N ({i#′ (α) · j#′ (α) −1 | α ∈
π1 (U ∩ V ; x 0 )}):
[D1 ]W1 · [D2 ]W2 · . . . · [D N ]W N = [D1 ]W1 · [D2 ]W2 · . . . · [D N ]W N · [RN ]W N
= [D1 ]W1 · . . . · [D N −1 ]W N −1 · [L N ]W N · [UN ]W N
= [D1 ]W1 · . . . · [D N −1 ]W N −1 · [RN −1 ]W N · [UN ]W N .
If now W N −1 = W N then we can apply (2.3) once more and get
[D1 ]W1 · [D2 ]W2 · . . . · [D N ]W N = [D1 ]W1 · . . . · [D N −2 ]W N −2 · [L N −1 ]W N −1 · [UN −1 ]W N −1 · [UN ]W N
(2.4)
In case W N −1 , W N , say W N −1 = U and W N = V , then L N = RN −1 ∈ Ω(U ∩ V ; x 0 ). Hence
mod N (... ) ′
[RN −1 ]V = j#′ ([RN −1 ]U∩V ) = i# ([RN −1 ]U∩V ) = [RN −1 ]U .
In
π1 (U; x 0 ) ⋆ π1 (V ; x 0 )
G :=
N ({i #′ (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 )})
the computation (2.4) is still possible. By induction on all squares in the row from right to the
left we find that
=1
z }| {
[D1 ]W1 · . . . · [D N ]W N = [L 1 ]W1 ·[D1 ]W1 · . . . · [D N ]W N = [U1 ]W1 · . . . · [UN ]W N .

49
2 Homotopy Theory

A second induction on all rows from bottom to top yields in G


mod N (... )
[ω1 ]U · [ω2 ]V · [ω3 ]U · . . . = [ε x0 ]W1 · [ε x0 ]W2 · [ε x0 ]W3 · . . . = 1 .

We have shown that in π1 (U; x 0 ) ⋆ π1 (V ; x 0 )

[ω1 ]U · [ω2 ]V · [ω3 ]U · . . . ∈ N ({i #′ (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 )})

and hence  
ker(i # ⋆ j# ) ⊂ N i #′ (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 ) .

Corollary 2.69. Let X be a topological space and U, V ⊂ X such that U ∪ V = X and


x 0 ∈ U ∩ V . Assume that U, V are path connected and that U ∩ V is 1-connected. Then

π1 (X ; x 0 )  π1 (U; x 0 ) ⋆ π1 (V ; x 0 )

where the isomorphism is induced by the inclusion maps.

Proof. By assumption π1 (U ∩ V, x 0 ) = {1} and hence

N ({i#′ (α) · j#′ (α) −1 | α ∈ π1 (U ∩ V ; x 0 )}) = {1} .

The assertion then follows from Theorem 2.68. 

Example 2.70. Consider the figure 8 space and the two subsets U and V as indicated in the
picture:

X
V U
b b

x0 x0

It is easy to see that U ≃ S1 and also V ≃ S1 . Moreover, the intersection

U ∩ V ≃ point

is 1-connected. Hence we have

π1 (X ; x 0 )  π1 (S1 ; x 0 ) ⋆ π1 (S1 ; x 0 )  Z ⋆ Z .

50
2.4 The Seifert-van Kampen theorem

Example 2.71. Let X be a connected n-dimensional manifold with n ≥ 3.

M
b
V
p

Let p ∈ X and U := X \ {p}. Now let V be an open neighborhood of p homeomorphic to Rn ,


which is contractible and hence 1-connected. Then the space

U ∩ V ≈ Rn \ {0} ≃ S n−1

is 1-connected. By Corollary 2.69 we then deduce

π1 (X ; x 0 )  π1 (U; x 0 ) ⋆ π1 (V ; x 0 ) = π1 (X \ {p}; x 0 ) .

Removing a point from a manifold of dimension at least 3 does not change its fundamental group.

Example 2.72. Let M and N be two connected manifold of dimension n ≥ 3.

M N

Let X = M#N.

U ∩V

| {z }
=U | {z }
=V

51
2 Homotopy Theory

Now choose U and V as in the picture. Then we have that U ≈ M \ {p} and V ≈ N \ {q}. Since
U ∩ V ≈ S n−1 × (0, 1) ≃ S n−1
is 1-connected we find by Corollary 2.69 once again that
π1 (M#N )  π1 (U ) ⋆ π1 (V )  π1 (M ) ⋆ π1 (N ) .
For example, if M = N = T 3 then π1 (T 3 #T 3 ) = (Z3 ) ⋆ (Z3 ).

Remark 2.73. We now see easily that the torus T n with n ≥ 3 cannot be homotopy equivalent
to the connected sum T n ≃ M#N of two non-1-connected manifolds M and N.2 If it were
possible then π1 (T n )  π1 (M ) ⋆ π1 (N ) would not be abelian but we know that π1 (T n )  Zn , a
contradiction.

2.5 The fundamental group of surfaces


Our aim is to prove that orientable compact connected surfaces of different genus (as depicted
by the pastries in Example 1.5) are not homotopy equivalent and therefore not homeomorphic.

2
Definition 2.74. We call Fg := T · · · #T}2 a surface of genus g ≥ 1.
| # {z
g−times

Fg = ···

| {z }
g−times

Remark 2.75. We also put F0 := S2 .

We now want to compute π1 (Fg ) and show that the fundamental groups for surfaces of different
genus are not isomorphic. This then shows in particular that they are not homotopy equivalent.

Proposition 2.76. For any g ≥ 1


2
· · · #T}2 ≈ D 2 / ∼
| # {z
Fg = T
g−times

2 If one allows simply connected summands then it is of course possible, T n ≈ T n #S n .

52
2.5 The fundamental group of surfaces

where x ∼ y iff x = y or x, y ∈ ∂D 2 and are identified according to the following scheme:

b−1
g a1
a−1 b1
g
a1−1

b−1
1
··· ≈ D2
a2

b2
a2−1

This identification is to be understood as follows: Each line segment labelled by ai (or b j


respectively) is identified with with ai−1 (or b−1
j ) with respect to the direction indicated by the
arrow. Note that there are 2g labels ai , bi , i = 1, .., g.

Proof. We do an induction on g.
Induction basis for g = 1: We see that the labelled disk D 2 / ∼ is homeomorphic to the labelled
W 2 / ∼ which is homeomorphic to the two-dimensional torus, i.e. to F1 .
a1

b−1
1 a1 a1

≈ b−1
1 b1 ≈ b−1
1 b1 ≈

a1−1 b1

a1−1

Inductive step, g − 1 ⇒ g: We perform a cut along the line c.

53
2 Homotopy Theory

b−1
g a1 a1
b1 b1
a−1
g b−1
g
a1−1 a−1 a1−1
g
c
b−1
1 b−1
1
≈ c
a2
a2
b2
b2
a2−1
a2−1

The endpoints of c in the two pieces are identified. This yields a homeomorphism where the
interior of the now closed loop c is cut out of the two remaining disks.

a1
b1
b−1
g a1−1
a−1
g b−1 a2
g
a−1 b2
c b−1 g
1 a2−1 b−1
1 a1
c
c b−1 c
≈ 2
a2

b2 a1−1 b1

a2−1

By induction hypothesis the left-hand disc is homeomorphic to a surface of genus g − 1 with


a disc removed that is bounded by c. The right-hand disc becomes a torus, also with a disc
removed that is bounded by c.
Gluing these two spaces together along c, we obtain a space which is homeomorphic to a surface
of genus g:

c c ≈

| {z } | {z }
(g−1)−times g−times


54
2.5 The fundamental group of surfaces

Remark 2.77. We note that F0 ≈ D 2 /∂D 2 .

Remark 2.78. Let G be a group. Assume G is generated by g1, ..., gn ∈ G as a group. We have
group homomorphisms ϕi : Z → G, ϕi (k) = gik . Repeated application of the universal property
of free products of groups yields the group homomorphism

(...(ϕ1 ⋆ ϕ2 ) ⋆ ϕ3 ) ⋆ · · · ⋆ ϕ n =: ϕ1 ⋆ · · · ⋆ ϕ n : (...(Z ⋆ Z) ⋆ Z) ⋆ · · · ⋆ Z =: Z ⋆ · · · ⋆ Z → G

with
(ϕ1 ⋆ · · · ⋆ ϕ n )((k1, i 1 ), ..., (kr , i r )) = gik11 · · · gikrr ,
where i j ∈ {1, ..., n} is an index used to make the Z-factors formally disjoint.
The fact that g1, ..., gn generate G is equivalent to the fact that ϕ1 ⋆ · · · ⋆ ϕ n : Z ⋆ · · · ⋆ Z → G
is onto. It follows that
Z⋆···⋆Z
G
ker(ϕ1 ⋆ · · · ⋆ ϕ n )
If the normal subgroup ker(ϕ1 ⋆ · · · ⋆ ϕ n ) is also finitely generated as a group, with generators
x 1, ..., x m , then we call G finitely presentable and

hg1, ..., gn | x 1, ..., x m i

a presentation of G.

Example 2.79. For the cartesian product Z2 of two copies of Z we have the presentation

Z2  ha, b | aba−1 b−1 i.

The cyclic group Z/2Z of order 2 is presentable as

Z/2Z  ha | a2 i.

Remark 2.80. A word of caution: isomorphic groups may have several, quite different presen-
tations. For example

hx, y | xyxy−1 x −1 y−1 i  hx, y, w, z | xyxy−1 x −1 y−1, xyxw −1, zy−1 x −1 i

because the generators xyxw −1 and zy−1 x −1 on the right-hand side can be used to eleminate w
and z. It is therefore often not obvious whether two presentations give rise to isomorphic groups.

Theorem 2.81. For any g ∈ N we have

π1 (Fg )  ha1, b1, ..., ag , bg | a1 b1 a1−1 b−1 −1 −1


1 · · · a g bg a g bg i.

55
2 Homotopy Theory

Proof. Recall that Fg ≈ D 2 / ∼ as in Proposition 2.76. To apply the Seifert-van Kampen


theorem 2.68 we put U := D̊ 2 and V := (D 2/ ∼) \ D 2 ( 21 ). We have Fg = U ∪ V .

b−1
g a1
b1
a−1
g
a1−1

b−1
1
V
a2

b2

a2−1

The subset U is contractible and therefore π1 (U ) = {1}. The subset V is homotopy equivalent
to the boundary ∂D subject to the identifications of the equivalence relation, V ≃ ∂D/ ∼. The
identifications induced by ∼ generate a bouquet of 2g circles, one for each relation ai and bi .
The bouquet is denoted by S1 ∨ · · · ∨ S1 , where the wedge sum “∨” of two topological spaces X
and Y is defined to be the disjoint union of X and Y with identification of two base points x 0 ∈ X ,
y0 ∈ Y such that X ∨ Y := X ∪ Y /{x 0 ∼ y0 }. For the bouquet of circles all S1 ’s are joined at the
same base point. Graphically we can depict the bouquet of circles as follows

S1 ∨ · · · ∨ S1 =

So we have
V ≃ ∂D 2 / ∼ ≈ |
S1 ∨ {z
· · · ∨ S}1 .
2g times

56
2.5 The fundamental group of surfaces

The fundamental group of V follows immediately from Example 2.70 by induction:


π1 (V )  | ⋆···⋆}
Z ⋆ Z {z Z
2g times

with generators again denoted by a1, b1, a2, b2, ..., ag , bg . For the intersection of U and V we
have U ∩ V = D̊ 2 \ D̄ 2 ( 21 ) ≃ S1 and thus by Theorem 2.42 we have π1 (U ∩ V )  Z. A generator
of π1 (U ∩ V ) is given by a loop c of degree 1.

b−1
g a1
b1
a−1
g
a1−1

b−1
1
c
a2

b2

a2−1

The inclusion map i ′ : U ∩ V → U induces the trivial homomorphism i #′ because π1 (U ) is


trivial. For the inclusion map j ′ : U ∩ V → V we note that the induced isomorphism maps c
onto a1 b1 a1−1 b−1 −1 −1
1 · · · a g bg a g bg .
By the Seifert-van Kampen theorem 2.68 we find
π1 (U ) ⋆ π1 (V )
π1 (Fg ) 
N ( j#′ (α)
· i#′ (α) −1 | α ∈ π1 (U ∩ V ))
π1 (V )
= ′
N ( j# (α) | α ∈ π1 (U ∩ V ))
Z⋆···⋆Z
=
N ( j#′ (c))
= ha1, b1, ..., ag , bg | a1 b1 a1−1 b−1 −1 −1
1 · · · a g bg a g bg i. 

Example 2.82. For the two-dimensional torus T 2 we find with Example 2.79
π1 (T 2 ) = π1 (F1 ) = ha, b | aba−1 b−1 i  Z2,
in agreement with Remark 2.48.

57
2 Homotopy Theory

Corollary 2.83. For g, g ′ ∈ N, g , g ′ we have Fg ; Fg ′ and hence Fg 0 Fg ′ .

Proof. The statement follows once we see that π1 (Fg )  π1 (Fg ′ ). Attention here: as noted in
2.80 different presentations can yield isomorphic groups.
For any group G let [G, G] be the normal subgroup generated by all commutators aba−1 b−1 ,
a, b ∈ G. The abelian factor group G/[G, G] is called the abelianization of G. We now calculate
the abelianization of π1 (Fg ).
π1 (Fg )
= ha1, b1, ..., ag , bg | a1 b1 a1−1 b−1 −1 −1 −1 −1
1 · · · a g bg a g bg , a1 b1 a1 b1 , ...
[π1 (Fg ), π1 (Fg )]
..., a1 a2 a1−1 a2−1, ...i
= ha1, b1, ..., ag , bg | a1 b1 a1−1 b−1 −1 −1
1 , ..., a1 a2 a1 a2 , ...i
 Z2g ,
where the second equality follows because the simple commutators ai bi ai−1 b−1
i , i = 1, ..., g
−1 −1 −1 −1
generate the relation a1 b1 a1 b1 · · · ag bg ag bg .
Hence if Fg ≃ Fg ′ then π1 (Fg )  π1 (Fg ′ ) and thus
π1 (Fg )/[π1 (Fg ), π1 (Fg )]  π1 (Fg′ )/[π1 (Fg′ ), π1 (Fg′ )].

Thus Z2g  Z2g and therefore g = g ′. 

Remark 2.84. This proves the uniqueness part of the classification for surfaces, see Example 1.5.

2.6 Higher homotopy groups


We now generalize the definition of π1 (X ; x 0 ).

Definition 2.85. Let W n = [0, 1] × . . . × [0, 1] be the n-cube. Let X be a topological space
| {z }
n times
and x 0 ∈ X . Then

πn (X ; x 0 ) := {[σ]∂W n | σ ∈ C (W n, X ), σ(∂W n ) = {x 0 }}

is called the n-th homotopy group of X with base point x 0 . Here [σ]∂W n denotes the homotopy
class of σ relative to ∂W n .

The group structure on πn (X ; x 0 ) is obtained as follows: For σ, τ ∈ C (W n, X ) with σ(∂W n ) =


τ(∂W n ) = {x 0 } define σ ⋆ τ by

 σ(2t 1, t 2, . . . , t n ),
(σ ⋆ τ)(t 1, . . . , t n ) = 
0 ≤ t 1 ≤ 1/2,
 τ(2t 1 − 1, t 2, . . . , t n ), 1/2 ≤ t 1 ≤ 1.

58
2.6 Higher homotopy groups

Then σ ⋆ τ ∈ C (W n, X ) with (σ ⋆ τ)(∂W n ) = {x 0 }.


Wn Wn Wn

σ σ⋆τ

X t 2, . . . , t n
b
t1
x0

Now put [σ]∂W n · [τ]∂W n := [σ ⋆ τ]∂W n . The proof that this yields a well-defined group
multiplication on πn (X ; x 0 ) for n ≥ 2 is literally the same as in the case for n = 1. The neutral
element is represented by the constant map ε nx0 : W n → X , ε nx0 (t 1, . . . , t n ) = x 0 , and [σ]−1
∂W n is
represented by σ(1 − t 1, t 2, . . . , t n ).
Unlike for the case n = 1 the higher homotopy groups are abelian:

Proposition 2.86. Let X be a topological space and x 0 ∈ X . Then for n ≥ 2 the group
πn (X ; x 0 ) is abelian.

Proof. For σ, τ ∈ C (W n, X ) with σ(∂W n ) = τ(∂W n ) = {x 0 } we need to show that σ ⋆ τ and


τ ⋆ σ are homotopic relative to ∂W n . The homotopy is obtained by precomposing with the
homotopy of the n-cube indicated in the following picture (which illustrates the case n = 2):

59
2 Homotopy Theory

σ τ σ σ σ
deform deform deform
−→ −→ −→
τ τ τ

τ σ
deform
−→

Remark 2.87. This proof also shows that replacing t 1 by any other variable in the definition of
σ ⋆ τ gives the same group multiplication on πn (X ; x 0 ).

For f ∈ C (X, Y ) with f (x 0 ) = y0 we get a group homomorphism

f # : πn (X ; x 0 ) → πn (Y ; y0 )

defined by [σ]∂W n 7→ [ f ◦ σ]∂W n .

Remark 2.88. Lemma 2.18, 2.19, Corollary 2.20, Propositions 2.21, 2.22, Theorem 2.23 and
Corollary 2.24 also hold for πn (X ; x 0 ). In particular, if f : X → Y is a homotopy equivalence
then the map f # : πn (X ; x 0 ) → πn (Y ; f (x 0 )) is an isomorphism. For γ ∈ Ω(X ; x 0, x 1 ) there is
an isomorphism
Φγ : πn (X ; x 1 ) → πn (X ; x 0 )
given by [σ]∂W n 7→ [σ ′]∂W n where


 σ(2t),
σ ′ (t) = 
kt kmax ≤ 21 ,
 γ(2 − 2kt kmax ),

1
2 ≤ kt kmax ≤ 1.

γ
x0 x1 σ

60
2.6 Higher homotopy groups

Remark 2.89. By Exercise 1.7 we have the maps

π ϕ
Wn / W n /∂W n

/ Sn

and we see that σ ∈ C (W n, X ) with σ(∂W n ) = {x 0 } corresponds uniquely to f σ ∈ C (S n, X )


with f σ (s0 ) = x 0 , where s0 = ϕ(∂W n ) such that σ = f σ ◦ ϕ ◦ π. Thus there is a canonical
bijection
1:1
πn (X ; x 0 ) ←→ {[ f ] {s 0 } | f ∈ C (S n, X ), f (s0 ) = x 0 } .

Remark 2.90. The Seifert-van-Kampen theorem for πn works only under very restrictive as-
sumptions. For this reason the computation of πn for explicit examples can be very difficult. We
will be able to compute πn (S m ) for n ≤ m but many of the πn (S m ) for n > m are actually still
unknown.

Remark 2.91. The definition of πn (X ; x 0 ) also works for n = 0. A map σ ∈ C (W 0, X )


corresponds to the point σ(W 0 ) ∈ X . Two such maps are homotopic iff the corresponding points
can be joined by a path. Hence

π0 (X ; x 0 ) = {path components of X }.

But there is no (natural) group structure on π0 (X ; x 0 ). More precisely, π0 (X ; x 0 ) is a pointed set,


i.e., a set with a distinguished point, namely the path component containing x 0 . This corresponds
to the neutral element in πn (X ; x 0 ) for n ≥ 1.

Definition 2.92. Let W, E and B be topological spaces and let p ∈ C (E, B). We say that p has
the homotopy lifting property, abbreviated by HLP, for W iff for every f ∈ C (W, E) and every
h ∈ C (W × [0, 1], B) with h(w, 0) = p( f (w)) for all w ∈ W there exists H ∈ C (W × [0, 1], E)
such that
H (w, 0) = f (w) ∀w ∈ W and h = p◦H.
In other words, there exists an H ∈ C (W × [0, 1], E) such that the diagram

f
∈ W /
w
❴ ✉: E
H✉ ✉
p

  ✉ h

(w, 0) ∈ W × [0, 1] /B

commutes.

61
2 Homotopy Theory

Example 2.93. Consider the spaces E = {point}, B = R, W = b


e0
W 0 and the map given by p(e) = 0.
No h : W × [0, 1] → B except the constant path ε 0 can be lifted
because it leaves the image of p. Here the problem is the lack
of surjectivity of p.
| R
h
0

Example 2.94. Now consider E = [0, ∞) × {0} ∪ (−∞, 0] × {1} ⊂ R2 , B = R and let p be the
projection onto the first factor, p(t, s) = t.

f
b
E

W = {w0 }

| B
h

The map p is surjective but still does not have the HLP for W = W 0 . For example, choose
h(w0, t) = t, f (w0 ) = (0, 1).

Definition 2.95. A map p ∈ C (E, B) is called a Serre fibration or weak fibration iff it has the
HLP for all W n , n ≥ 0. The space E is called the total space and B is called the base space of
the fibration. For b0 ∈ B we call p−1 (b0 ) the fiber over b0 .

Example 2.96. For topological spaces F and B put E := B × F and p = pr1 , the projection on
the B-factor. Then p has the HLP for all W . In particular, p is a Serre fibration. Namely, let
f ∈ C (W, B × F) and h ∈ C (W × [0, 1], B) be given such that p( f (w)) = h(w, 0) for all w ∈ W .
Now write f (w) = ( β(w), ϕ(w)) with β ∈ C (W, B) and ϕ ∈ C (W, F). Hence
h(w, 0) = p( f (w)) = β(w).
Now put H (w, t) := (h(w, t), ϕ(w)). Then H ∈ C (W × [0, 1], B × F) and
p(H (w, t)) = pr1 (h(w, t), ϕ(w)) = h(w, t) ,
H (w, 0) = (h(w, 0), ϕ(w)) = ( β(w), ϕ(w)) = f (w) ,
as required. Hence p has the HLP for any W .

62
2.6 Higher homotopy groups

Definition 2.97. A map p ∈ C (E, B) is called fiber bundle with fiber F iff for each b ∈ B there
exists an open subset U ⊂ B with b ∈ U and a homeomorphism Φ : p−1 (U ) → U × F such
that the diagram
Φ
U×F o ≈ p−1 (U )
tt
pr 1 ttt
tt
 zttt p | p −1 (U )
U
commutes.

Lemma 2.98. Every fiber bundle is a Serre fibration.

Proof. Let f ∈ C (W n, E) and h ∈ C (W n × [0, 1], B) such that h(w, 0) = p( f (w)) for all w ∈ W .
Now subdivide W n × [0, 1] such that h maps each subcube entirely into an open subset U as in
the definition of the fiber bundle.

W n × [0, 1]

[0, 1]

Q11
Wn

By Example 2.96 products have the HLP, hence we can extend the map f to a continuous map
H11 : (W n × {0}) ∪ Q11 → E such that p ◦ H11 = h.

63
2 Homotopy Theory

W n × [0, 1]

[0, 1]

Q11 Q12
Wn
Next we want to extend the lift to Q12 . Now there seems to be a problem because the required
lift H12 need not only coincide with f along the edge Q12 ∩ (W n × {0}) but also with H11 along
Q11 ∩ Q12 .
But there are homeomorphisms of a cube onto itself mapping two edges onto one as indicated in
the picture.

Thus we can apply the HLP and extend the lift to (W n × {0}) ∪ Q11 ∪ Q12 . Iteration of this
procedure proves the assertion. 

Example 2.99. Let G be a Lie group, e.g. a closed subgroup of GL(n; K ) with K = R or K = C.
Let H ⊂ G be a closed subgroup. We equip the space B = G/H = {g · H | g ∈ G} with the
quotient topology. Such a space is called a homogeneous space. Then G → G/H with g 7→ g · H
is a fiber bundle with fiber H. The proof of this fact requires some technical work, see [8, p. 120
ff].

64
2.6 Higher homotopy groups

Example 2.100. Let G = SO(n + 1) and

B 0
( ! )

0 1
H= B ∈ SO(n) .

Then H is a closed subgroup of G isomorphic to SO(n). We now show that G/H ≈ S n . Consider
the map f : G → S n with A 7→ A · en+1 where en+1 the (n + 1)-st unit vector of the canonical
basis. The map f is continuous and surjective. We observe

f (A) = f ( Ã) ⇐⇒ A · en+1 = Ã · en+1


⇐⇒ Ã−1 · A · en+1 = en+1
!
−1 ⋆ 0
⇐⇒ Ã · A =
⋆ 1
⇐⇒ Ã−1 · A ∈ H
⇐⇒ π(A) = π( Ã)

where the map π : G → G/H is the canonical projection. We conclude that the map f descends
to a bijective map f¯ : G/H → S n . By the universal property of the quotient topology the map
f¯ : G/H → S n is continuous. Since G is compact the space G/H is also compact. Moreover,
the sphere S n is a Hausdorff space, hence the map f¯ is a homeomorphism. Thus we obtain a
fiber bundle SO(n + 1) → S n with fiber SO(n).

Let p : E → B be any Serre fibration and fix e0 ∈ E. Put b0 := p(e0 ) ∈ B and let F := p−1 (b0 ).
Then e0 ∈ F. Let ι : F → E be the inclusion map. We obtain the following two homomorphisms:

ι# : πn (F; e0 ) → πn (E; e0 ),
p# : πn (E; e0 ) → πn (B; b0 ).

Now we construct a map ∂ : πn (B; b0 ) → πn−1 (F; e0 ). We define

Box0 := {1} and Boxk := (W k × {1}) ∪ (∂W k × [0, 1]) for k ≥ 1.

Then we have

∂W n = (W n−1 × {0}) ∪ Boxn−1,


(W n−1 × {0}) ∩ Boxn−1 = ∂W n−1 × {0} .

Consider the homeomorphism η k : Boxk → W k obtained by the central projection from


( 12 , . . . , 12 , −1).

65
2 Homotopy Theory

b
x

b
η k (x)

b
( 12 , . . . , 21 , −1)

This homeomorphism maps the faces out of which Boxk is built onto the regions depicted
in the right figure. In order to define ∂ : πn (B; b0 ) → πn−1 (F; e0 ) let σ ∈ C (W n, B) with
σ(∂W n ) = {b0 }.

F E

tn
b e0

σ p

b
b0
t 1, . . . , t n−1

Since (t 1, . . . , t n−1 ) 7→ σ(t 1, . . . , t n−1, 0) = b0 is constant we can lift it to the constant map
(t 1, . . . , t n−1 ) 7→ e0 . Now the HLP of p for W n−1 yields a continuous map Σ : W n → E with

66
2.6 Higher homotopy groups

Σ(t 1, . . . , t n−1, 0) = e0 and p ◦ Σ = σ. From σ(∂W n ) = {b0 } we have Σ(∂W n ) ⊂ F. We put


σ̃ := Σ ◦ η −1 n−1 : W
n−1 → F. We want to define the map ∂([σ]
∂W n ) := [σ̃]∂W n−1 . We have to
check well-definedness of this map:

a) We have to show that [σ̃]∂W n−1 does not depend on the particular choice of the lift Σ.
Let Σ ′ ∈ C (W n, E) be another lift of σ with Σ ′ (t 1, . . . , t n−1, 0) = e0 . Then Σ−1 • Σ ′ is a lift of
σ −1 • σ. Here • denotes the concatenation with respect to the variable t n , Σ−1 (t 1, . . . , t n ) =
Σ(t 1, . . . , t n−1, 1−t n ) and similarly for σ −1 . Since σ −1 • σ ≃∂W n ε bn0 we can find a homotopy
h : W n+1 → B relative to ∂W n with

h(t 1, . . . , t n , 0) = (σ −1 • σ)(t 1, . . . , t n ) and h(t 1, . . . , t n, 1) = b0 .

Then h(Boxn ) = {b0 }. We apply the HLP of p for W n+1 to get a lift H ∈ C (W n+1, E) of h
with

H (t 1, . . . , t n, 0) = Σ−1 • Σ ′ (t 1, . . . , t n ).

From h(Boxn ) = {b0 } we have H (Boxn ) ⊂ F. Then we get a homotopy in F relative to


∂W n−1 from σ̃ = Σ ◦ η −1 ′ ′ −1
n−1 to σ̃ = Σ ◦ η n−1 as shown in the picture.

Σ′

Σ−1
t n+1 tn

t 1, . . . , t n−1

b) We also have to show that σ ≃∂W n σ ′ in B implies σ̃ ≃∂W n−1 σ̃ ′ in F.


Let h : W n × [0, 1] → B be a homotopy in B from σ to σ ′ relative to ∂W n .

67
2 Homotopy Theory

σ′

B
h
b
b0

t n+1 tn

t 1, . . . , t n−1
σ

The HLP for W n yields a lift H : W n+1 → E of h with H (t 1, . . . , t n−1, 0, t n+1 ) = e0 . We thus
obtain a homotopy
Ĥ (t 1, . . . , t n−1, s) = H (η −1
n−1 (t 1, . . . , t n−1 ), s)

in F from σ̃ to σ̃ ′ relative to ∂W n−1 and hence

[σ̃]∂W n−1 = [σ̃ ′]∂W n−1 .

We have shown that the map ∂ : πn (B; b0 ) → πn−1 (F, e0 ) is well defined.

Lemma 2.101. For n ≥ 2 the map ∂ : πn (B; b0 ) → πn−1 (F, e0 ) is a group homomorphism.

Proof. Let σ, τ ∈ C (W n, B) such that σ(∂W n ) = τ(∂W n ) = {b0 }. Choose a lift H ∈ C (W n, E)


of σ ⋆ τ with H (t 1, . . . , t n−1, 0) = e0 . Restriction yields lifts Σ of σ and T of τ up to stretching
in the t 1 -direction. Moreover, we have

(Σ ◦ η −1 −1 −1
n−1 ) ⋆ (T ◦ η n−1 ) ≃ ∂W n−1 H ◦ η n−1

68
2.6 Higher homotopy groups

The homotopy is given by shrinking the marked region in the t 1 -direction.

b b b

We conclude that

∂([σ]∂W n ) · ∂([τ]∂W n ) = [Σ ◦ η −1 −1
n−1 ]∂W n−1 · [T ◦ η n−1 ]∂W n−1
= [(Σ ◦ η −1 −1
n−1 ) ⋆ (T ◦ η n−1 )]∂W n−1
= [H ◦ η −1
n−1 ]∂W n−1
= ∂([σ ⋆ τ]∂W n )
= ∂([σ]∂W n · [τ]∂W n ). 

Hence ∂ : πn (B; b0 ) → πn−1 (F, e0 ) is a homomorphism if n ≥ 2. For n = 1 this statement


does not make sense because π0 (F, e0 ) is not a group. But ∂ still maps the neutral element of
π1 (B, b0 ) to the distinguished element of π0 (F, e0 ).

Theorem 2.102 (Long exact homotopy sequence of a Serre fibration). Let p : E → B be a


Serre fibration, e0 ∈ E, b0 = p(e0 ) ∈ B and F = p−1 (b0 ). Let ι : E → B be the inclusion
map. Then the following sequence is exact:

∂ ι# p# ∂ ι#
... / πn (F; eo ) / πn (E; e0 ) / πn (B; b0 ) / πn−1 (F; e0 ) / ...

∂ ι# p#
... / π0 (F; e0 ) / π0 (E; e0 ) / π0 (B; b0 )

69
2 Homotopy Theory

Remark 2.103. Exactness means that the image of the incoming map equals the kernel of the
outgoing map. The question arises what this means on the π0 -level where we do not have
homomorphisms. The image is defined for an arbitrary map. For the kernel we recall that π0 is
a set together with a distinguished element which corresponds to the neutral element of a group.
It is therefore natural to define the kernel of a map to be the set of all elements in the domain of
the map which are mapped to the distinguished element. Having clearified this, exactness of the
above sequence also makes sense on the π0 -level.

Proof. a) Exactness at πn (E; e0 ) for n ≥ 0:


i) im(ι# ) ⊂ ker(p# ):
Since p ◦ ι is the constant map, we have for any [σ]∂W n ∈ πn (F; e0 )

p# (ι# ([σ]∂W n )) = (p ◦ ι)# ([σ]∂W n ) = [p ◦ ι ◦ σ]∂W n = [ε bn0 ]∂W n = 0

ii) ker(p# ) ⊂ im(ι# ):


Let [τ]∂W n ∈ πn (E; e0 ) with p# ([τ]∂W n ) = [p ◦ τ]∂W n = 0. Hence p ◦ τ ≃∂W n ε bn0 .
Let h : W n × [0, 1] → B be a homotopy in B relative to ∂W n from p ◦ τ to ε bn0 . Lift
the map h to a homotopy H : W n × [0, 1] → E with initial conditions τ, i.e. H (·, 0) = τ.
The red area in the diagram gets mapped to F by H because it gets mapped to b0 by h.

[0, 1]

Wn

We obtain a homotopy in E relative to ∂W n from τ to H ◦ η −1 −1


n−1 . Hence τ ≃ ∂W H ◦ η n−1
n

and we conclude that

[τ] E, ∂W n = [H ◦ η −1 −1
n−1 ] E, ∂W = ι# ([H ◦ η n−1 ] F, ∂W ) ∈ im ι# .
n n

b) Exactness at πn (F; e0 ) for n ≥ 0:


i) im ∂ ⊂ ker ι# :
Let [σ]∂W n+1 ∈ πn+1 (B; b0 ). Then we have ∂([σ]∂W n+1 ) = [Σ ◦ η −1
n ] F, ∂W n where Σ is
n
a lift of σ with initial conditions ε e0 .

70
2.6 Higher homotopy groups

ε en0

The map Σ yields a homotopy relative to ∂W n in E from ε en0 to Σ ◦ η −1


n−1 . Hence

ι# (∂([σ]∂W n+1 )) = ι# ([Σ ◦ η −1 −1 n


n−1 ] F, ∂W n ) = [Σ ◦ η n−1 ] E, ∂W n = [ε e 0 ] E, ∂W n = 0.

ii) ker ι# ⊂ im ∂:
Let [τ]∂W n ∈ πn (F; e0 ) with 0 = ι# ([τ]∂W n ). Hence τ ≃∂W n ε en0 in E. Let H be a
homotopy in E relative to ∂W n from ε en0 to τ. Then H is a lift of h := p ◦ H. Since
H maps ∂W n+1 to F, the map h maps ∂W n+1 to b0 . Thus h represents an element in
πn+1 (B; b0 ). By definition of ∂ we have [H ◦ η −1
n−1 ]∂W = ∂([h]∂W n+1 ).
n

[0, 1]
ε en0

Wn

In the image of Boxn under η −1


n we let the interior cube grow and thereby obtain a
homotopy in F relative to ∂W from H ◦ η −1
n
n−1 to τ.

71
2 Homotopy Theory

e0
e0

e0 τ e0 deform
−→ e0 τ e0 deform
−→ τ

e0
e0

Therefore [τ]∂W n = [H ◦ η −1
n−1 ]∂W = ∂[h]∂W n+1 ∈ im(∂).
n

c) Exactness at πn (B; b0 ) for n ≥ 1:


i) im p# ⊂ ker ∂:
Let [τ]∂W n ∈ πn (E; e0 ). Then τ is a lift of p ◦ τ with initial conditions ε en−1
0
. Since τ
n −1 n−1
maps the boundary of W to e0 we have τ ◦ η n−1 = ε e0 . Thus
∂(p# ([τ]∂W n ) = ∂([p ◦ τ]∂W n )) = [τ ◦ η −1 n−1
n−1 ]∂W n−1 = [ε e 0 ]∂W n−1 = 0.

ii) ker ∂ ⊂ im p# :
Let [σ]∂W n ∈ πn (B; b0 ) with ∂([σ]∂W n ) = 0. Let Σ be a lift of σ with initial condition
ε en−1
0
. Then Σ ◦ η −1
n−1 represents ∂([σ]∂W ) = 0. Hence we have in F
n

Σ ◦ η −1 n−1
n−1 ≃ ∂W n−1 ε e 0

e0

maps to F

e0 F F e0
Σ

e0

Now choose a homotopy in F relative to ∂W n−1 from Σ◦η −1 n−1


n−1 to ε e 0 . Use this homotopy
to continuously extend Σ to the larger cube with boundary values e0 ; call this extension
τ. We then have
[τ]∂W n ∈ πn (E; eo ) and p# ([τ] E, ∂W n ) = [p ◦ τ]∂W n .
Now p ◦ τ is homotopic relative to ∂W n to σ as shown below.

72
2.6 Higher homotopy groups

b0
deform deform
−→ −→

b0 σ b0 σ σ

Thus [σ]∂W n = [p ◦ τ]∂W n ∈ im p# . 

Definition 2.104. A fiber bundle with discrete fiber is called a covering.

Corollary 2.105. If p : E → B is a covering with e0 ∈ F, b0 = p(e0 ) ∈ B, then the map

p# : πn (E, e0 ) → πn (B; b0 )

is an isomorphism for all n ≥ 2.

Proof. The assertion follows from the long exact sequence:


ι# p# ∂
{0} = πn (F; eo ) / πn (E; e0 ) / πn (B; b0 ) / πn−1 (F; e0 ) = {0}. 

Example 2.106. The map Exp : R → S1 with t 7→ e2πit is a covering with fiber Z. Hence
πk (S1 ; 1)  πk (R, 0) = {0}
for all k ≥ 2. More generally, Exp : Rn → T n = S1 × . . . × S1 with
(t 1, . . . , t n ) 7→ (e2πit 1 , . . . , e2πit n )
is a covering with fiber Zn . Hence πk (T n )  πk (Rn ) = {0} for all k ≥ 2.

Example 2.107. Consider the real projective space, defined by RPn := S n / ∼, where x ∼ y
⇐⇒ x = y or x = −y. Then the map p : S n → RPn with x 7→ [x]∼ is a covering with fiber
Z/2Z. Hence πk (RPn )  πk (S n ) for all k ≥ 2. For n ≥ 2 we investigate the sequence:
{0} = π1 (S n ) / π1 (RPn ) / π0 (Z/2Z) / π0 (S n ) = {0}

73
2 Homotopy Theory

We deduce that π1 (RPn )  π0 (Z/2Z)  Z/2Z as sets. But then π1 (RPn )  Z/2Z also as groups
because there is only one group of order 2 (up to isomorphism).

Example 2.108. We know that SO(n + 1)/SO(n) ≈ S n from Example 2.99. In the case of n = 2
this means that SO(3)/SO(2) ≈ S2 . We also know that SO(2) ≈ S1 . For k ≥ 3 we consider the
long exact sequence

··· / πk (SO(2)) / πk (SO(3)) / πk (S 2 ) / πk−1 (SO(2))

We know that πk (S1 ) = {0} and πk−1 (S1 ) = {0}. Hence we find

πk (SO(3))  πk (S2 ) for all k ≥ 3.

The map f : S3 → SO(3) given by

* +
x 2 + y 2 − u2 − v 2 2(yu − xv) 2(yv − xu)
f (x, y, u, v) = .. 2(yu + xv) x2 − y 2 + u2 − v 2 2(uv − xy) //
, 2(yv − xu) 2(uv + xy) x 2 − y 2 − u2 + v 2 -

satisfies f (−x, −y, −z, −v) = f (x, y, z, v) and therefore induces a continuous map
f¯ : RP3 → SO(3). This map is bijective and thus a homeomorphism. Hence we have
SO(3) ≈ RP3 . It follows that

πk (S2 )  πk (SO(3))  πk (RP3 )  πk (S3 )

for all k ≥ 3. Later we will see (Example 3.121 on page 151) that π3 (S3 )  Z and consequently
π3 (S2 )  Z.

Example 2.109. By the same proof as for SO(n + 1)/SO(n) ≈ S n we get that

SU(n + 1)/SU(n) ≈ S2n+1 .

Therefore there is a fiber bundle SU(n + 1) → S2n+1 with fiber SU(n). For n ≥ 1 we consider
the following part of the long exact homotopy sequence:
ι# p#
π0 (SU(n)) / π0 (SU(n + 1)) / π0 (S 2n+1 ) = {0}.

Hence the map ι# : π0 (SU(n)) → π0 (SU(n+1)) is onto. Since SU(1) = {1} we have π0 (SU(1)) =
{0} and thus π0 (SU(n)) = {0} by induction on n. Thus SU(n) is path-connected for all n ≥ 1.
Now let us analyze π1 (SU(n)). Consider
ι# p#
π1 (SU(n)) / π1 (SU(n + 1)) / π1 (S 2n+1 ) = {0}.

Again, we conclude that the map ι# : π1 (SU(n)) → π1 (SU(n+1)) is onto. By the same induction
as before we find π1 (SU(n)) = {0} for all n ≥ 1. Thus SU(n) is simply connected for all n ≥ 1.

74
2.7 Exercises

2.7 Exercises
2.1. Let X be a set and let x 0 ∈ X . Determine π1 (X ; x 0 ) where

a) X carries the discrete topology;

b) X carries the coarse topology.

2.2. Let X be a topological space and let ω : S1 → X be continuous. Show that the following
are equivalent:

(i) ω is homotopic to a constant map.

(ii) ω has a continuous extension D 2 → X .

2.3. Let X be a topological space and let f , g : X → S n be continuous. Assume f (x) , −g(x)
for all x ∈ X . Show that f and g are homotopic.

2.4. Let X and Y be topological spaces. Show that X × Y is contractible if and only if X and Y
are contractible.

2.5. Let X1 , X2 be topological spaces, x i ∈ X i . Put X := X1 × X2 and x := (x 1, x 2 ) ∈ X . Let


pi : X → X i be the canonical projections. Show that

(p1#, p2# ) : π1 (X ; x) → π1 (X ; x 1 ) × π1 (X2 ; x 2 )

is a group isomorphism.

2.6. Let X = [0, 1] × [0, 1] ⊂ R2 and let A ⊂ X be the comb space. Show that there is no
retraction X → A.

2.7. For f ∈ C (S1, R2 ) and p ∈ R2 \ f (S1 ) consider f p ∈ C (S1, S1 ) given by

f (z) − p
f p (z) = .
| f (z) − p|
Then
U ( f , p) := deg( f p )
is called the winding number of f around p.

a) Show that for p, q ∈ R2 \ f (S1 ) which can be joined by a continuous path in R2 \ f (S1 ) we
have U ( f , p) = U ( f , q).

b) Compute U ( f n, p) for all p ∈ R2 \ f (S1 ) and all n ∈ Z where f n (z) = z n .

75
2 Homotopy Theory

2.8. Show that the system of equations



cos 1 + x 2 y3 + sin(xy2 ) − x 2 = 0,
1
y+ = 0,
cosh(x + y + 10)
has a solution (x, y) ∈ R2 .

2.9. a) Compute π1 (D 2 \ {x 0 }; x 1 ) for all x 0 , x 1 ∈ D 2 .


b) Show that each homeomorphism f : D 2 → D 2 maps the boundary onto itself, f (∂D 2 ) =
∂D 2 .

2.10. On [0, 1] × [−1, 1] consider the equivalence relation ∼ given by (t, s) ∼ (t ′, s ′ ) iff (t, s) =
(t ′, s ′ ) or |t − t ′ | = 1 and s ′ = −s. The quotient space M := [0, 1] × [−1, 1]/ ∼ is called the
Möbius strip. The image S of [0, 1] × {0} is called the chord of M, that of [0, 1] × {−1, 1} is the
boundary ∂ M of M.
a) Show that ∂ M is homeomorphic to S1 .
b) Show that S is a strong deformation retract of M.
c) Determine π1 (M; x 0 ) for some x 0 ∈ ∂ M and the subgroup ι# (π1 (∂ M; x 0 )) ⊂ π1 (M; x 0 )
where ι : ∂ M ֒→ M is the inclusion map.
d) Show that ∂ M is not a retract of M.

2.11. Let X = {(t, nt ) | 0 ≤ t ≤ 1, n ∈ N} ∪ {(s, 0) | 12 ≤ s ≤ 1} ⊂ R2 equipped with the


induced topology. Show that X is connected but not path-connected.

2.12. Let G1 and G2 be groups. Show that G1 ∗ G2  G2 ∗ G1


a) using the construction of the free product;
b) using the universal property.

2.13. Let G1 and G2 be groups. Show:


a) If g ∈ G1 ∗ G2 has finite order then g is conjugate to an element in i1 (G1 ) or in i 2 (G2 ).
b) If G1 and G2 are nontrivial then G1 ∗ G2 contains elements of infinite order.

2.14. Let X be a path-connected topological space. Show that the suspension ΣX (see Exer-
cise 1.11) is also path-connected.

2.15. Show that the map C → C, z 7→ z 2 , has the homotopy lifting property for W 0 but not for
W 1.

76
2.7 Exercises

2.16. Let X = R3 \ (S1 × {0}) = R3 \ {(x, y, z) ∈ R3 | x 2 + y2 = 1, z = 0}. Show:

π1 (X, {0})  Z

and draw a generator of the fundamental group.

2.17. On S n consider the equivalence relation ∼ given by x ∼ y ⇔ x = y or x = −y. The


quotient RPn := S n / ∼ is called the n-dimensional real projective space.
Show inductively using the Seifert-van Kampen theorem that for n ≥ 2

π1 (RPn )  Z/2Z.

Hint: Exercise 2.10 may be helpful for the induction base n = 2.

2.18. Decide by proof or counter-example whether or not the following assertion holds true:
The Seifert-van Kampen theorem also holds if one only assumes that U ∩ V is connected rather
than path-connected.

2.19. Let E = B × F and p = pr1 : E → B be the product fibration. Show that the boundary
map ∂ : πn+1 (B; b0 ) → πn (F; e0 ) is trivial in this case.

2.20. Let p : E → B be a Serre fibration. Show that for [σ]∂W n+1 ∈ πn+1 (B; b0 ) the class
[Σ ◦ η −1
n ]∂W n ∈ π n (F; e0 ) does not depend on the choice of lift Σ of σ with the “initial condition”
ε en0 .

2.21. Show that the inclusion SU(n) ֒→ U(n) induces an isomorphism

πk (SU(n))  πk (U(n))

for all k ≥ 2.

2.22. The complex projective space is defined as CP n := S2n+1 / ∼ where z ∼ w iff there exists
u ∈ S1 ⊂ C such that z = u · w. Here we have regarded S2n+1 = {z ∈ Cn+1 | |z| = 1} as a subset
of Cn+1 . The map p : S2n+1 → CP n , z 7→ [z]∼ , is called the Hopf fibration.
Compute
πk (CP n )
for all k ≤ 2n.
Hint: You can use πk (S m ) = {0} for all 1 ≤ k < m and m ≥ 2.

2.23. Let p : X → Y and q : Y → Z be Serre fibrations. Prove that q ◦ p : X → Z is a Serre


fibration as well.

2.24. Let p : E → B be a Serre fibration, e0 ∈ E, b0 = p(e0 ) and F = p−1 (b0 ). Show

77
2 Homotopy Theory

a) If F is simply connected then π1 (E; e0 )  π1 (B; b0 ).

b) If E is contractible then πn+1 (B; b0 )  πn (F; e0 ) for all n ≥ 1.

i p
2.25. Let 0 → A → − A′ − → A′′ → 0 be an exact sequence of abelian groups. Show that the
following three conditions are equivalent:

(i) There exists an isomorphism Ψ : A′ → A × A′′ such that the following diagram commutes:

i p
0 /A / A′ / A′′ /0

= Ψ =
  
0 /A / A × A′′ / A′′ /0

where the arrows A → A× A′′ and A× A′′ → A′′ are given by the canonical maps a 7→ (a, 0)
and (a, a ′′ ) 7→ a ′′, respectively.

(ii) There exists a homomorphism p′ : A′′ → A′ such that p ◦ p′ = id A′′ .

(iii) There exists a homomorphism r : A′ → A such that r ◦ i = id A .

If these conditions hold then we say that the exact sequence is split.

i p
2.26. a) Show that every exact sequence 0 → A → − A′ − → A′′ → 0 of vector spaces (i.e. all
spaces are vector spaces over some fixed field and all homomorphisms are linear maps) is
split.
2
b) Show that the exact sequence 0 → Z →
− Z → Z/2Z → 0 is not split.

78
3 Homology Theory
Homotopy groups are in general hard to compute. For instance, for spheres not all homotopy
groups are known even now. In this chapter we introduce rougher invariants which are much
easier to determine, the homology groups.

3.1 Singular homology


We will use the notation

e0 = (0, . . . , 0) ∈ Rn
e1 = (1, . . . , 0) ∈ Rn
..
.
en = (0, . . . , 1) ∈ Rn

Now we define

∆n := convex hull of e0, . . . , en




n
n 

 1
X X
= t i ei t i ≥ 0, .

ti ≤
 i=0 i=0 
Then ∆n is called n-dimensional standard simplex.

Example 3.1. For n = 0, . . . , 3 the standard simplices can be visualized as:

n=0 ∆0 = {e0 } = point


n=1 ∆1 = [0, 1] = line segment

e2 b

n=2 ∆2 = triangle b b

e0 e1

79
3 Homology Theory

∆3

n=3 ∆3 = tetrahedron b b

Definition 3.2. Let X be a topological space. A singular n-simplex in X is a continuous map


∆n → X .

Now we fix a commutative ring R with 1. The most important examples will be
R = R, Q, C, Z, Z/nZ.

Definition 3.3. We define the set of singular n-chains by

Sn (X ; R) := free R-module generated by all n-dimensional singular simplices in X


m

α i σi α i ∈ R, σi ∈ C (∆n, X ), n ∈ N0 .
 X 

= formal linear combinations
i=1

Then Sn (X ; R) carries an obvious structure as an R-module.

i
For n ≥ 0 we consider the affine linear map Fn+1 : ∆n → ∆n+1 given by


e j,
(e j ) = 
j <i
 e j+1,
i
Fn+1 (3.1)
 j ≥i

Note that Fni maps ∆n to the face of ∆n+1 opposite to ei .

Example 3.4. Consider the special case F21 :

80
3.1 Singular homology

X
e2 σ
b

F21

b
b

e0 e1 e0 e1

σ (i)

Definition 3.5. If σ is an (n + 1)-dimensional singular simplex in X then σ (i) := σ ◦ Fni is


called the i-th face of σ.

The boundary of a singular n-simplex in X is given by:


n
X
∂σ := (−1) i σ (i) .
i=0

We see that the boundary of a singular n-simplex is a singular (n − 1)-chain. We extend ∂ to


chains by linearity. The boundary of a singular n-chain in X is thus given by
m
X  Xm
∂ αjσj = α j ∂σ j .
j=0 j=0

Hence we obtain a linear map ∂ : Sn (X ; R) → Sn−1 (X ; R) and we set ∂(0-chain) := 0.

Lemma 3.6. ∂ ◦ ∂ = 0.

Proof. It suffices to prove ∂∂σ = 0 for all n-simplices σ. For j < i we have
j
Fni ◦ Fn−1 = Fnj ◦ Fn−1
i−1
. (3.2)

We compute
n
∂∂σ = ∂ * (−1) i σ ◦ Fni +
X

, i=0 -
Xn  
= (−1) i ∂ σ ◦ Fni
i=0

81
3 Homology Theory

n
X n−1
X
i j
= (−1) (−1) j σ ◦ Fni ◦ Fn−1
i=0 j=0
X X
j j
= (−1) i+ j σ ◦ Fni ◦ Fn−1 + (−1) i+ j σ ◦ Fni ◦ Fn−1
0≤ j <i ≤n 0≤i ≤ j ≤n−1
X X
j ′ ′ j′ ′
= (−1) i+ j σ ◦ Fn i−1
◦ Fn−1 + i −1
(−1) j +i −1 σ ◦ Fn ◦ Fn−1
0≤ j <i ≤n 0≤ j ′ <i ′ ≤n
= 0.
In the last step we used (3.2) for the first sum and changed the summation indices from i → j ′
and j → i ′ − 1 in the second sum. 

Now we define the set of singular n-cycles by


Z n (X ; R) := ker(∂ : Sn (X ; R) → Sn−1 (X ; R))
={c ∈ Sn (X ; R) | ∂c = 0}
and the set of singular n-boundaries by
Bn (X ; R) := im(∂ : Sn+1 (X ; R) → Sn (X ; R))
={c ∈ Sn (X ; R) | ∃ b ∈ Sn+1 (X ; R) such that c = ∂b} .

Lemma 3.6 says Bn (X ; R) ⊂ Z n (X ; R).

Definition 3.7. The quotient


Z n (X ; R)
Hn (X ; R) :=
Bn (X ; R)
is called n-th singular homology of X with coefficients in R.

Remark 3.8. The n-th homology Hn (X ; R) is an R-module.

Example 3.9. Assume that X = {point}. Then there is only one singular n-simplex, namely
the constant map σ n : ∆n → X . In other words, Sn (X ; R) = R · σ n . Consequently,
σ n(i) = σ n ◦ Fni = σ n−1 and



n 

0, n odd,

X
i
=  σ n−1,

∂σ n = (−1) σ n−1 n even, n , 0,

 0,

i=0 n = 0.
This implies

 R · σ n,
Z n (X ; R) = 
n odd or n = 0,
 0,
 n even and n , 0,

82
3.1 Singular homology

and

 R · σ n,
Bn (X ; R) = 
n odd,
 0,
 n even.
We conclude that

 R, if n = 0,
Hn (X ; R)  
 0, otherwise.

As in homotopy theory we not only associate groups to spaces but also homomorphisms to
maps. Let f : X → Y be a continuous map. Then we obtain an R-module homomorphism
Sn ( f ) : Sn (X ; R) → Sn (Y ; R) by setting
m
X  m
X
Sn ( f ) α i · σi := α i · ( f ◦ σi ).
i=1 i=1

Lemma 3.10. The diagram

Sn ( f )
Sn (X ; R) / Sn (Y ; R)

∂ ∂
 S n−1 ( f ) 
Sn−1 (X ; R) / Sn−1 (Y ; R)

commutes for all n ≥ 0.

Proof. We compute
∂(Sn ( f )(σ)) = ∂( f ◦ σ)
Xn
= (−1) i ( f ◦ σ) ◦ Fni
i=0
n
X
= (−1) i f ◦ (σ ◦ Fni )
i=0
n
X 
= Sn−1 ( f ) (−1) i σ ◦ Fni
i=0
= Sn−1 ( f )(∂σ). 

This lemma implies Sn ( f )(Z n (X ; R)) ⊂ Z n (Y ; R) and Sn ( f )(Bn (X ; R)) ⊂ Bn (Y ; R). Hence
we obtain a well-defined R-module homomorphism Hn ( f ) : Hn (X ; R) → Hn (Y ; R) where
Hn ( f )([x]) := [Sn ( f )(X )]. Here the square brackets denote the homology classes of the
n-cycles. One sees directly from the definition that Hn (·) has the functorial properties
(i) Hn (id X ) = idH n (X ;R) ,

83
3 Homology Theory

(ii) Hn ( f ◦ g) = Hn ( f ) ◦ Hn (g).

Exactly as for homotopy groups these functorial properties imply that a homeomorphism f :
X → Y induces an isomorphism Hn ( f ) : Hn (X ; R) → Hn (Y ; R). Homeomorphic spaces have
isomorphic homology groups.

3.2 Relative homology


For a topological space X and A ⊂ X we call (X, A) a pair of spaces and set

C ((X, A), (Y, B)) := { f ∈ C (X, Y ) | f (A) ⊂ B} .

We abbreviate Sn (X ) = Sn (X ; R) if R is understood. We observe that Sn (A) ⊂ Sn (X ) and


∂(Sn (A)) ⊂ Sn−1 (A). Writing Sn (X, A) = Sn (X, A; R) := Sn (X ; R)/Sn (A; R), the map ∂
induces a well-defined homomorphism ∂¯ such that

Sn (X ) / Sn (X, A)

∂ ∂¯
 
Sn−1 (X ) / Sn−1 (X, A)

commutes. Since ∂ ◦ ∂ = 0 and since Sn (X ) → Sn (X, A) is onto we also have that ∂¯ ◦ ∂¯ = 0. Set

Z n (X, A) = Z n (X, A; R) := ker ∂¯ : Sn (X, A) → Sn−1 (X, A) ,

Bn (X, A) = Bn (X, A; R) := im ∂¯ : Sn+1 (X, A) → Sn (X, A) .

We define the relative singular homology of (X, A) by


Z n (X, A; R)
Hn (X, A) = Hn (X, A; R) := .
Bn (X, A; R)
Now consider the preimage of Z n (X, A) under Sn (X ) → Sn (X, A) and set

Z n′ (X, A) := {c ∈ Sn (X ) | ∂c ∈ Sn−1 (A)},


Bn′ (X, A) := {c ∈ Sn (X ) | ∃b ∈ Sn+1 (X ) such that c + ∂b ∈ Sn (A)}.

Since Z n (X, A) = Z n′ (X, A)/Sn (A) and Bn (X, A) = Bn′ (X, A)/Sn (A) we obtain

Z n′ (X, A)/Sn (A) Z n′ (X, A)


Hn (X, A) = = .
Bn′ (X, A)/Sn (A) Bn′ (X, A)

Remark 3.11. For A = ∅ we have the special cases

Z n′ (X, ∅) = Z n (X ),
Bn′ (X, ∅) = Bn (X ),
Hn (X, ∅) = Hn (X ).

84
3.2 Relative homology

Example 3.12. Let X = S1 × [0, 1] be the cylinder over S1 and A = S1 × {0} ⊂ X .

To construct an element in the relative homology H1 (X, A) take a 1-simplex

σ : ∆1 → X,
(te1 + (1 − t)e0 ) 7→ (cos (2πt), sin (2πt), 1).

|
>
σ X

∆1
A

Since σ is a closed curve in X we find for its boundary

∂σ = σ(e1 ) − σ(e0 ) = 0.

85
3 Homology Theory

Therefore σ ∈ Z1 (X ) ⊂ Z1′ (X, A) and, as we will see later, represents a nontrivial element in
H1 (X ). For the 2-simplices σ1 and σ2 defined as indicated in the following picture

σ
>
>
σ1 X

∆2

σ2

we find for the 2-chain σ1 + σ2 the boundary ∂(σ1 + σ2 ) = σ + a where a ∈ S1 (A). Hence,
modulo a ∈ S1 (A), we have σ ∈ B1 (X, A) and therefore 0 = [σ] ∈ H1 (X, A).

Let (X, A) be a pair of spaces. The inclusion map i : A ֒→ X induces a homomorphism


Hn (i) : Hn (A) → Hn (X ). Furthermore we have the inclusion map j : (X, ∅) → (X, A) which
induces the homomorphism Hn ( j) : Hn (X ) = Hn (X, ∅) → Hn (X, A).
We define the connecting homomorphism or boundary operator

∂ : Hn (X, A) → Hn−1 (A), (3.3)


[c] 7→ [∂c],

where c ∈ Z n′ (X, A). Note that ∂c ∈ Z n−1 (A) since ∂ 2 = 0. The connecting homomorphism is
well defined because replacing c by another representative c + ∂b + a where b ∈ Sn+1 (X ) and
a ∈ Sn (A) yields

c + ∂b + a 7→ ∂(c + ∂b + a) = ∂c + ∂a.

Since ∂a ∈ Bn−1 (A) we get [∂c + ∂a] = [∂c] ∈ Hn−1 (A). Since ∂ : Z n′ (X, A) → Sn−1 (A) is a
homomorphism, the connecting homomorphism is also a homomorphism.

86
3.3 The Eilenberg-Steenrod axioms and applications

Lemma 3.13. The connecting homomorphism is natural, i.e. the diagram

Hn ( f )
Hn (X, A) / Hn (Y, B)

∂ ∂
 H n−1 ( f | A ) 
Hn−1 (A) / Hn−1 (B)

commutes for every f ∈ C ((X, A), (Y, B)).

Proof. We compute
X    X 
Hn−1 ( f | A ) ∂ α i σi = Hn−1 ( f | A ) ∂ α i σi
i i
X X
j 
= Hn−1 ( f | A ) αi (−1) j σ j ◦ Fn−1
i j
X X
j 
= αi (−1) j ( f | A ) ◦ (σi ◦ Fn−1 )
i j
X X
j 
= αi (−1) j ( f ◦ σi ) ◦ Fn−1
i j
 X 
= ∂ α i ( f ◦ σi )
i
X 
=∂ α i ( f ◦ σi )
i
X  
= ∂ Hn ( f ) α i σi . 
i

3.3 The Eilenberg-Steenrod axioms and applications


Now we list the most important properties of homology theory known as Eilenberg-Steenrod
axioms. The first axiom is exactly Example 3.9.

Dimension Axiom. (
R, n = 0
Hn ({point}; R) 
0, otherwise
The next axiom deals with homotopy invariance. Two continuous maps f 0, f 1 : (X, A) → (Y, B)
of pairs of spaces are called homotopic (in symbols f 0 ≃ f 1 ) iff there exists an H ∈ C (X ×[0, 1], Y )

87
3 Homology Theory

such that for all x ∈ X

H (x, 0) = f 0 (x),
H (x, 1) = f 1 (x),
H (A × [0, 1]) ⊂ B.

Homotopy Axiom.
Let f 0, f 1 ∈ C ((X, A), (Y, B)) be homotopic, f 0 ≃ f 1 . Then the induced maps on homology
coincide, i.e.,
Hn ( f 0 ) = Hn ( f 1 ) : Hn (X, A) → Hn (Y, B)
holds for all n.

Remark 3.14. Let (X, A) ≃ (Y, B), i.e., there exist f ∈ C ((X, A), (Y, B)) and
g ∈ C ((Y, B), (X, A)) such that f ◦ g ≃ id(Y, B) and g ◦ f ≃ id (X, A) . It then follows as be-
fore that Hn (X, A; R)  Hn (Y, B; R).

Exactness Axiom. For any pair of spaces (X, A) and inclusion maps i : A ֒→ X ,
j : (X, ∅) ֒→ (X, A), the sequence

H n+1 ( j)
Hn+1 (X, A) o ···

 H n (i) H n ( j)
Hn (A) / Hn (X ) / Hn (X, A)


H n−1 (i) 
··· o Hn−1 (A)

is exact and natural.

Notation: For A ⊂ X we call


[
Å = U with U open in X
U ⊂A

the interior of A and \


Ā = B with B closed in X
B⊂X
the closure of A.

Excision Axiom.
For every pair of spaces (X, A) and every U ⊂ A with Ū ⊂ Å the homomorphism

Hn ( j) : Hn (X \ U, A \ U ) → Hn (X, A)

88
3.3 The Eilenberg-Steenrod axioms and applications

induced by the inclusion map

j : (X \ U, A \ U ) ֒→ (X, A)

is an isomorphism.
The proofs of axioms A2, A3, and A4 will be given later. Before that we wil show their usefulness
by studying some basic examples.

Remark 3.15. (cf. Exercise 3.1)


1.) If X , ∅ is path-connected then H0 (X ; R)  R and generators are represented by every
0-simplex σ : ∆0 → X . In other words, the isomorphism H0 (X ; R) → R is given by
P P
j α j σ j 7→ j αj.
L
2.) If X k , k ∈ K, are the path-components of X then Hn (X ; R)  k ∈K H n (X k ; R).

Theorem 3.16. For n ≥ 1 we have


(
n R, if m = 0 or m = n
Hm (S ; R) 
0, otherwise
(
n n−1 R, if m = n
Hm (D , S ; R) 
0, otherwise

Proof. a) For n ≥ 1 the sphere S n is path-connected and therefore H0 (S n ; R)  R.


For n = 0 we have that S0 = {x, y} with the discrete topology and therefore H0 (S0 ; R)  R2 .
b) We have the exact sequence
H0 (S0 ) −→ H0 (D 1 ) −→ H0 (D 1, S0 ) −→ 0


R2 R

(a, b) 7−→ a+b

Since the map (a, b) 7→ a + b is onto, the map H0 (D 1 ) → H0 (D 1, S0 ) must be zero. Hence
H0 (D 1, S0 ) = 0.
For n ≥ 2 we have the following exact sequence
H0 (S n−1 ) −→ H0 (D n ) −→ H0 (D n, S n−1 ) −→ 0
id


R −→ R

Again the second arrow has to be trivial and therefore H0 (D n, S n−1 ) = 0. This settles the
case m = 0.

89
3 Homology Theory

c) We will now use the exact sequence

H1 (D n ) −→ H1 (D n, S n−1 ) −→ H0 (S n−1 ) −→ H0 (D n )

For n = 1 we have
H1 (D 1 ) −→ H1 (D 1, S0 ) −→ H0 (S0 ) −→ H0 (D 1 )


H1 ({point}) R2 R


0 (a, b) 7−→ a+b

which implies H1 (D 1, S0 )  ker((a, b) 7→ a + b)  R.


For n ≥ 2 we have

0 
H1 (D n ) −→ H1 (D n, S n−1 ) −→ H0 (S n−1 ) −→ H0 (D n )
=

from which we get H1 (D n, S n−1 ) = 0.

d) Consider the pair of spaces (S n , D−n ) where D−/(+)


n = {x ∈ S n |x 0 ≥ 0 (x 0 ≤ 0)} is the lower
(resp. upper) hemisphere.

Sn

D−n

For n ≥ 1 we have


H1 (D−n ) −→ H1 (S n ) −→ H1 (S n, D−n ) −→ H0 (D n ) −→ H0 (S n )
=

Since the last arrow is an isomorphism, the connecting homomorphism


H1 (S n , D−n ) → H0 (D−n ) must be zero. Therefore H1 (S n ) → H1 (S n, D−n ) is onto.
Since H1 (D−n ) = 0 we get H1 (S n )  H1 (S n, D−n ).

e) Put U−n := {x ∈ S n | x 0 < 21 }.

90
3.3 The Eilenberg-Steenrod axioms and applications

Sn

D−n

U−n

By the excision axiom the inclusion


(S n \ U−n, D n \ U−n ) ֒→ (S n , D n )
induces an isomorphism
Hm (S n \ U−n, D n \ U−n )  Hm (S n, D n )
since Ū−n ⊂ D̊−n . We also have that
(S n \ U−n, D n \ U−n ) ≃ (D+n, S n−1 ) ≈ (D n, S n−1 )
where the homeomorphism is given by a vertical projection. This gives us the isomorphism
Hm (S n \ U−n, D n \ U−n )  Hm (D n, S n−1 ).
In particular,
(
n n R, for n = 1
H1 (S )  H1 (S , D−n ) n
 H1 (D , S n−1
)= (3.4)
0, otherwise
This concludes the case m = 1.
f) Finally we treat the case m ≥ 2 by induction. Observe that
Hm (D−n ) −→ Hm (S n ) −→ Hm (S n , D−n ) −→ Hm−1 (D n )
=

0 0
and
Hm (D n ) −→ Hm (D n, S n−1 ) −→ Hm−1 (S n−1 ) −→ Hm−1 (D n )
=

0 0
This yields
Hm (S n )  Hm (S n , D−n )  Hm (D n, S n−1 )  Hm−1 (S n−1 ) (3.5)
Induction over m concludes the proof. 

91
3 Homology Theory

Remark 3.17. Let us describe a generator of H1 (S1 )  Z geometrically. We use the isomor-
phisms in (3.4) and start with H1 (D 1, S0 ). The map c : ∆1 → D 1 , t 7→ cos(π(1 − t)), is a singular
1-simplex with ∂c = const1 − const−1 . Under the isomorphism H0 (S0 )  R2 it maps to (1, −1)
which generates the kernel of the map R2 → R given by (a, b) 7→ a + b. Hence c represents a
generator of H1 (D 1, S0 ).
The isomorphism H1 (D+1 , S0 ) → H1 (D 1, S0 ) induced by vertical projection gives us a generator
of H1 (D+1 , S0 ), namely the homology class represented by c ′ : ∆1 → D+1 , t 7→ eiπ (1−t ) .
The isomorphism H1 (D+1 , S0 ) → H1 (S1, D−1 ) is induced by the inclusion. Hence [c ′] ∈
H1 (S1, D−1 ) is a generator. Now

F : S1 × [0, 1] → S1, (t, s) 7→ eiπ (1−(s+1)t ) ,

is continuous and satisfies

F (t, 0) = c ′ (t),
F (t, 1) = eiπ (1−2t ) =: c ′′ (t),
F (0, s) = −1 ∈ D−1 ,
F (1, s) = e−iπ s ∈ D−1 .

Thus c ′ ≃ c ′′ as maps (∆1, {0, 1}) → (S1, D−1 ). Using the homotopy axiom we compute

[c ′ ] = [c ′ ◦ id∆1 ] = H1 (c ′ )[id∆1 ] = H1 (c ′′ )[id∆1 ] = [c ′′ ◦ id∆1 ] = [c ′′].

Therefore [c ′′] ∈ H1 (S1, D−1 ) is a generator. Since ∂c ′′ = const c ′′ (1) − constc ′′ (0) = const−1 −
const−1 = 0, the 1-simplex c ′′ represents a homology class in H1 (S1 ). Moreover, since inclusion
induces an isomorphism H1 (S1 ) → H1 (S1, D−1 ) we find that [c ′′] ∈ H 1 (S1 ) is a generator.
Using the homotopy (t, s) 7→ eiπ (s−2t ) we see that t 7→ e−2iπt also represents a generator of
H1 (S1 ). Finally, playing the same game with lower and upper hemispheres interchanged shows
that t 7→ e2iπt represents a generator of H1 (S1 ) as well.

As a first application of Theorem 3.16 we now prove Brouwer’s fixed point theorem in all
dimensions.

Theorem 3.18 (Brouwer’s fixed point theorem). Let f : D n → D n , n ≥ 1, be a continuous


map. Then f has a fixed point, i.e., there exists an x ∈ D n such that f (x) = x.

Proof. We assume that the map f does not have a fixed point and then derive a contradiction.
Let n ≥ 2 (the case n = 1 having been treated in Remark 1.6).

92
3.3 The Eilenberg-Steenrod axioms and applications

Dn
Now consider the continuous map g : D n → S n−1 as in
x
the picture. Denote the inclusion map by ι : S n−1 → D n f (x)
• g(x)
and note that g ◦ ι = idS n−1 .
∂D n = S n−1

By the functorial properties of homology we obtain the following commutative diagram:


H n−1 (id)=id
Hn−1 (S n−1 ; Z)❙ Z / Hn−1 (S n−1 ; Z)  Z
5
❙❙❙
❙❙❙❙ ❦❦ ❦❦❦❦❦
❙❙❙ ❦
H n−1 (ι) ❙❙❙❙ ❦❦❦❦
) ❦❦❦ H n−1 (g)
Hn−1 (D n ; Z) = 0
Since the identity Z → Z does not factor through 0 we run into a contradiction. 

Now we are in the position to answer the third question on page 3.

Theorem 3.19. For n , m the space Rn is not homeomorphic to Rm .

Proof. Let us assume there exists a homeomorphism f : Rn → Rm . Then


f : Rn \ {0} → Rm \ { f (0)}
is also a homeomorphism. Since Rn \ {point} ≃ S n−1 we obtain an isomorphism on the level of
homology groups:
H j (S n−1 )  H j (Rn \ {0})  H j (Rm \ { f (0)})  H j (S m−1 ).
By Theorem 3.16 this is a contradiction for j = n − 1 or j = m − 1 unless n = m. 

Proposition 3.20. For n ≥ 1 let s : S n → S n be the reflection given by

s(x 0, x 1, . . . , x n ) = (−x 0, x 1, . . . , x n ).

Then the map


Hn (s) : Hn (S n ) → Hn (S n )
is given by Hn (s) = − id.

93
3 Homology Theory

Proof. The proof is given by induction. First consider the case n = 1. Let c : ∆1 → S1 be a
1-simplex generating H1 (S1 ) and let S1 (s)(c) be its image under the induced homomorphism on
1-chains.

S1 (s)

b
∧ b

We need to show that H1 (s)[c] = [S1 (s)c] = −[c]. We construct two 2-simplices. The first one
is given by

e2
b σ1

b

b b

e0 e1

Applying the boundary operator yields

∂σ1 = S1 (s)c − const + c.

The second 2-simplex is the constant map.

94
3.3 The Eilenberg-Steenrod axioms and applications

e2
b σ2

b b

e0 e1

We apply the boundary operator again and we get

∂σ2 = const − const + const = const

It follows that for the chain σ1 + σ2 that

∂(σ1 + σ2 ) = S1 (s)c + c

and hence
0 = [∂(σ1 + σ2 )] = [S1 (s)c + c] = [S1 (s)c] + [c] .
This yields the desired result in the case of n = 1.
The induction step n − 1 ⇒ n follows from the following commutative diagram:
  
Hn (S n ) / Hn (S n , D n ) o
− Hn (D+n, S n−1 ) / Hn−1 (S n−1 )

H n (s) H n (s) H n (s) H n−1 (s)


   
  
Hn (S n ) / Hn (S n , D n ) o
− Hn (D+n, S n−1 ) / Hn−1 (S n−1 )

where the horizontal isomorphisms are the ones in (3.5). Commutativity of the last square
follows from Lemma 3.13 and that of the first and the second one from the fact that the horizontal
isomorphisms are induced by inclusion maps (which commute with s). Note here that we have to
choose the lower hemisphere with respect to a different coordinate than the one which is reflected
by s so that s(D−n ) = D−n . 

Remark 3.21. Let a : S n → S n with a(x) = −x for all x ∈ S n be the antipodal map then
Hn (a) = (−1) n+1 . This follows from the fact that a is the composition of n + 1 reflections.

95
3 Homology Theory

Definition 3.22. A vector field on S n is a map v : S n → Rn+1 such that v(x) ⊥ x for all
x ∈ Sn .

x
v(x)
b

Theorem 3.23 (Combing a hedgehog). The n-dimensional sphere S n admits a continuous


vector field without zeros iff n is odd.
In particular, every continuous vector field on S2 has a zero.

Loosely speaking, this means that every continuously combed hedgehog has a “bald” spot.
Proof. If n is odd we simply set

v(x) := (−x 1, x 0, −x 3, x 2, . . . , −x n , x n−1 ) .

Now let n be even and let us assume that v : S n → Rn+1 is a


continuous vector field without a zero. Then we can put w(x) := x
v (x) n n
|v (x) | . We define the continuous map F : S × [0, 1] → S by
∧ b

F (x, t) := x cos(πt) + w(x) sin(πt) . w(x)


−x
Since F (x, 0) = x and F (x, 1) = −x = a(x) with a(x) the
antipodal map we have found that a ≃ id. Hence Hn (a) =
Hn (id) = 1.

This contradicts Hn (a) = (−1) n+1 = −1. 

3.4 The degree of a continuous map


For n ≥ 1 we consider a continuous map f : S n → S n . Then the homomorphism

Hn ( f ) : Hn (S n ; Z)  Z → Hn (S n ; Z)  Z

is given by multiplication with a number which we denote deg( f ) ∈ Z.

96
3.4 The degree of a continuous map

Definition 3.24. The number deg( f ) is called the degree of the map f . The degree can be
defined in the same way for continuous maps f : (D n, S n−1 ) → (D n, S n−1 ).

Examples 3.25. 1.) For a reflection we have deg(s) = −1.

2.) For the antipodal map we have seen that deg(a) = (−1) n+1 .

Lemma 3.26. The degree of a function has the following properties

(i) deg(id) = 1;

(ii) deg(const) = 0;

(iii) deg( f ◦ g) = deg( f ) deg(g);

(iv) If f ≃ g then deg( f ) = deg(g);

(v) If the map f is a homotopy equivalence then deg( f ) = ±1;

(vi) For f : (D n+1, S n ) → (D n+1, S n ) we have deg( f ) = deg( f S n ).

Proof. The first and third assertion follow directly from the functorial property of Hn ( f ). The
fourth and fifth statement follow from the homotopy axiom. The second assertion follows from
the fact that the homomorphism induced by a constant map factors through Hn (pt) = 0. The last
statement of the lemma follows fromm the commutativity of the following diagram:

H n+1 ( f )
Hn+1 (D n+1, S n ) / Hn+1 (D n+1, S n )

 ∂ ∂ 
 H n+1 ( f | S n ) 
Hn (S n ) / Hn (S n )

Theorem 3.27. (i) Every f ∈ C (S n, S n ) without fixed points satisfies deg( f ) = (−1) n+1 .

(ii) Every f ∈ C (S n, S n ) without an antipodal point, i.e., f (x) , −x for all x ∈ S n , satisfies
deg( f ) = 1.

(iii) For n even every f ∈ C (S n, S n ) has a fixed point or an antipodal point.

97
3 Homology Theory

Proof. (i) Let f ∈ C (S n, S n ) be without a fixed point. Then the line segment joining f (x)
and −x does not contain the origin.

x b

b
0

f (x) b b
−x

Hence we can define the continuous map


(1 − t) f (x) − t x
F : S n × [0, 1] → S n , F (x, t) := .
|(1 − t) f (x) − t x|
Since F (x, 0) = f (x) and F (x, 1) = −x = a(x) with a being the antipodal map the map F
is a homotopy for f ≃ a. Thus

deg( f ) = deg(a) = (−1) n+1 .

(ii) Now let f ∈ C (S n, S n ) be without antipodal points. Then we can define


(1 − t) f (x) + t x
G(x, t) := .
|(1 − t) f (x) + t x|

x b

b
0

b
−x
b

f (x)

Since G(x, 0) = f (x) and G(x, 1) = x we have that f ≃ id and deg( f ) = deg(id) = 1
follows.

(iii) Finally, assume that f ∈ C (S n, S n ) has neither fixed points nor antipodal points. Then
deg( f ) = (−1) n+1 by (i) and by deg( f ) = 1 by (ii). Thus n must be odd. 

98
3.4 The degree of a continuous map

Let µ : S n × S n → S n for n ≥ 1 be a continuous map. We choose p ∈ S n arbitrarily and define

j1 : S n → S n × S n , x 7→ (x, p),
n n n
j2 : S → S × S , x 7→ (p, x) .

We then get the following diagram:

(H n ( j1 ), H n ( j2 )) H n (µ)
Z2  Hn (S n ) ⊕ Hn (S n ) / Hn (S n × S n ) / Hn (S n )  Z
6

(d1, d2 )

with d µ ∈ Z.

Definition 3.28. The pair of numbers (d1, d2 ) ∈ Z2 is called the bidegree of the map µ.

Remark 3.29. The bidegree does not depend on the choice of p. If one chooses another p′ , then
a path from p to p′ will yield a homotopy between the corresponding embedding maps jν and jν′ .
Hence they induce the same homomorphisms on homology and therefore the same bidegrees.

Examples 3.30. 1.) Consider the case n = 1, S1 ⊂ C and let the map µ be given by µ(z1, z2 ) =
z1 z2 . Choose p = 1 ∈ S1 . Then µ ◦ j1 = id and thus

d1 = deg(µ ◦ j1 ) = deg(id) = 1.

Similarly, we get d2 = 1. Hence the bidegree by (d1, d2 ) = (1, 1).

2.) In the case of n = 3, S3 ⊂ H we consider the map µ given by quaternionic multiplication,


µ(h1, h2 ) = h1 h2 . Similar reasoning shows that the bidegree is again given by (d1, d2 ) =
(1, 1).

Remark 3.31. The quaternions H form a division algebra isomorphic to R4 as a vector space.
The algebra H is associate and noncommutative. The standard vector space basis we be denoted
by 1, i, j, k. Therefore any h ∈ H can be uniquely written as

h = h0 + h1 i + h2 j + h3 k .

Quaternionic multiplication is now determined by the relations i 2 = j 2 = k 2 = i j k = −1. With


the help of the conjugation
h ∗ = h0 − h1 i − h2 j − h3 k

we can define |h| := h∗ h. We regard S3 ⊂ H as the set of unit-length quaternions.

99
3 Homology Theory

Proposition 3.32. Let n = 1 or n = 3 and let k ∈ Z. The map f k : S n → S n with f k : z 7→ z k


given by complex multiplication in the case n = 1 and by quaternionic multiplication in the
case n = 3 has degree k.

Proof. The proof is by induction on k. For k = 0 and k = 1 the statement is trivial because
constant maps have degree 0 while the identity has degree 1. The case of k = −1 has already
been shown for n = 1, since here z 7→ z −1 = z̄ is a reflection and hence deg( f −1 ) = −1. On the
other hand for k = −1, n = 3 we note that S1 ⊂ S2 ⊂ S3 regarding C ⊂ H. Now the following
commutative diagram

H1 (S1 )  H2 (S2 )  H3 (S3 )


H1 ( f −1 | S 1 )=deg( f −1 | S 1 )=−1 H3 ( f −1 )=deg( f −1 )
 
H1 (S1 )  H2 (S2 )  H3 (S3 )

implies that deg( f −1 ) = −1 in the quaternionic case too. The horizontal isomorphisms are
obtained as the composition of the isomorphisms

Hk (S k )  Hk+1 (D k+1, S k )  Hk+1 (S k+1, D−k+1 )  Hk+1 (S k+1 ).

We used that the quaternionic multiplication restricted to the complex numbers C is just complex
multiplication.
Now we are ready to carry out the induction over k. We consider k > 0 and we show that the
statement for k − 1 implies that for k. Let µ be complex (resp. quaternionic) multiplication. Then

deg( f k ) = deg(µ ◦ ( f k−1, f 1 ))


!
deg( f k−1 )
= (1, 1) ·
deg( f 1 )
= deg( f k−1 ) + deg( f 1 )
= k − 1 + 1 = k.

The case k < 0 is treated similarly. 

Now we can show that the fundamental theorem of algebra 2.34 also holds for quaternionic
polynomials.

Theorem 3.33 (Fundamental theorem of algebra for quaternions). Every quaternionic


polynomial
p(z) = z k + α 1 z k−1 + . . . + α k

100
3.4 The degree of a continuous map

of positive degree k has a quaternionic root.

Proof. Suppose the polynomial p has no root. Then we can define the continuous map
p̂ : S3 → S3 with p̂(z) := | p(z) 3 3 p(t z)
p(z) | . Now consider F (z, t) : S × [0, 1] → S with F (z, t) := | p(t z) | .
We observe that F (z, 0) = const and F (z, 1) = p̂(z), hence we have p̂(z) ≃ const and conse-
quently deg( p̂) = 0.
On the other hand, we can put for z ∈ S3 and t > 0
 
t k p zt
G(z, t) :=  
|t k p zt |
and observe that the expression
z
k
t p = z k + tα 1 z k−1 + . . . + t k α
t
extends continuously to t = 0. The map G : S3 × [0, 1] → S3 satisfies G(z, 0) = f k (z) and again
G(z, 1) = p̂(z). Thus p̂ ≃ f k and hence deg( p̂) = k. This contradicts k > 0. 

Now let U ⊂ S n be open, n ≥ 1. Consider a continuous map f : U → S n and a point p ∈ S n


such that f −1 (p) is compact. Then we have the following diagram (with deg p ( f ) ∈ Z):

(i) (ii)
Z  Hn (S n ; Z) / Hn (S n , S n \ f −1 (p); Z) o Hn (U, U \ f −1 (p); Z) (3.6)

deg p ( f ) Hn ( f )
 (iii) 
Z  Hn (S n ; Z) 
/ Hn (S n, S n \ {p}; Z)

We observe:
Concerning (i): This homomorphism is induced by the inclusion
S n = (S n , ∅) ֒→ (S n, S n \ f −1 (p)).

Concerning (ii): The inclusion (U, U \ f −1 (p)) ֒→ (S n, S n \ f −1 (p)) induces an isomorphism


by the excision axiom.
Concerning (iii): Consider the exact homology sequence of (S n, S n \ {p})

0 = Hn (S n \ {p}) → Hn (S n ) → Hn (S n, S n \ {p}) → Hn−1 (S n \ {p}) → Hn−1 (S n ) .
| {z }



0 for n ≥ 2
=


 Z for n = 1

Hence in both cases n = 1 and n ≥ 2 the homomorphism (iii) is an isomorphism.

101
3 Homology Theory

Definition 3.34. The number deg p ( f ) is called the local degree of f over p.

Examples 3.35. 1.) If p < im( f ) then deg p ( f ) = 0 because Hn (S n , S n ) = 0.

2.) If f : U → S n is the inclusion map then deg p ( f ) = 1 for all p ∈ U. Namely, in this case the
homomorphisms (i) and (iii) in (3.6) coincide and Hn ( f ) is the inverse of (ii).

3.) For a homeomorphism f : U → f (U ) ⊂ S n we have that deg p ( f ) = ±1 for all p ∈ f (U ).


Namely, the homomorphisms (i) and Hn ( f ) in (3.6) are isomorphisms in this case, hence
multiplication by deg p ( f ) is an isomorphism, thus deg p ( f ) = ±1.

Proposition 3.36. Assume that f −1 (p) ⊂ K ⊂ V ⊂ U where K is compact and V open. Then
the degree deg p ( f ) is given by:


Hn (S n ) / Hn (S n , S n \ K ) o Hn (V, V \ K )
deg p ( f ) H n ( f |V )
 
Hn (S n ) 
/ Hn (S n, S n \ {p})

Hence we can replace f −1 (p) by a larger compact set in U and also U by a smaller open
neighborhood of f −1 (p). For this reason we call deg p ( f ) local.

Proof. The assertion follows from the commutativity of the following diagram:

Hn (S n6 , S n O \ f −1 (p)) o Hn (U, U \O f −1❙(p))
♠♠ ❙❙❙❙
♠♠♠♠ ❙❙H❙n❙( f )
♠ ❙❙❙❙
♠♠♠
♠♠♠ ❙❙)

Hn (S n ) ◗ Hn (S
5
n , S n \ {p}) o Hn (S n )
◗◗◗ ❦❦
◗◗◗
◗◗◗ ❦❦❦ ❦❦❦❦
◗◗◗ ❦❦
( 
❦❦❦❦ H n ( f |V )
Hn (S n , S n \ K ) o Hn (V, V \ K )

where all but two arrows are induced by inclusions. 

Remark 3.37. For f : S n → S n we have that deg( f ) = deg p ( f ) for all p ∈ S n . Simply choose
K = V = S n in Proposition 3.36.

102
3.4 The degree of a continuous map

Lemma 3.38. Let f : (D+n, S n−1 ) → (D+n, S n−1 ) be continuous and p ∈ D̊+n such that f −1 (p) ⊂
D̊+n is compact. Then  
deg( f ) = deg p f | D̊+n .

Proof. We extend f to a continuous map F : (S n , D−n ) → (S n, D−n ) by mapping the circular arc
from a point x on the equator S n−1 to the south pole to the corresponding arc from f (x) to the
south pole. Then by Remark 3.37 and Proposition 3.36 with K = f −1 (p), V = D̊+n and U = S n
we find
deg(F) = deg p (F) = deg p (F | D̊+n ) = deg p ( f | D̊+n ) . (3.7)
On the other hand, the commutative diagram

(D+n, S n−1 ) / (S n, D n ) o
− Sn
f F F
  
(D+n, S n−1 ) / (S n, D n ) o
− Sn

yields on the level of homology


 
Hn (D+n, S n−1 ) / Hn (S n , D n ) o
− Hn (S n )
· deg( f ) H n (F) · deg(F)
  
 
Hn (D+n, S n−1 ) / Hn (S n , D n ) o
− Hn (S n )

Hence deg(F) = deg( f ). This together with (3.7) concludes the proof. 

Proposition 3.39. Let f : U ⊂ S n → S n be continuous, let p ∈ S n with f −1 (p) compact and


let g ∈ C (S n, S n ). Then
deg p ( f ◦ g) = deg p ( f ) · deg(g).

Proof. This follows from the commutative diagram:

Hn (S n / Hn (S n , S n \ f −1 (p))  Hn (U, U \ f −1 (p))


O )
deg p ( f ) Hn ( f )
- 
deg(g)=H n (g) 1 Hn (S n , S n \ {p})  Hn (S n )
O
deg p ( f ◦g)
H n ( f ◦g)

Hn (S n ) / Hn (S n , S n \ ( f ◦ g) −1 (p))  Hn (g −1 (U ), g −1 (U ) \ ( f ◦ g) −1 (p))

103
3 Homology Theory

Remark 3.40. If X i are the path-components of X then


M
Hm (X )  Hm (X i ),
i
see Exercise 3.1. The isomorphism is induced by the inclusion maps of the connected components
into X . Similarly one sees that for A ⊂ X and Ai = A ∩ X i
M
Hm (X, A)  Hm (X i , Ai ) .
i

Proposition 3.41 (Additivity of the degree). Let f : U → S n be a continuous map. Let


p ∈ S n be such that f −1 (p) is compact. Let U = ∪rλ=1 Uλ with Uλ open and put f λ := f |Uλ .
Suppose f λ−1 (p) are pairwise disjoint. Then
r
X
deg p ( f ) = deg p ( f λ ) .
λ=1

Before proving the proposition we give some examples.

Example 3.42 U
Assume that f −1 (p) is a finite set, i.e.
b

f −1 (p) = {p1, . . . , pr }. Now choose Uλ such


that pλ ∈ Uλ and pν < Uλ for ν , λ. It
follows that
r
X b

deg p ( f ) = deg p ( f λ ) .
λ=1

If the map f is a local homeomorphism, then


deg p ( f ) = ±1 for all λ. b

Example 3.43. Consider the map f k : S1 → S1 with f (z) = z k , k > 0, and set p = 1. We write
f k−1 (1) = {k-th unit roots} = {ξ1, . . . , ξ k }
and find that f k |small neighborhood of ξ λ is a homeomorphism. Hence deg( f k, λ ) = ±1. Since k > 0,
the restriction of f k to a small neighborhood of ξ λ is homotopic to an embedding of a (k times
larger) neighborhood of ξ λ into S1 . Hence deg( f k, λ ) = 1. We conclude deg1 ( f k ) = k.

104
3.5 Homological algebra

Proof of Proposition 3.41. Choose open neighborhoods Vλ such that f λ−1 (p) ⊂ Vλ ⊂ Uλ with
Vλ ∩ Vµ = ∅ for λ , µ. Now put V = ∪rλ=1 Vλ . Proposition 3.36 tells us deg p ( f ) = deg p ( f |V ).
The commutative diagram
deg p ( f |V )

H n ( f |V ) ,
Hn (S n ) / Hn (V, V \ f −1 (p)) / Hn (S n, S n \ {p}) o Hn (S n
O O  O )
*.1+/
.. .. //
.. //
.  (1, ...,1) (1, ...,1)

,1-
Lr  Lr ⊕λ Hn ( f λ ) Lr Lr

λ=1 Hn (S n ) /
λ=1 Hn (Vλ, Vλ \ f λ−1 (p)) /
λ=1 Hn (S n, S n \ {p}) o 2 λ=1 Hn (S n )

*.deg p ( f 1 ) +/
.. .. //
.. . //
. /
, deg p ( f r ) -

yields

* +/ * . +
deg p ( f 1 ) 1
deg p ( f |V ) = (1, . . . , 1) · ... ..
. .
// · . .. // = deg p ( f 1 ) + . . . + deg p ( f r ). 
, deg p ( f r ) - ,1-

3.5 Homological algebra


Before continuing with topological considerations we clarify some of the underlying algebra.
Throughout this section let R be a commutative ring with unit element.

Definition 3.44. A complex of R-modules K∗ is a sequence

∂n+1 ∂n
... / K n+1 / Kn / K n−1 / ...

of R-modules K n together with homomorphisms ∂n such that

∂n ◦ ∂n+1 = 0 (3.8)

for all n ∈ Z. We define the space of n-cycles

Z n K∗ := {x ∈ K n | ∂n x = 0} = ker(∂n ) ,

and the space of n-boundaries

Bn K∗ := {∂n+1 x ∈ K n+1 | x ∈ K n+1 } = im(∂n+1 ).

105
3 Homology Theory

By (3.8) Bn K ⊂ Z n K so that we can define the n-th homology by


Z n K∗
Hn K∗ := .
Bn K∗

Example 3.45. For


 Sn (X, A; R) = Sn (X ; R)/Sn (A; R),
Kn = 
n≥0
 0,
 n<0

we obtain relative singular homology Hn K∗ = Hn (X, A; R).

Definition 3.46. Given two complexes K∗ and K∗′ a chain map ϕ∗ : K∗ → K∗′ is a sequence of
homomorphisms ϕ n : K n → K n′ such that
ϕ n+1
K n+1 / K′
n+1
∂n+1 ∂′n+1
 ϕn 
Kn / K′
n

commutes for all n ∈ Z.

Example 3.47. For a continuous map f : (X, A) → (Y, B) the homomorphisms ϕ n = Sn ( f )


constitute a chain map.

Given a general chain map ϕ∗ we conclude from



ϕ n ◦ ∂n+1 = ∂n+1 ◦ ϕ n+1

that ϕ n (Z n K∗ ) ⊂ Z n K∗′ and also that ϕ n (Bn K∗ ) ⊂ Bn K∗′ . Hence ϕ∗ induces homomorphisms
Hn (ϕ∗ ) : Hn K∗ → Hn K∗′ by Hn (ϕ∗ )([z]) = [ϕ n z]. This construction is functorial in the sense
that
Hn (ϕ∗ ◦ ψ∗ ) = Hn (ϕ∗ ) ◦ Hn (ψ∗ ) and Hn (idK∗ ) = idH n K∗ .

ϕ∗ ψ∗
Definition 3.48. A sequence · · · −→ K∗′ −→ K∗ −→ K∗′′ −→ · · · of chain maps is called exact
ϕn ψn
iff the sequence · · · −→ K n′ −→ K n −→ K n′′ −→ · · · is exact for every n ∈ Z.

106
3.5 Homological algebra

i∗ p∗
Proposition 3.49. If 0 → K∗′ −→ K∗ −→ K∗′′ → 0 is an exact sequence of complexes then the
H n (i) H n (p)
sequence Hn K∗′ −→ Hn K∗ −→ Hn K∗′′ is exact for every n ∈ Z.

Proof. a) By assumption we have that p∗ ◦ i∗ = 0 and hence


0 = Hn (p∗ ◦ i∗ ) = Hn (p∗ ) ◦ Hn (i∗ ).
Therefore im Hn (i ∗ ) ⊂ ker Hn (p∗ ).
b) It remains to show that ker Hn (p) ⊂ im Hn (i).
Let z ∈ Z n K∗ represent [z] ∈ ker Hn (p∗ ). Hence pn z = ∂ ′′ x ′′ for some x ′′ ∈ K n+1
′′ . Since

pn+1 is surjective we can choose x ∈ K n+1 with pn+1 x = x ′′. We compute


pn (z − ∂ x) = ∂ ′′ x ′′ − ∂ ′′ pn+1 x = ∂ ′′ x ′′ − ∂ ′′ x ′′ = 0 .
By exactness of the complex there exists y ′ ∈ K n′ such that z − ∂ x = i n y ′. Now we get
i n−1 ∂ ′ y ′ = ∂i n y ′ = ∂[z − ∂ x] = ∂z − ∂∂ x = 0 .
Since i n−1 is injective it follows that ∂ ′ y ′ = 0, i.e., y ′ ∈ Z n K∗′ represents an element in
homology. Finally we see
Hn (i∗ )([y ′]) = [i n y ′] = [z − ∂ x] = [z].
This shows ker Hn (p∗ ) ⊂ im Hn (i ∗ ). 

i∗ p∗
Definition 3.50. Let 0 → K∗′ −→ K∗ −→ K∗′′ → 0 be an exact sequence of complexes. We
construct the connecting homomorphism

∂∗ : Hn K∗′′ → Hn−1 K∗′

as follows: Let z ′′ ∈ Z n K∗′′ represent an element [z ′′] ∈ Hn K∗′′. Since pn is surjective we can
choose x ∈ K n with pn x = z ′′. For ∂ x ∈ K n−1 we observe

pn−1 ∂ x = ∂ ′′ pn x = ∂ ′′ z ′′ = 0.

By exactness there is a unique y ′ ∈ K n−1


′ with i n−1 y ′ = ∂ x. Moreover,

i n−2 ∂ ′ y ′ = ∂i n−1 y ′ = ∂∂ x = 0.

Since i n−2 is injective we have ∂ ′ y ′ = 0, i.e., y ′ ∈ Z n−1 K∗′ . Now put

∂∗ [z ′′] := [y ′].

107
3 Homology Theory

Lemma 3.51. The conncecting homomorphism ∂∗ : Hn K∗′′ → Hn−1 K∗′ is well defined, i.e.,
independent of the choices made in its construction.

Proof. There are two choices in the construction of the connecting homomorphism: that of the
preimage x with pn x = z ′′ and that of the representing cycle z ′′ itself.
As to the choice of x let pn x = pn x̄ = z ′′. For the corresponding elements y ′, ȳ ′ ∈ K n−1
′ we
have and i n−1 y = ∂ x and i n−1 ȳ = ∂ x̄. Since pn (x − x̄) = 0 there exists an ω ∈ K n′ with
′ ′ ′

x − x̄ = i n (ω ′ ). We compute
i n−1 (y ′ − ȳ ′ ) = ∂(x − x̄) = ∂(i n ω ′ ) = i n−1 ∂ω ′ ,
from which we conclude that y − ȳ ′ = ∂ω ′ and therefore [y ′ ] = [ ȳ ′ ].
As to the choice of the representing cycle z ′′ it suffices to show that if z ′′ is a boundary then
[y ′ ] = 0. Let z ′′ = ∂ ′′ ζ ′′ be a boundary. Since pn+1 is onto we can choose ξ ∈ K n+1 with
pn+1 ξ = ζ. Then x = ∂ξ is an admissible choice because
pn x = pn ∂ξ = ∂ ′′ pn+1 ξ = ∂ ′′ ζ = z ′′ .
Thus ∂ x = ∂∂ξ = 0 and hence y ′ = 0. 

It is easy to see that the connecting homomorphism is indeed a homomorphism. Now we are
ready to prove the Exactness Axiom.

i∗ p∗
Proposition 3.52. If 0 → K∗′ −→ K∗ −→ K∗′′ → 0 is an exact sequence of complexes then the
long homology sequence

∂∗ ∂∗
··· / Hn+1 K ′′ / Hn K ′ / Hn K∗ H n (p∗ )/ Hn K ′′
H n (i ∗ )
/ Hn−1 K ′ / ···
∗ ∗ ∗ ∗

is also exact.

Proof. In view of Proposition 3.49 it remains to show ker H (i) = im ∂∗ and ker ∂∗ = im H (p).
a) im ∂∗ ⊂ ker H (i):
Using the notation of Definition 3.50 we compute
H (i)∂∗ [z ′′] = H (i)∂∗ [px] = H (i)[i −1 ∂ x] = [∂ x] = 0 .

b) ker H (i) ⊂ im ∂∗ :
Let H (i)[z ′] = 0. Then we have [iz ′ ] = 0 and therefore iz ′ = ∂ x. Put z ′′ := px. Then
∂ ′′ z ′′ = ∂ ′′ px = p∂ x = piz ′ = 0. Hence z ′′ ∈ Z n K∗ represents an element in homology. We
compute
∂∗ [z ′′] = ∂∗ [px] = [i−1 ∂ x] = [z ′] .
Hence [z ′] ∈ im ∂∗ .

108
3.5 Homological algebra

c) im H (p) ⊂ ker ∂∗ :
This follows from
∂∗ H (p)[z] = ∂∗ [pz] = [i−1 |{z}
∂z ] = 0 .
=0

d) ker ∂∗ ⊂ im H (p):
Let ∂∗ [z ′′] = 0. We write z ′′ = px and compute

0 = ∂∗ [z ′′] = ∂∗ [px] = [i−1 ∂ x] .

This implies that i −1 ∂ x is a boundary in K ′. Hence we have that i −1 ∂ x = ∂ ′ x ′ and therefore

∂(x − ix ′ ) = ∂ x − i∂ ′ x ′ = 0 .

This finally leads to

H (p)[x − ix ′ ] = [px − pix ′ ] = [px] = [z ′′] ,

hence [z ′′] ∈ im H (p). 

Example 3.53. For K n′ = Sn (A), K n = Sn (X ) and


Sn (X )
K n′′ = Sn (X, A) =
Sn (A)
Proposition 3.52 now yields the Exactness Axiom for singular homology.

Example 3.54. Consider a triple of spaces (X, A, B) with B ⊂ A ⊂ X and set

K n′ = Sn (A)/Sn (B) = Sn (A, B)


Kn = Sn (X )/Sn (B) = Sn (X, B)
K n′′ = Sn (X )/Sn (A) = Sn (X, A) .

Then the canonical sequence

0 / K′ / Kn / K ′′ /0
n n

is again exact. By Proposition 3.52 we obtain the long exact homology sequence for a triple:

... /H (X, A)
❣ ❣❣❣ n+1
∂∗ ❣❣❣❣ ❣
❣❣
❣❣❣❣❣
s❣❣❣❣❣
Hn (A, B) / Hn (X, B) / Hn (X, A)
❣❣ ❣❣❣
∂∗ ❣❣❣❣❣

❣❣❣ ❣❣❣❣❣
s❣❣
Hn−1 (A, B) / ...

109
3 Homology Theory

Proposition 3.55. The long homology sequence is natural, i.e., if the diagram of chain maps

i∗ p∗
0 / K′ / K∗ / K ′′ /0
∗ ∗

ϕ ∗′ ϕ∗ ϕ ∗′′
 j∗  q∗ 
0 / L′ / L∗ / L ′′ /0
∗ ∗

is commutative with exact rows then the diagram

H n+1 (p∗ ) ∂∗ ∂∗
... / Hn+1 K ′′ / Hn K ′ / Hn K∗ H n (p∗ )/ Hn K ′′
H n (i ∗ )
/ ...
∗ ∗ ∗

H n+1 (ϕ ∗′′ ) H n (ϕ ∗′ ) H n (ϕ ∗ ) H n (ϕ ∗′′ )


H n+1 (q∗ )  ∂∗  H n ( j∗ )  H n (q∗ )  ∂∗
... / Hn+1 L ′′ / Hn L ′ / Hn L ∗ / Hn L ′′ / ...
∗ ∗ ∗

is commutative as well.

Proof. By assumption we have ϕ ◦ i = j ◦ ϕ ′ and ϕ ′′ ◦ p = q ◦ ϕ, hence

Hn (ϕ) ◦ Hn (i) = Hn ( j) ◦ Hn (ϕ ′ ),
Hn (ϕ ′′ ) ◦ Hn (p) = Hn (q) ◦ Hn (ϕ).

We are left to show that the diagram


∂∗
Hn K ′′ / Hn−1 K ′

H n (ϕ ′′ ) H n−1 (ϕ ′ )
 ∂∗ 
Hn L ′′ / Hn L ′

commutes. We calculate

Hn−1 (ϕ ′ )∂∗ [px] = Hn−1 (ϕ ′ )[i−1 ∂ x] = [ϕ ′i −1 ∂ x] = [ j −1 ϕ∂ x]


= [ j −1 ∂ϕx] = ∂∗ [pϕx] = ∂∗ [ϕ ′′ px] = ∂∗ Hn (ϕ ′′ )[px] ,

which proves the assertion. 

Example 3.56. Let f ∈ C ((X, A), (Y, B)). For K n′ = Sn (A), K n = Sn (X ), K n′′ = Sn (X, A)
and ϕ n′ = Sn ( f | A ), ϕ n = Sn ( f ), and ϕ n′′ the induced homomorphism on Sn (X, A) we recover
Lemma 3.13.

Definition 3.57. Let ϕ, ψ : K → K ′ be chain maps. A chain homotopy between ϕ and ψ is a


sequence of homomorphisms

hn : K n → K n+1

110
3.5 Homological algebra

such that

ϕ n − ψ n = ∂n+1 hn + hn−1 ∂n ,
for all n ∈ Z. The maps ϕ and ψ are called homotopic if such a homotopy exists. In this case
we write ϕ ≃ ψ.

Proposition 3.58. (i) The relation “≃” is an equivalence relation on the set of all chain
maps K → K ′.

(ii) If ϕ ≃ ψ : K → K ′ and ϕ ′ ≃ ψ ′ : K ′ → K ′′ then ϕ ′ ϕ ≃ ψ ′ ψ : K → K ′′.

Proof. (i) First we show that “≃” is an equivalence relation.


a) To show that ϕ ≃ ϕ choose h = 0.
b) The statement ϕ ≃ ψ ⇒ ψ ≃ ϕ follows from replacing h by −h.
c) If ϕ ≃ ψ and ψ ≃ χ then ψ ≃ χ .
h k h+k

(ii) Assume that ϕ ≃ ψ and ϕ ′ ≃′ ψ ′. Now we calculate


h h

ϕ ′ ϕ − ϕ ′ψ = ϕ ′ (∂ ′ h + h∂) = ∂ϕ ′ h + ϕ ′ h∂ ,

hence ϕ ′ ϕ ≃′ ϕ ′ψ. Similarly we see


ϕh

ϕ ′ψ − ψ ′ ψ = (∂ ′ h ′ + h ′ ∂)ψ = ∂ ′ h ′ψ + h ′ψ∂ ,

hence ϕ ′ψ ≃′ ψ ′ ψ. Combining both equivalences we get the desired result, namely


ϕ ′ ϕ ≃ ϕ ′ψ ≃ ψ ′ ψ .

Homotopic chain maps induce the same homomorphisms on homology:

Proposition 3.59. If ϕ ≃ ψ : K → K ′ then H (ϕ) = H (ψ) : H K → H K ′.

Proof. Let z ∈ Z n K. We compute

H (ϕ)[z] = [ϕz] = [ψz + ∂ ′ hz + h∂z] = [ψz] = H (ψ)[z]

because ∂z = 0 and [∂ ′ hz] = 0. 

111
3 Homology Theory

Definition 3.60. A chain map ϕ : K → K ′ is called a homotopy equivalence if there exists a


chain map ψ : K ′ → K such that ψϕ ≃ id K and ϕψ ≃ idK ′ . If such a homotopy equivalence
exists we write K ≃ K ′.

Corollary 3.61. If K ≃ K ′ then H K  H K ′.


ϕ H (ϕ)

3.6 Proof of the homotopy axiom


We put I := [0, 1] and define the affine linear map
j
Tn : ∆n+1 → ∆n × I

for j = 0, . . . , n by

 (ek , 0),
Tn (ek ) = 
j k ≤ j,
 (ek−1, 1),
 k > j.
In the case n = 0 we have
T00 : ∆1 → ∆0 × I = {e0 } × I ,
which can be visualized as

b
(e0, 0)
T00

b b b
(e0, 1)
e0 e1

In the case n = 1 we find

112
3.6 Proof of the homotopy axiom

e2 (e0, 1) (e1, 1)
b b b

T10

b b b b

e0 e1 (e0, 0) (e1, 0)

e2 (e0, 1) (e1, 1)
b b b

T11

b b b b

e0 e1 (e0, 0) (e1, 0)

Lemma 3.62. The composition of the operators T and the affine linear map F (defined in (3.1)
on page 80) yields:
j+1 i j
Tn ◦ Fn+1 = (Fni × id) ◦ Tn−1 ( j ≥ i),
j i+1 j
Tn ◦ Fn+1 = (Fni × id) ◦ Tn−1 ( j < i),
Tni ◦ Fn+1
i
= Tni−1 ◦ i
Fn+1 (1 ≤ i ≤ n),
Tn0 ◦ Fn+1
0
= i 1,
Tnn ◦ Fn+1
n+1
= i 0,

with i t : ∆n → ∆n × I being the inclusion map x 7→ (x, t).

Proof. Let us prove the first formula where j ≥ i. If k < i then


j+1 i j+1 j
Tn ◦ Fn+1 (ek ) = Tn (ek ) = (ek , 0) = (Fni × id)(ek , 0) = (Fni × id) ◦ Tn−1 (ek ).

If i ≤ k ≤ j then

Tnj+1 ◦ Fn+1
i j
(ek ) = Tnj+1 (ek+1 ) = (ek+1, 0) = (Fni × id)(ek , 0) = (Fni × id) ◦ Tn−1 (ek ).

If k > j then
j+1 i j+1 j
Tn ◦ Fn+1 (ek ) = Tn (ek+1 ) = (ek , 1) = (Fni × id)(ek−1, 1) = (Fni × id) ◦ Tn−1 (ek ).

The proofs of the other formulas are similar exercises in index shifting. 

113
3 Homology Theory

j
Now let σ : ∆n → X be a singular n-simplex. Note that (σ × id) ◦ Tn ∈ C (∆n+1, X × I). We
define Pσ ∈ Sn+1 (X × I) by
n
X
Pσ := (−1) j (σ × id) ◦ Tnj .
j=0

We extend P linearly to chains and get a linear map

P : Sn (X ) → Sn+1 (X × I).

This homomorphism descends to relative chains

P : Sn (X, A) → Sn+1 (X × I, A × I).

The operator P is called prism operator.

Lemma 3.63. Let (X, A) be a pair of spaces. The operator P is a chain homotopy for

S(i 0X ), S(i 1X ) : S(X, A) → S(X × I, A × I) ,

where i tX : X → X × I denotes the inclusion map x 7→ (x, t).

Proof. We have to show that


S(i 0X ) − S(i 1X ) = P∂ + ∂P. (3.9)
Let σ : ∆n → X be a singular n-simplex. We compute, using Lemma 3.62:
n
X
P∂σ = P (−1) i (σ ◦ Fni )
i=0
n
X n−1
X
j
= (−1) i (−1) j (σ ◦ Fni × id) ◦ Tn−1
i=0 j=0
X
j
= (−1) i+ j (σ × id) ◦ (Fni × id) ◦ Tn−1
0≤ j <i ≤n
X
j
+ (−1) i+ j (σ × id) ◦ (Fni × id) ◦ Tn−1
0≤i ≤ j ≤n−1
X
= − (−1) i+ j+1 (σ × id) ◦ Tnj ◦ Fn+1
i+1

0≤ j <i ≤n
X
j+1
− (−1) i+ j+1 (σ × id) ◦ Tn i
◦ Fn+1 .
0≤i ≤ j ≤n−1

114
3.6 Proof of the homotopy axiom

On the other hand, we find


n
X
j
∂Pσ = ∂ (−1) j (σ × id) ◦ Tn
j=0
n
X n+1
X
j
= (−1) j (−1) i (σ × id) ◦ Tn ◦ Fn+1
i

j=0 i=0
X
j
= (−1) i+ j (σ × id) ◦ Tn ◦ Fn+1
i
(i < j)
0≤i< j ≤n
Xn
+ (σ × id) ◦ Tni ◦ Fn+1
i
(i = j)
i=0
n+1
X
− (σ × id) ◦ Tni−1 ◦ Fn+1
i
(i = j + 1)
i=1
X
+ (−1) i+ j (σ × id) ◦ Tnj ◦ Fn+1
i
(i > j + 1)
1≤ j+1<i ≤n+1
X ′ ′
= (−1) i+ j +1 (σ × id) ◦ Tnj +1 ◦ Fn+1
i

0≤i ≤ j ′ ≤n−1
× id) ◦ i1 − (σ × id) ◦ i0
+(σ X
′ j ′
+ (−1) i + j+1 (σ × id) ◦ Tn ◦ Fn+1
i +1

0≤ j <i ′ ≤n

In the last step we changed the summation indices to j ′ = j − 1 and i ′ = i − 1. We see that
P∂σ + ∂Pσ = (σ × id) ◦ i1 − (σ × id) ◦ i0
= i 1X ◦ σ − i 0X ◦ σ
= Sn (i 1X )σ − Sn (i 0X )σ
which proves the lemma. 

Proposition 3.64. If f ≃ g : (X, A) → (Y, B) then we have

S( f ) ≃ S(g) : S(X, A) → S(Y, B)

Proof. Let F be a homotopy for f and g, i.e., F ∈ C ((X × I, A × I), (Y, B)) with
f = F ◦ i 1X , g = F ◦ i0X .
Then by (3.9) we get
S( f ) − S(g) = S(F)S(i 1X ) − S(F)S(i 0X ) = S(F)P∂ + S(F)∂P = S(F)P∂ + ∂S(F)P.
Hence S(F)P is a chain homotopy for S( f ) and S(g). 

115
3 Homology Theory

Corollary 3.65. The homotopy axiom holds for singular homology.

3.7 Proof of the excision axiom


We want to show that the inclusion (X \ U, A \ U ) ֒→ (X, A) induces an isomorphism

Hn (X \ U, A \ U )  Hn (X, A)

provided Ū ⊂ Å.
X

A U

Let X be a topological space and let U = {Ui }i ∈I be an open cover of X . A chain


P
σ = j α j σ j ∈ Sn (X ) is called U -small iff for each j there exists an i such that σ j (∆n ) ⊂ Ui .
We denote

SnU (X ) := {σ ∈ Sn (X ) | σ is U -small}

= im * Sn (Ui ) −−−−−−−−−→ Sn (X ) +
M ⊕i S n ( j i )

, i ∈I -
with ji : Ui → X the inclusion map. For A ⊂ X we put

SnU (X )
SnU (X, A) := .
SnU (A)

Theorem 3.66 (Small chain theorem). The inclusion SnU (X, A) → Sn (X, A) induces an iso-
morphism in homology.

Before coming to the proof we show that the small chain theorem implies the excision axiom.
To this extent let (X, A) be a pair of spaces and let U ⊂ A be such that Ū ⊂ Å. Now set U1 := Å

116
3.7 Proof of the excision axiom

and U2 := X \ Ū. Then U = {U1, U2 } forms an open cover of the space X . We compute

SnU (X )
SnU (X, A) =
SnU (A)
Sn ( Å) + Sn (X \ Ū)
=
Sn ( Å) + Sn (A \ Ū)
Sn (X \ Ū )
=
(Sn ( Å) + Sn (A \ Ū)) ∩ Sn (X \ Ū)
Sn (X \ Ū)
= .
Sn (A \ Ū)
Similarly we get

Sn ( Å \ U ) + Sn (X \ Ū) Sn (X \ Ū)
SnU (X \ U, A \ U ) = = .
Sn ( Å \ U ) + Sn (A \ Ū) Sn (A \ Ū)
In the following commutative diagram all arrows are induced by inclusions.

U (X \ U, A \ U ) = / S U (X, A)
Sm n

 
Sm (X \ U, A \ U ) / Sn (X, A)

By the small chain theorem the vertical arrows induce isomorphisms on homology and the
excision theorem is proved.
It remains to prove Theorem 3.66. We define the barycenter of ∆n by
n
1 X
Bn := ej .
n + 1 j=0

Example 3.67. For example, e2

B0 = e0,
1
B1 = (e0 + e1 ), B1
b

B2
2
1 e0 e1 e0 e1
B2 = (e0 + e1 + e2 ).
3

For an affine map σ : ∆n → ∆n+1 we define the affine map Cn σ : ∆n+1 → ∆n+1 by


 Bn+1,
(Cn σ)(ek ) = 
k = 0,
 σ(ek−1 ),
 k ≥ 1.

117
3 Homology Theory

Example 3.68. Consider the case σ = F20 : ∆1 → ∆2 .

e2
b

G
b

e0 e1

e0 e1 C1 G
e2
b

b b

e0 e1

Now we set

Skaff (∆n ) := σ ∈ Sk (∆n ) σ =


 X
α j σ j and each σ j is affine .
j

By linear extension we get a homomorphism

Cn : Snaff (∆n+1 ) → Sn+1


aff
(∆n+1 ) .

Lemma 3.69. The homomorphism Cn has the following properties:


P P
(i) ∂C0 (c) = c − ( j α j )B0 where c = j α j σ j ;

(ii) ∂Cn (c) = c − Cn−1 ∂c for n ≥ 1.

Proof. (i) It suffices to show the assertion for an affine simplex c. We then find

(C0 c)(e0 ) = B0, (C0 c)(e1 ) = c(e0 )

and hence

∂C0 (c)(e0 ) = (C0 (c))(e1 ) − (C0 (c))(e0 ) = c(e0 ) − B0 = (c − 1B0 )(e0 )

118
3.7 Proof of the excision axiom

as desired.

(ii) is left as an exercise. 

Lemma 3.70. To each topological space X and each n ∈ N0 we can associate homomorphisms

Sdn : Sn (X ) → Sn (X ),
Q n : Sn (X ) → Sn+1 (X ),

such that

(i) Sd∗ is a chain map, i.e., ∂ ◦ Sdn = Sdn−1 ◦ ∂;

(ii) Q∗ is a chain homotopy between id and Sd∗ , i.e., id −Sdn = ∂ ◦ Q n + Q n−1 ◦ ∂;

(iii) Sd∗ and Q∗ are natural, i.e., for every f ∈ C (X, Y ) the following diagrams commute:

Sd n Qn
Sn (X ) / Sn (X ) Sn (X ) / Sn+1 (X )

Sn ( f ) Sn ( f ) Sn ( f ) S n+1 ( f )
 Sd n   Qn 
Sn (Y ) / Sn (Y ) Sn (Y ) / Sn+1 (Y )

(iv) If the map σ : ∆n → ∆n is affine then each simplex σ j occuring in Sdn (σ) or in Q n (σ)
is affine and
n
diam(σ j ) ≤ diam(σ).
n+1

Proof. The construction of Sdn and Q n will be done recursively, the proof is by induction over
n. We start by considering the case n = 0. We put Sd0 := id : S0 (X ) → S0 (X ) and Q0 := 0. It
is obvious that the four assertions hold.
In the case n ≥ 1 we assume that Sdn−1 and Q n−1 are already defined and we set for a singular
simplex σ : ∆n → X :

Sdn (σ) := Sn (σ)(Cn−1 (Sdn−1 (∂ (id∆ n ) ) ) )


|{z}
n∈S aff (∆ n )
| {z }
∈S aff (∆ n )
| {zn−1 }
∈S aff
n−1
(∆ n )
| {z }
∈S aff n
n (∆ )

119
3 Homology Theory

and
Q n (σ) = Sn+1 (σ)(Cn (id∆ n −Sdn (id∆ n ) − Q n−1 ∂ id∆ n ) ).
| {z }
∈S aff n
n (∆ )
| {z }
∈S aff
n+1
(∆ n )

We have to verify assertions (i)-(iv). We check (iii):

Sdn (Sn ( f )(σ)) = Sn (Sn ( f )(σ))(Cn−1 (Sdn−1 (∂(id∆ n ))))


= Sn ( f ◦ σ)(Cn−1 (Sdn−1 (∂(id∆ n ))))
= Sn ( f ) ◦ Sn (σ)(Cn−1 (Sdn−1 (∂(id∆ n ))))
= Sn ( f )(Sdn (σ)).

The computation for Q n is similar.


Next we check (i): First we consider the case that X = ∆n and σ = id∆ n .

∂(Sdn (id∆ n )) = ∂(Cn−1 (Sdn−1 (∂ id∆ n )))


L. 3.69
= Sdn−1 (∂ id∆ n ) − Cn−2 ∂(Sdn−1 (∂ id∆ n ))
ind. hyp.
= Sdn−1 (∂ id∆ n ) − Cn−2 Sdn−2 |{z}
∂∂ id∆ n
=0
= Sdn−1 (∂ id∆ n ).

For a general simplex σ : ∆n → X we then find

∂(Sdn (σ)) = ∂(Sn (σ)(Sdn (id∆ n )))


= Sn (σ)(∂(Sdn (id∆ n )))
= Sn (σ)(Sdn−1 (∂ id∆ n ))
(iii)
= Sdn−1 (Sn (σ)(∂(id∆ n )))
= Sdn−1 (∂Sn (σ)(id∆ n ))
= Sdn−1 (∂σ).

Now we check (ii): Again we first consider the case X = ∆n and σ = id∆ n .

∂Q(id∆ n ) = ∂Cn (id∆ n −Sdn (id∆ n ) − Q n−1 ∂ id∆ n )


L. 3.69
= id∆ n −Sdn (id∆ n ) − Q n−1 ∂ id∆ n
 
−Cn−1 ∂ id∆ n −Sdn (id∆ n ) − Q n−1 ∂ id∆ n
ind. hyp.
= id∆ n −Sdn (id∆ n ) − Q n−1 ∂ id∆ n

−Cn−1 ∂ id∆ n −∂Sdn (id∆ n ) − (∂ id∆ n −Sdn−1 (∂ id∆ n )

∂∂ id∆ n ))
+ Q n−2 (|{z}
=0
(i)
= id∆ n −Sdn (id∆ n ) − Q n−1 ∂ id∆ n .

120
3.7 Proof of the excision axiom

The passage to general σ can now be done as before.


Finally we check (iv): It is clear from the recursive definition of Sdn and of Q n that each simplex
σ j occuring in Sdn (σ) or in Q n (σ) is again affine. The diameter of an affine simplex is the
maximal distance of any two of its vertices. We distinguish two cases:

1. The vertices p, q of σ j of maximal distance lie on ∂σ.

p q

Then we find by the induction hypothesis


n−1 n
d(p, q) ≤ diam(face of σ) < diam(σ) .
n n+1

2. One vertex is Bn :

Bn
pi


Then we find

1 X
n
d(pi , Bn ) = pi −
n + 1 j=0
pj

1 X n

= (pi − p j )
n + 1 j=0
n
1 X
≤ |pi − p j |
n + 1 j=0 | {z }
≤diam(σ ) and =0 for j=i
n
≤ diam(σ) .
n+1

121
3 Homology Theory

Conclusion 3.71. We have that Sd ≃ id and therefore

Sd − Sd2 = Sd ◦ ∂ ◦ Q + Sd ◦ Q ◦ ∂
= ∂ ◦ Sd ◦ Q + Sd ◦ Q ◦ ∂.

We conclude that Sd2 ≃ Sd ≃ id. Iterating this procedure we find that Sdr ≃ id for all r ∈ N.
Hence there exist homomorphisms Q n(r ) : Sn (X ) → Sn+1 (X ) with

Sdrn − id = ∂ ◦ Q n(r ) + Q n−1


(r )
◦∂.

For σ affine, we have for every σ j occuring in Sdr (σ) that


 n r
diam(σ j ) ≤ diam(σ)
n+1

Lemma 3.72. For σ : ∆n → X continuous there exists a ε > 0 such that all ε-balls ∩∆n are
completely contained in σ −1 (Ui ) for some Ui ∈ U .

Proof. Assume the assertion were false. Then for ε k = 1/k there exists a point pk ∈ ∆n such
that
B 1 (pk ) = {x ∈ ∆n | |x − pk | < ε k }
k

is not contained in any of the σ −1 (Ui ). After passing to a subsequence we have that pk → p ∈ ∆n
by compactness of ∆n . Choose i0 with p ∈ σ −1 (Ui 0 ). Since σ −1 (Ui 0 ) is open there exists a δ > 0
such that Bδ (p) ⊂ σ −1 (Ui 0 ). Now choose k so large, that |pk − p| < δ/2 and ε k = k1 < δ/2. We
then find
B 1 (pk ) ⊂ Bδ (p) ⊂ σ −1 (Ui 0 ),
k

a contradiction. 

Corollary 3.73. Assume that σ : ∆n → X is continuous. Then there exists an r (σ) ∈ N such
that every simplex σ j occuring in Sdr (σ) or in Q (r ) (σ) for r ≥ r (σ) is completely contained
in one of the Ui , i.e., Sdr (σ), Q (r ) (σ) ∈ SnU (X ).

Finally we can prove the small chain theorem.

Proof of Theorem 3.66. a) First we consider the case A = ∅.


i) We show injectivity: Assume that z ∈ Z nU (X ) with Hn ( j)([z]H nU ) = 0. Then there

122
3.8 The Mayer-Vietoris sequence

exists an x ∈ Sn+1 (X ) with ∂ x = z. Now we calculate

∂ Sdr x
|{z} = Sdr ∂ x
∈S U
n+1
(X ) for large r

= Sdr z
= z − ∂Q (r ) z − Q (r ) |{z}
∂z .
=0

Hence
z = ∂( Sdr x + Q (r ) z )
| {z }
∈S U
n (X ) for large r

and therefore [z]H nU = 0.


ii) We show surjectivity: Let [z] ∈ Hn (X ) be given. We know that Sdr z ∈ SnU (X ) for r
large enough. We compute

[Sdrn z] = [z − ∂Q n(r ) z − Q n−1


(r )
∂z ] = [z].
|{z}
| {z }
∈H n ( j)(H nU (X )) =0

b) Now we pass to general (X, A). Consider the commutative diagram with exact rows:

HnU (A) / H U (X ) / H U (X, A) / H U (A) / H U (X )


n n n−1 n−1
   
    
Hn (A) / Hn (X ) / Hn (X, A) / Hn−1 (A) / Hn−1 (X )

By part a) of the proof we know that the outer four arrows are isomorphisms. The Five
Lemma (Exercise 3.11) implies that the map HnU (X, A) → Hn (X, A) is also an isomorphism.


3.8 The Mayer-Vietoris sequence


Let X be a topological space and let A ⊂ X be a subset. Assume that U1, U2 ⊂ X are open with
U1 ∪ U2 = X , hence U = {U1, U2 } forms an open cover of X . Consider the exact sequence of
chain complexes
 
S n (i 1 )
−S n (i 2 ) (S n ( j1 ), S n ( j2 ))
0 / Sn (U1 ∩ U2 ) / Sn (U1 ) ⊕ Sn (U2 ) / S U (X ) /0
n

with the inclusion maps iν : U1 ∩ U2 → Uν and jν : Uν → X . We then get the following long
exact homology sequence:
 
H n (i 1 )
−H n (i 2 ) ∂
· · · → Hn (U1 ∩ U2 ) / Hn (U1 ) ⊕ Hn (U2 ) (H n ( j1 ), H n ( j2 )) / H U (X ) → Hn−1 (U1 ∩ U2 ) → · · ·
n

123
3 Homology Theory

By the small chain theorem HnU (X ) is canonically isomorphic to Hn (X ). Using this isomorphism
we can replace HnU (X ) in the above exact homology sequence by Hn (X ).

(H n ( j1 ), H n ( j2 )) ∂
· · · → Hn (U1 ) ⊕ H❱n (U2 ) / H U (X ) / Hn−1 (U1 ∩ U2 ) → · · ·
❱❱❱❱ n 4
❱❱❱❱ ✐✐✐✐
❱❱❱❱  ✐✐✐✐✐✐✐
(H n ( j1 ), H n ( j2 ))❱❱❱❱❱❱ ✐✐✐✐ ∂ M V
❱*  ✐✐✐✐
Hn (X )

The same reasoning applies to relative homology and we obtain

Theorem 3.74 (Mayer-Vietoris sequence). Let X be a topological space, let A ⊂ X and let
U1, U2 ⊂ X be open such that U1 ∪ U2 = X . Set Aν := A ∩ Uν and let

iν :(U1 ∩ U2, A1 ∩ A2 ) → (Uν, Aν ),


jν :(Uν, Aν ) → (X, A),

be the inclusion maps, ν = 1, 2. Then the following sequence is exact and natural
 
H n (i 1 )
∂MV / Hn (U1 ∩ U2, A1 ∩ A2 ) −H n (i 2 )
/ Hn (U1, A1 ) ⊕ Hn (U2, A2 )
···
❝ ❝ ❝❝ ❝ ❝ ❝❝❝❝❝

❝❝❝❝❝❝❝❝
❝❝❝❝ ❝❝❝❝❝❝❝❝❝❝❝ (H n ( j1 ), H n ( j2 ))
q ❝❝❝❝❝❝

Hn (X, A) M V / Hn−1 (U1 ∩ U2, A1 ∩ A2 ) / ···

Example 3.75. We give a new computation of the homology of S1 . To this extent we cover the
circle as indicated in the picture.

U1 U2

We directly see that


U1 ≈ U2 ≈ (0, 1) ≃ {p}
and
U1 ∩ U2 ≈ (0, 1) ⊔ (0, 1) ≃ {p1, p2 }
Consider the following part of the Mayer-Vietoris sequence.

124
3.8 The Mayer-Vietoris sequence

Hn (U1 ∩ U2 ) = Hn ({p1, p2 }) / Hn (U1 ) ⊕ Hn (U2 ) = Hn ({p}) ⊕ Hn ({p}) / 1


❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝❝ Hn (S )
❝❝❝❝❝❝❝❝❝
❝❝❝ ❝❝❝❝❝❝❝❝❝❝❝❝❝❝
❝❝ ∂ M V
q❝❝❝❝❝❝❝❝❝❝❝
Hn−1 (U1 ∩ U2 ) = Hn−1 ({p1, p2 }) / Hn−1 (U1 ) ⊕ Hn−1 (U2 ) = Hn−1 ({p}) ⊕ Hn−1 ({p})

If n ≥ 2 then all homologies of the point occuring in this diagram vanish. Hence Hn (S1 ) = 0
for all n ≥ 2. Since S1 is path-connected we have that H0 (S1 )  R. In the case n = 1 we find
 
11
/ H1 (S 1 ) / R2 11 / R2
0
To compute H1 (S1 ) we calculate
−x
! ( ! )
1 1
x
1
H1 (S )  ker = x ∈ R  R.
1 1

Example 3.76. Now we consider the space X = G2 as given in the picture.

U1 U2

It is easy to see that U1 ≈ U2 ≃ S1 and U1 ∩ U2 ≃ {p}. Since G2 is path-connected we find


H0 (G2 )  R.
Hn (U1 ∩ U2 ) → Hn (U1 ) ⊕ Hn (U2 ) → Hn (G2 ) → Hn−1 (U1 ∩ U2 )
| {z } | {z }
H n (S 1 ) ⊕H n (S 1 ) H n−1 ( { p })

In the case of n ≥ 2 we find the exact sequence


0 / Hn (G2 ) /0

and hence Hn (G2 ) = 0. For n = 1 we find


H ({p}) → H1 (S1 ) ⊕ H1 (S1 ) → H1 (G2 ) → H0 ({p}) → H0 (S1 ) ⊕ H0 (S1 ) .
| 1{z } | {z } | {z } | {z }
=0 R 2 R R 2

The last map in this diagram is given by

H0 ({p})  R / H0 (S 1 ) ⊕ H0 (S 1 )  R2
  ;
1
1

and hence injective. Consequently we find the isomorphism


H1 (G2 )  H1 (U1 ) ⊕ H1 (U2 )  R2 .

125
3 Homology Theory

3.9 Generalized Jordan curve theorem


In this section we will use the Mayer-Vietoris sequence to prove the Jordan curve theorem in
arbitrary dimensions.

Lemma 3.77. Let X be a topological space and let Ui ⊂ X be open with Ui ⊂ Ui+1 and
∪i ∈N Ui = X . Furthermore let ιn : Un ֒→ X and ιn, m : Un ֒→ Um for m ≥ n be the inclusion
maps. Then we have:

(i) For each α ∈ Hk (X ; R) there exists an n0 such that α ∈ im(Hk (ιn )) for all n ≥ n0 .

(ii) For each α n ∈ Hk (Un ; R) with Hk (ιn )(α n ) = 0 there exists an m0 such that
Hk (ιn, m )(α n ) = 0 for all m ≥ m0 .

Proof. (i) Let α ∈ Hk (X ). We represent α by


l
X
α j σ j ∈ Zk (X ; R), α j ∈ R, σ j ∈ C (∆k , X ).
j=1
S
Note that σ j (∆k ) ⊂ X is a compact subset and therefore C := lj=1 σ j (∆k ) ⊂ X is
compact. Hence there exists an n0 such that for all n ≥ n0 we have C ⊂ Un . We conclude
that
X l
α j σ j ∈ Zk (Un ; R)
j=1
for all n ≥ n0 and therefore
fX g  fX g 
α= αjσj = Hn (ιn ) αjσj .
H k (X ) H k (U n )

P
(ii) Again represent α n ∈ Hk (Un ) by lj=1 α j σ j . From Hn (ιn )(α n ) = 0 we know that there
P
exists a β ∈ Ck+1 (X ; R) such that lj=1 α j σ j = ∂ β. As before there exists a compact
subset C ′ ⊂ X such that β ∈ Ck+1 (C ′; R). Since there exists an m0 with C ′ ⊂ Um for all
m ≥ m0 and thus β ∈ Ck+1 (Um ; R) we have that Hk (ιn, m )(α n ) = 0. 

By an embedding we mean a continuous map f : X → Y which is a homeomorphism onto its


image. In other words, f is continuous, open and injective. In the considerations which follow
the domain will be compact and the target will be Hausdorff so that f is automatically open.

Proposition 3.78. Let n ∈ N and let Y be a compact topological space such that for any
embedding f : Y → S n
H∗ (S n \ f (Y ); R)  H∗ (point; R) .

126
3.9 Generalized Jordan curve theorem

Then the space [0, 1] × Y also has this property.

Proof. Let f : [0, 1] × Y → S n be an embedding. Suppose 0 , α ∈ Hi (S n \ f ([0, 1] × Y )). Put


U0 := S n \ f ([0, 12 ] × Y ) and U1 := S n \ f ([ 12 , 1] × Y ) .
| {z } | {z }
compact compact

Both U0 and U1 are open. We also find


U0 ∩ U1 = S n \ f ([0, 1] × Y ) and U0 ∪ U1 = S n \ f ({ 21 } × Y ) .
| {z }
≈Y

The Mayer-Vietoris sequence yields:

0 = Hi+1 (S n \ f ({ 12 } × Y )) / Hi (S n \ f ([0, 1] × Y ))


Hi (S n \ f ([0, 21 ] × Y )) ⊕ Hi (S n \ f ([ 21 , 1] × Y ))

Thus the inclusions S n \ f ([0, 1]×Y ) ֒→ S n \ f ([0, 12 ]×Y ) and S n \ f ([0, 1]×Y ) ֒→ S n \ f ([ 12 , 1]×Y )
induce an injective homomorphism
Hi (S n \ f ([0, 1] × Y )) → Hi (S n \ f ([0, 12 ] × Y ) ⊕ Hi (S n \ f ([ 21 , 1] × Y )) .
Hence
0 , Hi (inclusion)(α) ∈ Hi (S n \ f ([0, 21 ] × Y )) or
0 , Hi (inclusion)(α) ∈ Hi (S n \ f ([ 21 , 1] × Y )) .
By iterating this procedure we obtain a sequence of intervals Ik such that
I0 = [0, 1], Ik+1 ⊂ Ik , |Ik | = 2−k
and
0 , Hi (ι0, k )(α) ∈ Hi (S n \ f (Ik × Y ))
for all k. Here Vk := S n \ f (Ik × Y ) is open in S n and ιk,l : Vk ֒→ Vl for l ≥ k denotes the
inclusion map. We find
∩k ∈N Ik = {t}
∪k ∈N Vk = S n \ f ({t} × Y ) =: X
Now we apply the previous Lemma 3.77 for the inclusion ι : V0 ֒→ X . Hence, for i ≥ 1,
0 , Hi (ι)(α) ∈ Hi (X ) = Hi (S n \ f ({t} × Y ) ) = 0
| {z }
≈Y

giving a contradiction. The proof for i = 0 is similar (or in fact the same if one uses augmented
homology). 

127
3 Homology Theory

Corollary 3.79. If f : D r → S n is an embedding then

H∗ (S n \ f (D r ))  H∗ ({point})

Proof. The proof is by induction on r. For r = 0 we find

S n \ f (D 0 ) ≈ Rn ≃ {point}

and hence H∗ (S n \ f (D r ))  H∗ ({point}).


For the induction step “r − 1 ⇒ r” we observe D r ≈ W r = [0, 1] × W r −1 ≈ [0, 1] × D r −1 and
hence Proposition 3.78 applies. 

Theorem 3.80. Let r < n and let f : Sr → S n be an embedding. Then

H∗ (S n \ f (Sr ))  H∗ (S n−r −1 ).

Proof. Again the proof is done by induction on r. For r = 0 we find

S n \ f (S0 ) = S n \ {p, q} ≈ Rn \ {0} ≃ S n−1 .

For the induction step “r − 1 ⇒ r” we write Sr = D+r ∪ D−r . Then D+r ∩ D−r = Sr −1 . We put
U+ := S n \ f (D+r ) and U− := S n \ f (D−r ). Both sets U+, U− are open because f (D±r ) is compact.
In addition
U+ ∩ U− = S n \ f (Sr ) and U+ ∪ U− = S n \ f (Sr −1 ).
Now we look at the Mayer-Vietoris sequence and use Corollary 3.79:

H (U ) ⊕ H (U ) / Hi+1 (S n \ f (S r −1 )) / Hi (S n \ f (S r ))
| i+1 + {z i+1 −}
H i+1 (point) ⊕H i+1 (point)=0


H (U ) ⊕ H (U )
| i + {z i −}
 H i (point)⊕ H i (point)=0
if i ≥1

Thus for i ≥ 1 (and by similar reasoning also for i = 0) we get an isomorphism

Hi (S n \ f (Sr ))  Hi+1 (S n \ f (Sr −1 ))  Hi+1 (S n−r )  Hi (S n−r −1 ).

This proves the theorem. 

128
3.9 Generalized Jordan curve theorem

Theorem 3.81 (Jordan-Brouwer separation theorem). For any embedding f : S n−1 → S n


the complement S n \ f (S n−1 ) consists of exactly two path-components U and V . Both U and
V are open in S n and ∂U = ∂V = f (S n−1 ).

Proof. a) We know that H0 (S n \ f (S n−1 ))  H0 (S0 )  R2 , hence S n \ f (S n−1 ) has exactly two
path-components U and V .

b) Since f (S n−1 ) is compact the union U ∪ V = S n \ f (S n−1 ) is open and therefore both
connected components U and V are open.

c) Since U and V are open they contain no boundary point of U, thus ∂U ⊂ f (S n−1 ). Assume
that there exists p ∈ S n−1 with f (p) < ∂U. Then there exists W ⊂ S n open with f (p) ∈ W
and W ∩ U = ∅. Since f is continuous we can choose an open ball B ⊂ S n−1 with f (B) ⊂ W .
Since S n−1 \ B ≈ D n−1 we find that the space Y := S n \ f (S n−1 \ B) is path-connected because
of Corollary 3.79:

H0 (Y ) = H0 (S n \ f (S n−1 \ B))  H0 (point)  R .

Moreover, we have

Y = U ∪ V ∪ f (B) ⊂ U ∪ V ∪ W and U ∩ (V ∪ W ) = ∅

and hence
Y = (Y ∩ U ) ⊔ (Y ∩ (V ∪ W ))
| {z } | {z }
U ⊃V
can be written as a disjoint union of open non-empty subsets. This contradicts Y being
path-connected. 

Corollary 3.82 (Generalized Jordan curve theorem). For every embedding f : S n−1 → Rn
the complement Rn \ f (S n−1 ) consists of exactly two path components U and V . Both U and
V are open, U is bounded, V is unbounded and ∂U = ∂V = f (S n−1 ).

Proof. We use the stereographic projection to identify Rn with a subset of S n ,

S n = Rn ∪ {∞}, f : S n−1 → Rn ⊂ S n .

By the Jordan-Brouwer separation theorem the two path-components Ũ and Ṽ of S n \ f (S n−1 )


are both open and ∂Ũ = ∂Ṽ = f (S n−1 ). Let Ṽ be the component containing ∞. Then U = Ũ
and V = Ṽ \ {∞} are the two path-components of Rn \ f (S n−1 ). Clearly, V is unbounded and
∂U = ∂Ũ = ∂V = ∂Ṽ = f (S n−1 ). If U were unbounded then ∞ ∈ ∂U which is not the case.
Hence U is bounded. 

129
3 Homology Theory

Remark 3.83. Consider an embedding f : S n−1 → S n . If i ≥ 1 we find for the homology of the
components of the complement

H (U ⊔ V ) = Hi (U ) ⊕ Hi (V )  Hi (S0 ) = 0

and hence U and V have the same homology as a point. This makes us suspect that U, V ≈ D̊ n .
For n = 2 this is indeed true, but it is false for n ≥ 3. The Alexander horned sphere is an
example of an embedding of S2 into S3 where one component of the complement is not even
simply connected:

The red circle in the picture is a non-contractible loop giving rise to a non-trivial element in
the fundamental group. A very nice video illustrating this embedding of S2 can be found at
http://www.youtube.com/watch?v=d1Vjsm9pQlc.

3.10 CW-complexes
We now describe a type of topological spaces for which there is a particularly efficient way to
compute their homology. These space are obtained by gluing together balls of various dimensions.

Definition 3.84. A finite CW-complex is a pair (X, X) where X is a Hausdorff space,


`
X = n ∈N0 Xn , Xn ⊂ P(X ) and |X| < ∞ with the following properties:
`
(i) X = σ ∈X σ.

(ii) Set X n := ∪σ ∈Xm σ ⊂ X . For every σ ∈ Xn we have σ̄ \ σ ⊂ X n−1 .


m ≤n

130
3.10 CW-complexes

(iii) For every σ ∈ Xn there exists a surjective continuous map

ϕσ : D n → σ̄ ⊂ X

such that ϕσ | D̊ n : D̊ n → σ is a homeomorphism.

Definition 3.85. An element σ ∈ Xn is called an n-cell. The map ϕσ is called the characteris-
tic map of σ and X n is called the n-skeleton of (X, X). The map ϕσ S n−1=∂D n : S n−1 → X n−1
is called the attaching map of σ.

X2

b b
X0

X1

b b

Example 3.86. Consider X = S n for n ≥ 1. Then the choice

X0 = {{e1 }},
Xn = {σ n = S n \ {e1 }},
Xm = ∅, otherwise,

turns the n-sphere into a CW-complex.

131
3 Homology Theory

e1
Dn b
Sn

ϕσ n

The attaching map to the n-cell is the constant map S n−1 → {e1 }. We have

X 0 = X 1 = X 2 = . . . = X n−1 = {e1 },
S n = X n = X n+1 = . . .

Example 3.87. If a space X has the structure of a CW-complex there are in general many
different ways to write X as a CW-complex, i.e., there are many different X for the same X . Let
us look again at X = S n . We start with the case n = 0. Here the CW-structure is unique:


 {{e1 }, {−e1 }},
Xm = 
m=0
 ∅,
 m > 0.

For n > 0 we use that S n−1 ⊂ S n and define recursively




n−1

 XmS ,
 n n
m ≤ n−1
:= 
n
XmS 


{ D̊+ , D̊− }, m=n
 ∅,
 m>n

Sn
◦+
Dn

S n−1

◦−
Dn

132
3.10 CW-complexes

Therefore we have in this case

X 0 = S0, X 1 = S1, . . . , X n = S n .

Example 3.88. Real projective space RPn . We define the real projective space as

RPn = {1-dimensional real vector subspace of Rn+1 } = Rn+1 \ {0}/ ∼

where
x = (x 0, . . . , x n ) ∼ y = (y0, . . . , yn )
iff there exists a t , 0 such that x = t y. We consider the canonical projection map

π : Rn+1 \ {0} → RPn, x 7→ [x 0, . . . , x n ].

The restriction ψ := π S n : S n → RPn is continuous and surjective. Thus RPn is compact.


Clearly ψ(x) = ψ(y) iff x = ±y. We set

σk := [x 0, . . . , x n ] ∈ RPn | x k+1 = . . . = x n = 0, x k , 0 .

Then

σ̄k = [x 0, . . . , x n ] ∈ RPn | x k+1 = . . . = x n = 0 ≈ RPk .
For the characteristic map ϕk : D k → σ̄k we can take
q
ξ 7→ [ξ1, . . . , ξ k , 1 − |ξ | 2, 0, . . . , 0] .

It is clear that the map ϕk is continuous. Now we check that it is also surjective. Let [x ′, x k , 0] ∈
σ̄k where x ′ = (x 0, . . . , x k−1 ). Without loss of generality we assume that x k ≥ 0. Set

ξ := √ ′ 2x 2
∈ D k . We compute
| x | + | xk |

" s #
x′ |x ′ | 2
ϕk (ξ) = p , 1− ′ 2 ,0
|x ′ | 2 + |x k | 2 |x | + |x k | 2
" s #
x′ |x k | 2
= p , ,0
|x ′ | 2 + |x k | 2 |x ′ | 2 + |x k | 2
= [x ′, |x k |, 0]
= [x ′, x k , 0] .

Next we show that ϕk | D̊ k is injective. Let ϕk (ξ) = ϕk (η) for ξ, η ∈ D̊ k . This leads to the two
equations: q q
ξ = tη and 1 − |ξ | 2 = t 1 − |η| 2
for some t , 0. Squaring and adding both equations we find

|ξ | 2 + 1 − |ξ | 2 = t 2 |η| 2 + t 2 (1 − |η| 2 )

133
3 Homology Theory

p
leading
p to t 2 = 1 and consequently t = ±1. Since |ξ |, |η| < 1 it follows that 1 − |ξ | 2 > 0 and
1 − |η| 2 > 0 and therefore t > 0. Hence t = 1 and thus ξ = η.
The map ϕk : D k → σ̄k is closed, therefore the restriction

ϕk D̊ k : D̊ → σk

is bijective, continuous and closed, hence a homeomorphism. We find:

X 0 = {point} ⊂ |{z}
X 1 ⊂ |{z}
X 2 ⊂ · · · ⊂ |{z}
Xn .
≈RP1 ≈RP2 =RP n

Finally let us discuss the gluing map. For ξ ∈ ∂D k = S k−1 , i.e. |ξ | = 1, the gluing map is given
by ϕk (ξ) = [ξ, 0, 0] and hence ϕk = ψ : S k−1 → X k−1 ≈ RPk−1 .

Example 3.89. Complex projective space CPn . For the complex projective space the discussion
is analogous to the real case with complex parameters instead of real parameters,

CPn = {1-dimensional complex vector subspace of Cn+1 } = Cn+1 \ {0}/ ∼

with x ∼ y iff x = t y for some t ∈ C, t , 0. We find


 1,
|Xm | = 
m even and 0 ≤ m ≤ 2n
 0,
 otherwise
and
X 0 = {point} = X 1 ⊂ |{z}
X 2 = X 3 ⊂ |{z}
X 4 = X 5 ⊂ · · · ⊂ X 2n−1 = |{z}
X 2n .
≈CP1 ≈CP2 =CP n

Example 3.90. Quaternionic projective space HPn . Similarly, for the quaternionic projective
space

HPn = {1-dimensional quaternionic vector subspace of Hn+1 } = Hn+1 \ {0}/ ∼

with x ∼ y iff x = t y for some t ∈ H, t , 0, we find that


 1,
|Xm | = 
m divisible by 4 and 0 ≤ m ≤ 4n,
 0,
 otherwise,

and
X 0 = {point} = X 1 = X 2 = X 3 ⊂ |{z}
X 4 = . . . ⊂ X 4n−3 = · · · = |{z}
X 4n .
≈HP1 =HP n

Remark 3.91. Every compact differentiable manifold can be triangulated and is consequently a
finite CW-complex.

134
3.11 Homology of CW-complexes

3.11 Homology of CW-complexes


Throughout this section we assume that (X, X) is a finite CW-complex. Our goal is to prove
Theorem 3.100 which will provide us with an efficient way to compute the homology of X .

Lemma 3.92. The map

L ⊕σ∈Xn H i (ϕ σ )
Hi (D n, S n−1 ) / Hi (X n, X n−1 )
σ ∈Xn

is an isomorphism.

Once we have this lemma, Theorem 3.16 implies

Corollary 3.93. We have the isomorphisms


 R | Xn |,
Hi (X n, X n−1 )  
for i = n
 0,
 otherwise

Proof of Lemma 3.92. We set Ḋ n := D n \ {0} and


[
Ẋ n := X n−1 ∪ ϕσ ( Ḋ n ) = X n \ {ϕσ (0) | σ ∈ Xn }.
σ ∈Xn
x
The inclusion S n−1 ֒→ Ḋ n is a homotopy equivalence with homotopy inverse x 7→ |x | .

Dn

We define a a a
Y n := D n, Y n−1 := S n−1, Ẏ n := Ḋ n
σ ∈Xn σ ∈Xn σ ∈Xn

The inclusions yield homotopy equivalences Y n−1 → Ẏ n and X n−1 → Ẋ n .

135
3 Homology Theory

X2
b

X1

Now consider:
 Φ:=∪σ∈Xn ϕ σ
(Y n, Y n−1 )  / (Y n, Ẏ n ) / (X n, Ẋ n ) o ? _ (X n, X n−1 )

Both inclusions induce isomorphisms on Hi . Due to the diagram

Hi (Y n−1 ) / Hi (Y n ) / Hi (Y n, Y n−1 ) / Hi−1 (Y n−1 ) / Hi−1 (Y n )

 =  =
    
Hi (Ẏ n ) / Hi (Y n ) / Hi (Y n , Ẏ n ) / Hi−1 (Ẏ n ) / Hi−1 (Y n )

and the Five Lemma the map Hi (Y n, Y n−1 ) → Hi (Y n, Ẏ n ) is also an isomorphism. Similarly,
we get that the inclusion (X n, X n−1 ) ֒→ (X n, Ẋ n ) induces an isomorphism Hi (X n, X n−1 ) →
Hi (X n, Ẋ n ). The inclusions

(Y n \ Y n−1, Ẏ n \ Y n−1 ) ֒→ (Y n, Ẏ n ),
(X n \ X n−1, Ẋ n \ X n−1 ) ֒→ (X n, Ẋ n ),

induce isomorphisms on homology by the excision axiom. In addition, we have

(Y n \ Y n−1, Ẏ n \ Y n−1 ) ≈ (X n \ X n−1, Ẋ n \ X n−1 )

Hence we find
M
Hi (D n, S n−1 )  Hi (Y n, Y n−1 )
σ ∈Xn
 Hi (Y n, Ẏ n )
 Hi (Y n \ Y n−1, Ẏ n \ Y n−1 )
 Hi (X n \ X n−1, Ẋ n \ X n−1 )
 Hi (X n, Ẋ n )
 Hi (X n, X n−1 ) . 

Set K n (X, X) := Hn (X n, X n−1 )  R | Xn | . We define a homomorphism

∂n : K n (X, X) → K n−1 (X, X)

136
3.11 Homology of CW-complexes

by the following diagram:

/ Hn (X n, X n−1 ) ∂ / Hn−1 (X n−1 ) / Hn−1 (X n ) / ···


···
= ∂n
 *
··· / Hn−1 (X n−2 ) / Hn−1 (X n−1 ) / Hn−1 (X n−1, X n−2 ) / ···

Lemma 3.94. The sequence of groups K n (X, X) together with ∂n forms a complex.

The pair (K∗ (X, X), ∂∗ ) is called the cellular complex of (X, X).

Proof. The diagram


Hn (X n, X n−1 ) / Hn−1 (X n−1 )

= ∂n
 *

Hn−1 (X n−1 ) / Hn−1 (X n−1, X n−2 ) / Hn−1 (X n−2 )
2
=0 = ∂n−1
 *
Hn−1 (X n−2 ) / Hn−2 (X n−2, X n−3 )

shows ∂n−1 ◦ ∂n = 0 which is the claim. 

Lemma 3.95. For p ≥ q ≥ n or n > p ≥ q we have Hn (X p, X q ) = 0.

Proof. The proof is by induction on p−q. The assertion is certainly true for p−q = 0. To analyze
the situation p − q > 0 we look at the exact homology sequence for the triple (X p, X q+1, X q ):

ind. hyp.
Hn (X q+1, X q ) / Hn (X p, X q ) / Hn (X p, X q+1 ) = 0.

Since either q ≥ n or q < p < n we have n , q + 1. Thus Hn (X q+1, X q ) = 0 by Lemma 3.92


and hence Hn (X p , X q ) = 0. 

Corollary 3.96. For n > p we have that Hn (X p ) = 0.

Proof. Lemma 3.95 with q = 0 says Hn (X p, X 0 ) = 0. The claim now follows from the exact
sequence
0 = Hn (X 0 ) −→ Hn (X p ) −→ Hn (X p, X 0 ) = 0.

137
3 Homology Theory

Corollary 3.97. For q ≥ n we have Hn (X, X q ) = 0.

Proof. Choose p ≥ q so large that X p = X and use Lemma 3.95. 

Corollary 3.98. For r > n the inclusion X r ֒→ X induces an isomorphism

Hn (X )  Hn (X r ) .

Proof. The assertion follows from Corollary 3.97 and the exact sequence

0 = Hn+1 (X, X r ) −→ Hn (X r ) −→ Hn (X ) −→ Hn (X, X r ) = 0.

Lemma 3.99. For r > n and r ≥ q the inclusion induces an isomorphism

Hn (X, X q )  Hn (X r , X q ) .

Proof. Since r ≥ n + 1, Corollary 3.97 gives us Hn+1 (X, X r ) = Hn (X, X r ) = 0. The assertion
now follows from the exact homology sequence of the triple (X, X r , X q ):

Hn+1 (X, X r ) −→ Hn (X r , X q ) −→ Hn (X, X q ) −→ Hn (X, X r ).

Theorem 3.100. We have the following isomorphism:

Hn K∗ (X, X)  Hn (X )

138
3.11 Homology of CW-complexes

Proof. Consider the commutative diagram with exact columns and rows

Hn+1 (X n+1, X n
❘ ) H (X n−2 ) = 0
n−1
❘❘❘
❘❘❘∂n
∂ ❘❘❘
❘❘❘
 ) 
i∗
0 = Hn (X n−1 ) / Hn (X ) n / Hn (X n, X n−1 )
❘❘❘
/ Hn−1 (X n−1 )
❘❘❘ ∂n−1
❘❘❘
❘❘❘
 ❘) 
Hn (X n+1 ) Hn−1 (X n−1, X n−2 )


Hn (X n+1, X n ) = 0

Now we compute

Hn (X )  Hn (X n+1 )
Hn (X n )

∂Hn+1 (X n+1, X n )
i∗ Hn (X n )

i ∗ ∂Hn+1 (X n+1, X n )
ker(Hn (X n, X n−1 ) → Hn−1 (X n−1 ))

∂n Hn+1 (X n+1, X n )
ker(∂n−1 )

∂n Hn+1 (X n+1, X n )
= Hn K∗ (X, X)

and the theorem is proved. 

In the following examples we set α n := |Xn |.

Example 3.101. We consider X = S n for n ≥ 2. The CW-decomposition of S n from Exam-


ple 3.86 has
α 0 = α n = 1, α j = 0 otherwise.
Hence we have to compute the homology of the complex

R ←− 0 ←− · · · ←− 0 ←− R ←− 0 ←− · · ·

Since all arrows are 0 the homology coincides with the complex, hence


 R, j = 0 or n,
H j (S n )  
 0, otherwise,

which confirms our earlier findings.

139
3 Homology Theory

Example 3.102. Now look at X = CPn . Then we have for the CW-decomposition from Exam-
ple 3.89
α 0 = α 2 = . . . = α 2n = 1, α j = 0 otherwise.
Again all arrows in the complex

R ←− 0 ←− · · · ←− 0 ←− R ←− 0 ←− · · ·

must be zero, hence the homology is given by



 R, j = 0, 2, . . . , 2n,
H j (CPn )  
 0, otherwise.

Example 3.103. Consider X = HPn . For the CW-decomposition from Example 3.90 we have

α 0 = α 4 = α 8 = . . . = α 4n = 1, α j = 0 otherwise

and all arrows in

0 ←− R ←− 0 ←− 0 ←− 0 ←− R ←− · · · ←− 0 ←− R ←− 0 ←− · · ·

must be zero. We find for the homology



 R, j = 0, 4, 8, . . . , 4n,
H j (HPn )  
 0, otherwise.

3.12 Betti numbers and the Euler number


Throughout this section let R be a field.

Definition 3.104. The dimension b j (X ; R) := dim R H j (X ; R) is called j-th Betti number of


the space X (over the field R).

Now let (X, X) be a finite CW-complex. Denote the number of j-cells in (X, X) by α j . Then
we get the following estimate for the Betti numbers:

b j (X ; R) = dim R H j (X ; R)
ker(∂ j : K j (X, X) → K j−1 (X, X))
= dim R
im(∂ j+1 : K j+1 (X, X) → K j (X, X))
≤ dim R ker(∂ j : K j (X, X) → K j−1 (X, X))
≤ dim R K j (X, X)
= αj.

We conclude that b j (X ; R) ≤ α j . In particular, the Betti numbers are finite, b j (X ; R) < ∞.

140
3.12 Betti numbers and the Euler number

Definition 3.105. For a finite CW-complex (X, X) we call



X
χ(X, X) = (−1) i α i
i=0

the Euler number or Euler-Poincaré characteristic.

Note that the sum in this definition is finite. It ends at the highest dimension occuring in the cell
decomposition.

Proposition 3.106. We have the following relation between Euler and Betti numbers:

X
χ(X, X) = (−1) i bi (X ; R).
i=0

In particular, the Euler number does not depend on the cell decomposition because the Betti
numbers don’t. On the other hand, the Euler number does not depend on the coefficient field
because the α i ’s don’t. We will henceforth write χ(X ) instead of χ(X, X).
In order to prove the proposition we use the following

Lemma 3.107. Let


d1 d2 dn
0o V0 o V1 o ··· o Vn o 0o 0o ···

be a complex of finite-dimensional R-vector spaces. Then



X ∞
X
(−1) j dim H j V∗ = (−1) j dim Vj .
j=0 j=0

Proof of Lemma 3.107. To show the lemma we compute


X X
(−1) j dim Vj = (−1) j (dim(d j Vj ) + dim ker(d j ))
j j
X
= (−1) j (dim ker(d j ) − dim(d j+1 Vj+1 ))
j
!
X ker(d j )
= (−1) j dim
j
d j+1 Vj+1

141
3 Homology Theory

X
= (−1) j dim H j V∗ . 
j

Proof of Proposition 3.106. The proposition follows from Lemma 3.107 with Vj = K j (X, X) 
Rα j . 

Example 3.108. For X = S n we have


b0 = bn = 1, b j = 0 otherwise
and therefore

 2, n even
χ(S n ) = 
 0, n odd .

The case n = 2 contains Euler’s classical formula for polyhedra as a special case. It says that for
the alternating sum of the number of vertices, edges and faces of a polyhedron is always equal to
2. In particular, for platonic solids we have the following list:

α0 α1 α2
Tetrahedron 4 6 4
Cube 8 12 6
Octahedron 6 12 8
Dodecahedron 20 30 12
Icosahedron 12 30 20

Examples 3.109. We compute the Euler numbers of the projective spaces.


1.) For X = CPn we have b0 = b2 = . . . = b2n = 1 and b j = 0 otherwise. Hence χ(CPn ) = n+1.
2.) For X = HPn we have b0 = b4 = . . . = b4n = 1 and b j = 0 otherwise. Hence χ(HPn ) = n+1.
3.) In the case of X = RPn we do not know the Betti numbers yet. So we use Proposition 3.106
to compute the Euler number. For the cell decomposition described in Example 3.88 we
have
α 0 = . . . = α n = 1, α j = 0 otherwise.

142
3.13 Incidence numbers

Hence

 1,
χ(RPn ) = 
n even
 0,
 n odd

3.13 Incidence numbers


We return to a general commutative ring R with unit 1. Throughout this section let (X, X) be
a finite CW-complex. In order to compute the cellular homology of (X, X) we need a better
understanding of the homomorphism ∂n+1 : K n+1 (X, X) → K n (X, X). In the case of n = 0 we
find
∂1
K1 (X, X) / K0 (X, X)


H1 (X 1, X 0 ) / H0 (X 0 )


L L
H1 (ϕτ )(H1 (D 1, S0 )) R
τ ∈X1 σ ∈X0

Hence ∂1 is given by the (α 1 × α 0 )-matrix (∂στ ) where τ ∈ X1 and σ ∈ X0 . The entries ∂στ ∈ R
of this matrix are easily computed. Namely, recall that D 1 = [−1, 1] and that a generator of
H1 (D 1, S0 ) is represented by c : [0, 1] → [−1, 1], t 7→ 2t − 1. Then

∂1 H1 (ϕτ )([c]) = ∂[ϕτ ◦ c] = ϕτ (c(1)) − ϕτ (c(0)) = ϕτ (1) − ϕτ (−1).

Thus if ϕτ (−1) = ϕτ (1) then ∂στ = 0 for all σ. If ϕτ (−1) , ϕτ (1) then






1, for σ = ϕτ (1),

∂σ =  −1,
τ

for σ = ϕτ (−1),

 0,
 otherwise.

Example 3.110. We compute the homology of the following CW-complex consisting of two
0-cells and three 1-cells:
τ2
τ1 σ1 b b σ2
τ3

Clearly, ∂στ11 = ∂στ12 = 0. Depending on how the characteristic maps parametrize the 1-cells τ2
and τ3 we get
∂στ21 = ∂στ31 = 1 and ∂στ22 = ∂στ32 = −1,

143
3 Homology Theory

or possibly different signs which will not affect the homology however. Thus the cellular complex
is
*.0 1 1+
/
o o 2 ,o 0 −1 −1- 3 o
··· 0 R R 0o ···
The image of the matrix is {(x, −x) | x ∈ R} = R · (1, −1) and hence
H0 (X ; R)  R2/R · (1, −1)  R
and !
0 1 1
H1 (X ; R)  ker = {(x, y, −y) | x, y ∈ R}  R2 .
0 −1 −1

Returning to our general CW-complex (X, X) we find for n ≥ 1:


∂n+1
K n+1 (X, X) / K n (X, X)

Hn+1 (X n+1, X n ) Hn (X n, X n−1 )

L L
Hn+1 (ϕτ )(Hn+1 (D n+1, S n )) Hn (ϕσ )(Hn (D n, S n−1 ))
τ ∈Xn+1 σ ∈Xn
 
 τ)
(∂σ 
Rα n+1 / Rα n

Hence ∂n+1 is given by the (α n+1 × α n )-matrix (∂στ ) where τ ∈ Xn+1 and σ ∈ Xn . The entries
(∂στ ) ∈ R of this matrix are called the incidence numbers. We want to see how we can compute
them.
Fix σ ∈ Xn , τ ∈ Xn+1 , and p ∈ D̊ n . From the commutative diagram
∂n+1

+ prσ

Hn+1 (X n+1 n
O ,X )
/ Hn (X n )
O
/ Hn (X n, X n−1 ) / Hn (ϕσ )Hn (D n, S n−1 )
O
H n+1 (ϕ τ ) H n (ϕ τ | S n ) 
−1 ◦ϕ :ϕ −1 (σ )→D n ⊂S n )
deg p (ϕ σ
∂ τ τ
Hn+1 (D n+1, S n ) 
/ Hn (S n ) / Hn (S n )  Hn (D n, S n−1 )


Hn (S n , S n \ ϕτ−1 (ϕσ (p)))
O


Hn (ϕτ−1 (σ), ϕτ−1 (σ) \ ϕτ−1 (ϕσ (p))) 

 H n (ϕ τ )

Hn (σ, σ \ {ϕσ (p)})
−1 )
 H n (ϕ σ
 

Hn ( D̊ n, D̊ n \ {p}) / Hn (D n, D n \ {p})

144
3.13 Incidence numbers

we conclude that
∂στ = deg p (ϕσ
−1
◦ ϕτ : ϕτ−1 (σ) → D n ⊂ S n ).
We used the canonical isomorphism Hn (S n )  Hn (D n, S n−1 ). We have intepreted the incidence
numbers as certain local mapping degrees. In applications they can often be computed by
counting preimages.

Example 3.111. Look at X = RPn . We want to compute the homology of


∂1 ∂2 ∂3 ∂n
0o Ro Ro Ro ··· o Ro 0o ···
The operator ∂ j+1 is given by the (1 × 1)-matrix (∂στ ) with σ ∈ Xj and τ ∈ Xj+1 . The gluing map
ϕτ |S j : S j → X j ≈ RP j is given by the canonical projection. The image point ϕσ (p) ∈ X j has
exactly two preimages under ϕτ which we call x, −x ∈ S j . Put Φ := ϕσ −1 ◦ ϕ : ϕ−1 (σ) → D j .
τ τ
From the additivity of the local degree we obtain
deg p (Φ : ϕτ−1 (σ) → D j ) = deg p (Φ U (x) : U (x) → D j ) + deg p (Φ U (−x) : U (−x) → D j )
where U (x) is a small neighborhood of x. W.l.o.g. we assume U (−x) = −U (x). Since Φ U (x) is
a homeomorphism onto its image we have
deg p (Φ U (x) : U (x) → D j ) = ±1 =: ε .
Let a : S j → S j be the antipodal map. Then Φ = Φ ◦ a and hence
deg p (Φ U (−x) : U (−x) → D j ) = deg p (Φ ◦ a : U (−x) → D j )
= deg p (Φ : U (x) → D n ) · deg(a)
= ε · (−1) j+1 .
We conclude that ∂ j+1 = ε · (1 + (−1) j+1 ). For n even we obtain the complex
0 ±2 0 ±2
0o Ro Ro ··· o Ro Ro 0
whereas for n odd we find
0 ±2 0
0o Ro Ro ··· o Ro Ro 0.
For R = Z/2Z all arrows are zero so that

 Z/2Z,
H j (RPn ; Z/2Z) = 
j = 0, . . . , n,
 0,
 otherwise.
For R = Z the homology is computed to




Z, j = 0,







Z/2Z, j = 1, 3, n − 1 or n − 2 resp.,

H j (RP ; Z) =  Z,
n

j = n odd,






0, j = n even,

 0,
 otherwise.

145
3 Homology Theory

3.14 Homotopy versus homology


The final goal of this chapter is to compare homotopy groups and homology groups. We start by
examining the fundamental group of CW-complexes.

Proposition 3.112. Let (X, X) be a finite CW-complex and let x 0 ∈ X 0 . Then the inclusion
map j : X 2 ֒→ X induces an isomorphism

j# : π1 (X 2 ; x 0 ) → π1 (X ; x 0 ) .

Proof. We have to show that attaching a k−cell for k ≥ 3 does not alter π1 in the sense that the
inclusion induces an isomorphism on π1 . We assume that X 2 is path-connected, since otherwise
we may replace X 2 by the path-component that contains x 0 .
Let Y be a path-connected finite CW-complex and let Ỹ be obtained from Y by attaching a k-cell.
More precisely, Ỹ = Y ∪ϕ D k = Y ⊔ D k / ∼ where x ∼ ϕ(x) for all x ∈ S k−1 . Here ϕ : S k−1 → Y
is a continuous map. We have to show that the inclusion map induces an isomorphism

j# : π1 (Y ; x 0 ) → π(Ỹ ; x 0 )

if k ≥ 3. Let D k ( 21 ) ⊂ D̊ k be the closed k-dimensional subball of radius 21 . Cover Ỹ by the two


open subsets U1, U2 where

U1 = D̊ k , U2 = Ỹ \ D k ( 21 ) ≃ Y

Y U2

146
3.14 Homotopy versus homology

We then find U1 ∩ U2 = D̊ k \ D k ( 21 ) ≃ S k−1 . To apply the Seifert-van-Kampen Theorem 2.68


we calculate

π1 (U1 ) = {1},
π1 (U2 )  π1 (Y ),
π1 (U1 ∩ U2 )  π1 (S k−1 ) = {1}, (here we use k − 1 ≥ 2)

and we find
π1 (U1 ) ⋆ π1 (U2 )
π1 (Ỹ )   π1 (Y )
im π1 (U1 ∩ U2 )
the isomorphisms being induced by inclusions. 

Example 3.113. Consider complex-projective space X = CPn . We use the cell decomposition
from Example 3.89:

X 0 = {point} = X 1 ⊂ |{z}
X 2 = X 3 ⊂ |{z}
X 4 = X 5 ⊂ . . . ⊂ X 2n−1 ⊂ |{z}
X 2n
≈CP1 ≈CP2 =CP n

We are now able to calculate the fundamental group

π1 (CPn )  π1 (X 2 )  π1 (CP1 ) = π1 (S2 ) = {1}.

Hence complex-projective space CPn is simply connected.

Example 3.114. Similarly, for X = HPn we use the cell decomposition from Example 3.90:

X 0 = {point} = X 1 = X 2 = X 3 ⊂ |{z}
X 4 = . . . ⊂ X 4n−3 = . . . ⊂ |{z}
X 4n
≈HP1 =HP n

For the fundamental group we find

π1 (HPn ) = π1 (X 2 ) = π1 ({point}) = {1}.

Thus HPn is also simply connected.

Remark 3.115. Proposition 3.112 can be generalized as follows: For a finite CW-complex the
inclusion map j : X n+1 ֒→ X always induces an isomorphism j# : πn (X n+1 ; x 0 ) → πn (X ; x 0 )
where x 0 ∈ X 0 .

Now we relate homotopy and homology groups. Recall that for the n-dimensional cube
W n = [0, 1]n we have (W n, ∂W n ) ≈ (D n, S n−1 ). Fix a generator α n ∈ Hn (W n, ∂W n ; Z) 
Hn (D n, S n−1 ; Z)  Z. The elements of πn (X ; x 0 ) are homotopy classes relative to ∂W n of
maps f : W n → X with f (∂W n ) = {x 0 }. Then Hn ( f )(α n ) ∈ Hn (X, {x 0 }; Z). The long exact
homology sequence of the pair (X, {x 0 }) yields for n ≥ 1

/ Hn (X ; Z) / Hn (X, {x 0 }; Z) 0 / Hn−1 ({x 0 }; Z).


0 = Hn ({x 0 }; Z)

147
3 Homology Theory

Namely, if n ≥ 2 then Hn−1 ({x 0 }; Z) = 0. For n = 1 the arrow emanating from H0 ({x 0 }; Z)
is injective so that the incoming arrow must again be zero. In either case the inclusion

j : (X, ∅) ֒→ (X, {x 0 }) induces an isomorphism Hn ( j) : Hn (X ; Z) −→ Hn (X, {x 0 }; Z). Now set

h([ f ]) := Hn ( j) −1 Hn ( f )(α n ) ∈ Hn (X ; Z) .

Due to homotopy invariance the expression h([ f ]) only depends on the homotopy class of the
map f . Hence we have constructed a well-defined map

h : πn (X ; x 0 ) → Hn (X ; Z), n ≥ 1.

Proposition 3.116. The map h : πn (X ; x 0 ) → Hn (X ; Z) is a homomorphism.

Proof. Consider the map s1 : W1n = [0, 12 ] × [0, 1]n−1 → W n given by

(x 1, . . . , x n ) 7→ (2x 1, x 2, . . . , x n )

and the map s2 : W2n = [ 21 , 1] × [0, 1]n−1 → W n .

(x 1, . . . , x n ) 7→ (2x 1 − 1, x 2, . . . , x n ) .

Wn W1n W2n W n
s1

s2

For [ f 1 ], [ f 2 ] ∈ πn (X ; x 0 ) we have [ f 1 ] + [ f 2 ] = [g] with


 f 1 (s1 (x)),
g(x) = 
x 1 ≤ 12 ,
 f 2 (s2 (x)),
 x 1 ≥ 12 .

We represent α n by c1 + c2 ∈ Sn (W n ; Z), where cν ∈ S n (Wνn ; Z) and cν represents the generator


Hn (sν ) −1 (α n ) of Hn (Wνn, ∂Wνn ; Z).

148
3.14 Homotopy versus homology

n=1 n=2

σ1 σ3 c1 = σ1 + σ2
c2 = σ3 + σ4
σ2 σ4

c1 c2

Now the proposition follows from

h([ f 1 ] · [ f 2 ]) = h([g])
= Hn ( j) −1 Hn (g)(α n )
= Hn ( j) −1 Hn (g)([c1 + c2 ])
= Hn ( j) −1 Hn (g)([c1 ]) + Hn ( j) −1 Hn (g)([c2 ])
= Hn ( j) −1 Hn ( f 1 ◦ s1 )([c1 ]) + Hn ( j) −1 Hn ( f 2 ◦ s2 )([c2 ])
= Hn ( j) −1 Hn ( f 1 )Hn (s1 )([c1 ]) + Hn ( j) −1 Hn ( f 2 )Hn (s2 )([c1 ])
= Hn ( j) −1 Hn ( f 1 )(α n ) + Hn ( j) −1 Hn ( f 2 )(α n )
= h( f 1 ) + h( f 2 ). 

Definition 3.117. The map h : πn (X ; x 0 ) → Hn (X ; Z) is called Hurewicz homomorphism.

Proposition 3.118. The Hurewicz homomorphism h is natural, i.e., for every f ∈ C (X, Y )
with f (x 0 ) = y0 the following diagram commutes:

h / Hn (X ; Z)
πn (X ; x 0 )
f# Hn ( f )
 h

πn (Y ; y0 ) / Hn (Y ; Z)

Proof. The inclusion maps

j X : (X, ∅) ֒→ (X, {x 0 })
jY : (Y, ∅) ֒→ (Y, {y0 })

149
3 Homology Theory

satisfy

( jY ◦ f )(x) = jY ( f (x)) = ( f (x), y0 )


( f ◦ j X )(x) = f ((x, x 0 )) = ( f (x), f (x 0 )) = ( f (x), y0 )

and therefore jY ◦ f = f ◦ j X . On the level of homology groups we find

Hn ( jY ) ◦ Hn ( f ) = Hn ( f ) ◦ Hn ( j X )

and consequently
Hn ( f ) ◦ Hn ( j X ) −1 = Hn ( jY ) −1 ◦ Hn ( f ) .
Thus

(Hn ( f ) ◦ h)([σ]) = Hn ( f ) ◦ Hn ( j X ) −1 ◦ Hn (σ)(α n )


= Hn ( jY ) −1 ◦ Hn ( f ) ◦ Hn (σ)(α n )
= Hn ( jY ) −1 ◦ Hn ( f ◦ σ)(α n )
= h([ f ◦ σ])
= (h ◦ f # )([σ])

as claimed. 

Theorem 3.119 (Hurewicz). Let X be a topological space, x 0 ∈ X and n ≥ 2. Assume that

π0 (X ; x 0 ) = π1 (X ; x 0 ) = . . . = πn−1 (X ; x 0 ) = {1} .

We then have
H1 (X ; Z) = . . . = Hn−1 (X ; Z) = 0
and the map
h : πn (X ; Z) → Hn (X ; Z)
is an isomorphism.

For a proof of this theorem see e.g. [2, Sec. 4.2].

Remark 3.120. For n = 1 this theorem cannot be true as it stands because H1 (X ; Z) is always
abelian while π1 (X ; x 0 ) is not in general. However, h induces a homomorphism

h̄ : π1 (X ; x 0 )/[π 1 (X ;x0 ),π 1 (X ;x0 )] → H1 (X ; Z).

Already Poincaré showed that if π0 (X, x 0 ) = {1} (i.e., X is path-connected) then the map h̄ is an
isomorphism.

150
3.14 Homotopy versus homology

Example 3.121. Consider X = S n for n ≥ 2. We already know that S n is 1-connected. Now


apply the Hurewicz isomorphism with n = 2:

 Z,
π2 (S n )  H2 (S n ; Z) = 
n = 2,
 0,
 n ≥ 3.

For n ≥ 3 we find due to Hurewicz:



 Z,
π3 (S n )  H3 (S n ; Z) = 
n = 3,
 0,
 n ≥ 4.

By induction we then deduce that π1 (Sn ) = . . . = πn−1 (S n ) = 0 and πn (S n )  Z.

Example 3.122. We know that X = CPn is 1-connected. With the help of the Hurewicz
isomorphism we calculate
π2 (CPn )  H2 (CPn ; Z)  Z .

Example 3.123. We also know that X = HPn is 1-connected. We apply the Hurewicz isomor-
phism three times and we get

π2 (HPn )  H2 (HPn ; Z) = 0,
π3 (HPn )  H3 (HPn ; Z) = 0,
π4 (HPn )  H4 (HPn ; Z)  Z.

Example 3.124. Now let us analyze the space X = RPn for n ≥ 2. The map ψ : S n → RPn is a
two-fold covering and by Theorem 2.102
ψ♯
π1 (S n ) −−→ π1 (RPn ) → π0 ({p, q}) → π0 (S n )

is exact. Since S n is simply connected π1 (RPn ) → π0 ({p, q}) is an isomorphism of pointed


sets, thus π1 (RPn ) has exactly two elements. There is only one group with two elements, hence
π1 (RPn )  Z/2Z. By Corollary 2.105 ψ# : πk (S n ) → πk (RPn ) is an isomorphism for all k ≥ 2.
We conclude that π2 (RPn ) = . . . = πn−1 (RPn ) = 0 and πn (RPn )  Z.

Remark 3.125. Under the assumptions of the theorem of Hurewicz 3.119 not much can be
said about h : πk (X, x 0 ) → Hk (X ; Z) for k > n. For example, consider the Hopf fibration
S3 → S2 with fiber S1 . By Theorem 2.102 it induces an isomorphism π3 (S2 )  π3 (S3 )  Z.
But H3 (S2 ; Z) = 0 and hence h : π3 (S2 ) → H3 (S2 ; Z) is not injective.
On the other hand, for the 2-torus T 2 we have again by Corollary 2.105 π2 (T 2 )  π2 (R2 ) = 0
while one can compute H2 (T 2 ; Z)  Z. Thus h : π2 (T 2 ) → H2 (T 2 ; Z) is not surjective.

Remark 3.126. We have seen that Hk (S n ; Z) = 0 whenever k > n. But in general this is not
true for the higher homotopy groups of the sphere, e.g. π3 (S2 )  Z. The computation of πk (S n )
for k > n is a difficult problem and many of these groups are not known to date.

151
3 Homology Theory

3.15 Exercises
3.1. Let X be a topological space. Show:

a) If X is path-connected then
H0 (X ; R)  R

b) If X k , k ∈ K, are the path-components of X then


M
Hn (X ; R)  Hn (X k ; R)
k ∈K

3.2. Let Yk = {1, ..., k} be equipped with the discrete topology. Compute Hn (Yk ; R) without
using Exercise 3.1 directly with the help of the Eilenberg-Steenrod axioms.
Hint: Consider the pair (Yk , Yk−1 ).

3.3. Let X be a topological space. The augmented boundary operator is defined by

∂ # : S0 (X ; R) → R,

X  X
∂# α i σi = αi,

where σi ∈ C (∆0, X ).

a) Verify ∂ # ◦ ∂ = 0.

b) Compute the augmented homology

ker(∂ # : S0 (X ; R) → R)
H0# (X ; R) :=
im(∂ : S1 (X ; R) → S0 (X ; R))
for X = {point}.

3.4. a) Show that homeomorphisms σ : ∆n → D n represent generators of Hn (D n, S n−1 ; R).

b) Describe generators of Hn (S n ; Z). Make a drawing for n = 2.

3.5 (Topological invariance of the dimension). Let U ⊂ Rn and V ⊂ Rm be open and


nonempty. Show: If U and V are homeomorphic then n = m.
Hint: For p ∈ U and q ∈ V consider the pairs (U, U \ {p}) and (V, V \ {q}).

3.6. Let p be a complex polynomial without zeros on the unit circle S1 ⊂ C. Show: The degree
of the map
p(z)
p̂ : S1 → S1, p̂(z) = ,
|p(z)|
coincides with the number of zeros of p in the interior of the unit disk (counted with multiplicities).

152
3.15 Exercises

3.7 (Homotopy invariance of the local mapping degree). Let V ⊂ S n be open and F : V ×
S
[0, 1] → S n continuous. We put f t (x) := F (x, t). Let p ∈ S n such that t ∈[0,1] f t−1 (p) is
compact. Show:
deg p ( f 1 ) = deg p ( f 0 ).

3.8. Let (X, A) be a pair such that A is closed and a strong deformation retract of an open
neighborhood U. Show that Hn (X, A) = Hn (X/A) for n , 0.

3.9. Let Z = S1 × [0, 1] be the cylinder. Compute Hn (Z, S1 × {0} ∪ S1 × {1}) for all n. Sketch
generators of the nontrivial homology groups.
Hint: Use the homology sequence of the triple

(Z, S1 × {0} ∪ S1 × {1}, S1 × {0}).

3.10. Let (X, A) be a pair. Describe the 0th singular relative homology group H0 (X, A).

3.11 (Five lemma). Let the rows in the following commutative diagram of abelian groups be
exact:
f1 f2 f3 f4
A1 / A2 / A3 / A4 / A5

ϕ1 ϕ2 ϕ3 ϕ4 ϕ5
 g1  g2  g3  g4 
B1 / B2 / B3 / B4 / B5

Show that if ϕ1 , ϕ2 , ϕ4 , and ϕ5 are isomorphisms then so is ϕ3 .


Hint: ϕ3 is injective if ϕ1 is surjective and ϕ2 , ϕ4 are injective; ϕ3 is surjective if ϕ5 is injective
and ϕ2 , ϕ4 are surjective.

3.12. Suppose
f n+1 fn
· · · → G n+1 −−−−−−−−−−→ G n −−−−−−−−→ G n−1 → · · ·
is a long exact sequence of abelian groups and
f′ f′
′ n+1 ′ −−−−−n−−−→ ′
··· → G n+1 −−−−−−−−−−→ G n G n−1 → ···

is a subsequence, i.e., G n′ ⊂ G n and f n′ = f n |G′n . Prove that the subsequence is exact if and only
if the quotient sequence

· · · → G n+1 /G n+1 → G n /G n′ → G n−1 /G n−1

→ ···

is exact.

3.13. Let n ∈ N and m ∈ Z. Show that there exists f ∈ C (S n, S n ) with deg( f ) = m.

153
3 Homology Theory

3.14. Let f ∈ C (S n, S n ) with f (D+n ) ⊂ D+n and f (D−n ) ⊂ D−n . We identify S n−1 with D+n ∩ D−n
and therefore have f (S n−1 ) ⊂ S n−1 . Show

deg( f ) = deg( f |S n−1 ).

3.15. Show that H1 (R, Q; Z) is a free abelian group and find a basis as a Z-module.

3.16. Let M be an n-dimensional manifold, n ≥ 3. Let p ∈ M. Show that the inclusion map
M \ {p} ֒→ M induces an isomorphism

H j (M \ {p}; R)  H j (M; R)

for all j ∈ {1, ..., n − 2}.

3.17. a) Show that for disjoint closed subsets A, B ⊂ R2 we have

H1 (R2 \ (A ∪ B))  H1 (R2 \ A) ⊕ H1 (R2 \ B).

b) Let p1, ..., pn ∈ R2 be pairwise distinct. Compute H1 (R2 \ {p1, ..., pn }) and sketch generators.

3.18. Use the Mayer-Vietoris sequence to compute

a) the homology of the 2-torus;

b) the homology of surfaces Fg of genus g ≥ 2.

Sketch generators.

3.19. Let (X, X) be a finite CW-complex. Show that X is compact.

3.20. Describe a CW-decomposition for the surfaces of genus g ≥ 1.

3.21. Show:

a) Each nonempty CW-complex has at least one 0-cell.

b) Each CW-complex consisting of exactly two cells is homeomorphic to a sphere.

3.22. Let n ≥ 2 and k ≥ 1. Let X be the topological space obtained from k copies of S n by
identifying them all at one point. More formally,
k
[ 
X= { j} × S n ∼
j=1

where ( j, x) ∼ ( j ′, x ′ ) iff x = x ′ = e1 . Compute the homology of X .

154
3.15 Exercises

3.23. Find a CW-decomposition of the 2-torus with exactly one 0-cell, two 1-cells, and one
2-cell and use it to compute the homology.

155
Sources list for pictures

titlepage http://www.sxc.hu
p. 4 taken from [3]
p. 5 based on http://commons.wikimedia.org/wiki/File:Hilbert_curve.png
p. 6, 2. and 4. picture http://www.sxc.hu
p. 33 based on http://www.sxc.hu
p. 130 http://www.math.ohio-state.edu/∼fiedorow/

157
Bibliography
[1] Greenberg, M. J.; Harper, J. R.: Algebraic topology. A first course, Benjamin/Cummings
Publishing, 1981

[2] Hatcher, A.: Algebraic topology, Cambridge University Press, 2002

[3] Hilbert, D.: Ueber die stetige Abbildung einer Line auf ein Flächenstück, Math. Ann. 38
(1891), 459–460

[4] Lück, W.: Algebraische Topologie. Homologie und Mannigfaltigkeiten, Vieweg, 2004

[5] Ossa, E.: Topologie, Vieweg, 1992

[6] Peano, G.: Sur une courbe, qui remplit toute une aire plane, Math. Ann. 36 (1890), 157–160

[7] tom Dieck, T.: Topologie, Walter de Gruyter, 1991

[8] Warner, F. W.: Foundations of differentiable manifolds and Lie groups, Springer-Verlag,
New York, 1983

159
Index

Symbols h, Hurewicz homomorphism . . . . . . . . . . 148


Hn ( f ), map induced on homology . . . . . 83
a, antipodal map . . . . . . . . . . . . . . . . . . . . . . 95 Hn K∗ , n-th homology . . . . . . . . . . . . . . . . 106
Ā, closure of A . . . . . . . . . . . . . . . . . . . . . . . 88 Hn (ϕ∗ ), map induced on homology . . . 106
Å, interior of A . . . . . . . . . . . . . . . . . . . . . . . 88 Hn (X, A; R), relative n-th singular
b j (X ; R), Betti number . . . . . . . . . . . . . . . 140 homology . . . . . . . . . . . . . . . . . . . 84
Bn , barycenter . . . . . . . . . . . . . . . . . . . . . . . 117 Hn (X ; R), n-th singular homology . . . . . 82
Bn K∗ , n-boundaries . . . . . . . . . . . . . . . . . . 106 HPn , quaternionic projective space . . . . 134
Bn (X, A; R), relative singular n-boundaries K n (X, X), cellular complex . . . . . . . . . . . 136
84 N (S), normal subgroup generated by S . 41
Bn (X ; R), singular n-boundaries . . . . . . . 82 Ω (X ; x 0, x 1 ), set of paths . . . . . . . . . . . . . . 21
B(x, r), open ball of radius r centered at x 7 Ω (X ; x 0 ), set of loops . . . . . . . . . . . . . . . . . 21
χ(X, X), Euler number . . . . . . . . . . . . . . 141 P, prism operator . . . . . . . . . . . . . . . . . . . . 114
Cn , barycentric subdivision . . . . . . . . . . . 117 ϕσ , attaching map . . . . . . . . . . . . . . . . . . . 131
CPn , complex projective space . . . . 77, 134 π1 (X ; x 0 ), fundamental group . . . . . . . . . . 22
C X , cone over X . . . . . . . . . . . . . . . . . . . . . 14 πn (X ; x 0 ), n-th homotopy group . . . . . . . 58
C ((X, A), (Y, B)), set of continuous maps Q n , chain homotopy for subdivision map
between pairs of spaces . . . . . . . 84 119
C (X, Y ), set of continuous maps . . . . . . . . 17 RPn , real projective space . . . . . 73, 77, 133
deg, degree . . . . . . . . . . . . . . . . . . . . . . . 30, 97 Sdn , subdivision chain map . . . . . . . . . . . 119
deg p , local degree . . . . . . . . . . . . . . . . . . . 102 Skaff (∆n ), affine chains . . . . . . . . . . . . . . . . 118
D n , closed n-dimensional unit ball . . . . . . 6 Sn ( f ), map induced on singular chains . . 83
D±n , hemisphere . . . . . . . . . . . . . . . . . . . . . . 90 Sn (X, A; R), relative singular n-chains . . 84
∆n , standard simplex . . . . . . . . . . . . . . . . . . 79 Sn (X ; R), singular n-chains . . . . . . . . . . . . 80
∂, boundary of singular chain . . . . . . . . . . 81 SnU (X, A), small singular n-chains . . . . . 116
∂, connecting homomorphism for ΣX , suspension of X . . . . . . . . . . . . . . . . . . 14
homology groups . . . . . . . . 86, 107 σ (i) , i-th face of singular simplex . . . . . . . 81
∂, connecting homomorphism for Tni . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
homotopy groups . . . . . . . . . . . . . 67 TRn , standard topology of R n . . . . . . . . . . . . 7
¯ boundary of relative singular chain . . . 84
∂, TX , topology of X . . . . . . . . . . . . . . . . . . . . . 7
ε x0 , constant loop . . . . . . . . . . . . . . . . . . . . . 21 U ( f , p), winding number . . . . . . . . . . . . . . 75
Exp, covering map of S1 . . . . . . . . . . . . . . . 28 W n , n-dimensional cube . . . . . . . . . . . . . . . 13
f # , map induced on fundamental group . 24 X n , n-skeleton . . . . . . . . . . . . . . . . . . . . . . 131
Fg , surface of genus g . . . . . . . . . . . . . . . . . . 6 Z n K∗ , n-cycles . . . . . . . . . . . . . . . . . . . . . . 105
Fni , face map . . . . . . . . . . . . . . . . . . . . . . . . . 80 Z n (X, A; R), relative singular n-cycles . . 84
H, quaternions . . . . . . . . . . . . . . . . . . . . . . . . 99 Z n (X ; R), singular n-cycles . . . . . . . . . . . . 82

161
Index

≃, homotopic . . . . . . . . . . . . . . . . . . . . 17, 111 connected . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39


≃ A , homotopic relative to A. . . . . . . . . . . .17 1-connected . . . . . . . . . . . . . . . . . . . . . . . . . . 39
≈, homeomorphic . . . . . . . . . . . . . . . . . . . . . . 9 connecting homomorphism . . . . . . . 86, 107
G/H, homogeneous space . . . . . . . . . . . . . 64 continuous map . . . . . . . . . . . . . . . . . . . . . . . . 8
M#N, connected sum . . . . . . . . . . . . . . . . . 51 contractible . . . . . . . . . . . . . . . . . . . . . . . . . . 19
G1 ⋆ G2 , free product . . . . . . . . . . . . . . . . . 43 covering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
ω ⋆ η, concatenation . . . . . . . . . . . . . . . . . . 21 cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
A
D
Alexander horned sphere . . . . . . . . . . . . . 130
antipodal map . . . . . . . . . . . . . . . . . . . . . . . . 95 deformation retract . . . . . . . . . . . . . . . . . . . . 19
attaching map . . . . . . . . . . . . . . . . . . . . . . . 131 degree. . . . . . . . . . . . . . . . . . . . . . . . . . . .30, 97
augmented boundary operator . . . . . . . . . 152 dimension axiom . . . . . . . . . . . . . . . . . . . . . 87
augmented homology . . . . . . . . . . . . . . . . 152 discrete topology . . . . . . . . . . . . . . . . . . . . . . 8

B E

barycenter . . . . . . . . . . . . . . . . . . . . . . . . . . 117 Eilenberg-Steenrod axioms . . . . . . . . . . . . 87


base space . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 embedding . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Betti number . . . . . . . . . . . . . . . . . . . . . . . . 140 Euler number . . . . . . . . . . . . . . . . . . . . . . . 141
bidegree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 Euler’s formula for polyhedra . . . . . . . . . 142
Borsuk-Ulam theorem . . . . . . . . . . . . . . . . . 33 Euler-Poincaré characteristic . . . . . . . . . . 141
boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 exact sequence of complexes . . . . . . . . . . 106
boundary of a singular simplex . . . . . . . . . 81 exactness axiom . . . . . . . . . . . . . . . . . . . . . . 88
boundary operator . . . . . . . . . . . . . . . . . . . . 86 excision axiom . . . . . . . . . . . . . . . . . . . . . . . 88
augmented . . . . . . . . . . . . . . . . . . . . . . 152
F
Brouwer’s fixed point theorem . . . . 37 f., 92
face of a singular simplex . . . . . . . . . . . . . . 81
C fiber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
fiber bundle . . . . . . . . . . . . . . . . . . . . . . . . . . 63
cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
finite CW-complex . . . . . . . . . . . . . . . . . . . 130
cellular complex . . . . . . . . . . . . . . . . . . . . . 137
finitely presentable group . . . . . . . . . . . . . . 55
chain homotopy . . . . . . . . . . . . . . . . . . . . . 110
free product . . . . . . . . . . . . . . . . . . . . . . . . . . 42
chain map . . . . . . . . . . . . . . . . . . . . . . . . . . 106
fundamental group . . . . . . . . . . . . . . . . . . . . 22
characteristic map . . . . . . . . . . . . . . . . . . . 131
chord . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 H
closed subset . . . . . . . . . . . . . . . . . . . . . . . . . . 8
closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 ham-sandwich theorem . . . . . . . . . . . . . . . . 34
coarse topology . . . . . . . . . . . . . . . . . . . . . . . . 8 Hausdorff space . . . . . . . . . . . . . . . . . . . . . . 10
comb space . . . . . . . . . . . . . . . . . . . . . . . 19, 75 Heine-Borel theorem . . . . . . . . . . . . . . . . . . . 9
compact subset . . . . . . . . . . . . . . . . . . . . . . . . 9 Hilbert curve . . . . . . . . . . . . . . . . . . . . . . . . . . 4
complex . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 HLP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
complex projective space. . . . . . . . . .77, 134 homeomorphic . . . . . . . . . . . . . . . . . . . . . . . . 9
cone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 homeomorphism . . . . . . . . . . . . . . . . . . . . . . . 9

162
Index

homogeneous space . . . . . . . . . . . . . . . . . . . 64 open subset of Rn . . . . . . . . . . . . . . . . . . . . . . 7


homology . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 open subset of a topological space . . . . . . . 8
augmented . . . . . . . . . . . . . . . . . . . . . . 152
relative singular . . . . . . . . . . . . . . . . . . 84 P
singular . . . . . . . . . . . . . . . . . . . . . . . . . .82
pair of spaces. . . . . . . . . . . . . . . . . . . . . . . . . 84
homotopic chain maps . . . . . . . . . . . . . . . 111
path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
homotopic relative to A. . . . . . . . . . . . . . . .17
path-connected . . . . . . . . . . . . . . . . . . . . . . . 39
homotopy axiom . . . . . . . . . . . . . . . . . . . . . . 88
plane-filling curve . . . . . . . . . . . . . . . . . . . . . 4
homotopy equivalence . . . . . . . . . . . . 18, 112
pointed set . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
homotopy equivalent . . . . . . . . . . . . . . . . . . 18
presentation of a group . . . . . . . . . . . . . . . . 55
homotopy group . . . . . . . . . . . . . . . . . . . . . . 58
prism operator. . . . . . . . . . . . . . . . . . . . . . .114
homotopy lifting property . . . . . . . . . . . . . 61
projective space
homotopy relative to A . . . . . . . . . . . . . . . . 17
complex . . . . . . . . . . . . . . . . . . . . . 77, 134
Hopf fibration . . . . . . . . . . . . . . . . . . . . . . . . 77
quaternionic . . . . . . . . . . . . . . . . . . . . 134
Hurewicz homomorphism . . . . . . . . . . . . 149
real . . . . . . . . . . . . . . . . . . . . . 73, 77, 133
I
Q
incidence numbers . . . . . . . . . . . . . . . . . . . 144
induced topology of metric space . . . . . . . . 8 quaternionic projective space . . . . . . . . . 134
induced topology of subspace . . . . . . . . . . . 8 quaternions . . . . . . . . . . . . . . . . . . . . . . . . . . 99
interior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 R
L real projective space . . . . . . . . . . 73, 77, 133
lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 relative singular homology . . . . . . . . . . . . . 84
local degree . . . . . . . . . . . . . . . . . . . . . . . . . 102 retract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
long exact homology sequence of a pair of retraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
spaces . . . . . . . . . . . . . . . . . . . . . . . 88
S
long exact homotopy sequence of a Serre
fibration . . . . . . . . . . . . . . . . . . . . . 69 Seifert-van Kampen theorem . . . . . . . . . . . 47
loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 Serre fibration . . . . . . . . . . . . . . . . . . . . . . . . 62
simply-connected . . . . . . . . . . . . . . . . . . . . . 39
M
singular n-boundary. . . . . . . . . . . . . . . . . . . 82
Mayer-Vietoris sequence . . . . . . . . . . . . . 124 singular n-chain . . . . . . . . . . . . . . . . . . . . . . 80
Möbius strip . . . . . . . . . . . . . . . . . . . . . . . . . 76 singular n-cycle . . . . . . . . . . . . . . . . . . . . . . 82
singular n-simplex . . . . . . . . . . . . . . . . . . . . 80
N singular homology . . . . . . . . . . . . . . . . . . . . 82
skeleton . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
natural . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 small chain . . . . . . . . . . . . . . . . . . . . . . . . . 116
normal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40 small chain theorem . . . . . . . . . . . . . . . . . 116
normal subgroup generated by a set . . . . . 41 split exact sequence . . . . . . . . . . . . . . . . . . . 78
O standard simplex . . . . . . . . . . . . . . . . . . . . . . 79
standard topology of Rn . . . . . . . . . . . . . . . . 7
open (for the product topology) . . . . . . . . 13 strong deformation retract . . . . . . . . . . . . . 19

163
Index

subspace topology . . . . . . . . . . . . . . . . . . . . . 8 of product topology . . . . . . . . . . . . . . . 13


surface of genus g . . . . . . . . . . . . . . . . . . . . 52 of quotient topology . . . . . . . . . . . . . . 11
suspension . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
V
T
vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
total space . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
W
U
weak fibration . . . . . . . . . . . . . . . . . . . . . . . . 62
universal property winding number . . . . . . . . . . . . . . . . . . . . . . 75

164

You might also like